Next Article in Journal
Characterisation of Large-Sized REBaCuO Bulks for Application in Flux Modulation Machines
Previous Article in Journal
Mechanical and Fatigue Properties of Ti-6Al-4V Alloy Fabricated Using Binder Jetting Process and Subjected to Hot Isostatic Pressing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Nickel-Containing High-Entropy Alloys: From Fundamentals to Additive Manufacturing

by
Ashish Kumar Gupta
1,*,
Amit Choudhari
2,
Aditya Rane
1,
Abhishek Tiwari
3,*,
Prince Sharma
4,
Ashutosh Gupta
5,
Prathamesh Sapale
6,
Ravi Teja A. Tirumala
6,
Rajmohan Muthaiah
7 and
Abhishek Kumar
8
1
School of Mechanical and Aerospace Engineering, Oklahoma State University, Stillwater, OK 74078, USA
2
Department of Mechanical Engineering, Cleveland State University, Cleveland, OH 44115, USA
3
Department of Mechanical and Aerospace Engineering, Indian Institute of Technology Hyderabad, Sangareddy 502284, India
4
Department of Mechanical Engineering and Mechanics, Lehigh University, Bethlehem, PA 18015, USA
5
Department of Zoology, Dayanand Vedic College, Orai 285001, India
6
School of Chemical Engineering, Oklahoma State University, Stillwater, OK 74078, USA
7
School of Aerospace and Mechanical Engineering, University of Oklahoma, Norman, OK 73019, USA
8
J. Mike Walker ’66 Department of Mechanical Engineering, Texas A&M University, College Station, TX 77840, USA
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(15), 3826; https://doi.org/10.3390/ma17153826
Submission received: 11 June 2024 / Revised: 17 July 2024 / Accepted: 23 July 2024 / Published: 2 August 2024
(This article belongs to the Section Metals and Alloys)

Abstract

:
High-entropy alloys (HEAs) are recognized as a class of advanced materials with outstanding mechanical properties and corrosion resistance. Among these, nickel-based HEAs stand out for their impressive strength, ductility, and oxidation resistance. This review delves into the latest advancements in nickel-containing HEAs, covering their fundamental principles, alloy design strategies, and additive manufacturing techniques. We start by introducing HEAs and their unique properties, emphasizing the crucial role of nickel. This review examines the complex relationships between alloy composition, valence electron concentration (VEC), and the resulting crystal structures. This provides insights into design principles for achieving desired microstructures and mechanical properties. Additive manufacturing (AM) techniques like selective laser melting (SLM), electron beam melting (EBM), and laser metal deposition (LMD) are highlighted as powerful methods for fabricating intricate HEA components. The review addresses the challenges of AM processes, such as porosity, fusion defects, and anisotropic mechanical properties, and discusses strategies to mitigate these issues through process optimization and improved powder quality. The mechanical behavior of AM-processed nickel-based HEAs is thoroughly analyzed, focusing on compressive strength, hardness, and ductility. This review underscores the importance of microstructural features, including grain size, phase composition, and deformation mechanisms, in determining the mechanical performance of these alloys. Additionally, the influence of post-processing techniques, such as heat treatment and hot isostatic pressing (HIP) on enhancing mechanical properties is explored. This review also examines the oxidation behavior of nickel-containing HEAs, particularly the formation of protective oxide scales and their dependence on aluminum content. The interplay between composition, VEC, and oxidation resistance is discussed, offering valuable insights for designing corrosion resistant HEAs. Finally, this review outlines the potential applications of nickel-based HEAs in industries such as aerospace, automotive, and energy, and identifies future research directions to address challenges and fully realize the potential of these advanced materials.

1. Introduction

Metals account for roughly two-thirds of all elements on Earth and approximately 25% of Earth’s mass [1,2]. Metals are generally not used in their pure form because they fail to exhibit good mechanical, structural, and thermal properties unless alloyed [3,4]. This property limitation is overcome by alloying the pure metal with other elements. Traditional alloys provide basic essential features by adding basic essential elements, such as iron, titanium, or aluminum, and adding minor amounts of other factors to regulate the microstructure and enhance the characteristics [5,6]. For hundreds of years, this alloying method has been routinely employed. The quantity of potential compositions limits the traditional design of new alloys [7]. A novel and advanced class of alloys known as multi-principal element or high entropy alloys (HEA) are synthesized by mixing more than four elements in near equal ratios. Depending on the constituting elements, these alloys provide exceptional hardness, high yield stress, and enhanced corrosion resistance, plasticity, and thermal conductivity [8,9,10,11,12,13,14,15,16]. HEAs were first reported in 2004 individually by Yeh and Cantor [17,18]. The nano-structured fine-grained materials (grain size 0.1 µm) have garnered overwhelming interest as contrary to their coarse-grained counterpart [19,20,21]. The exceptional properties of the nano-structured materials originate from higher density grain boundaries, statistically increasing the defect density, hindering the dislocation motion and enhancing corrosion protection [22,23,24]. A widely used synthesis technique for nano-structured materials is ball/mechanical alloying (powder metallurgy), which usually activates a material to nearly 30 kJ/mol [25] followed by consolidation using sintering techniques such as spark plasma sintering. This process is compromised further by the thermal stability of the nano-structured grains as they tend to coarsen under working pressure–temperature conditions. Experiments with low melting nanocrystalline Sn, Cu, Al, Mg, and Pb have shown grain coarsening even at room temperature [26,27]. Moreover, it is technically challenging to counter the trade-off between grain coarsening in alloy powder during the sintering process [28]. It was further revealed that the properties of nano-structured alloys with 1–2 alloying elements did not improve mechanical performance much. Hence, the idea of high configurational entropy was leveraged to stabilize these nano-structured materials as incorporated within high entropy alloys.
Franz Carl Achard, a German physicist and metallurgist, studied alloys that had 5–7 components in equal quantities in the late 18th century, and is regarded as one of the pioneers of HEA research [29]. He studied more than 900 possible alloy combinations made from the metals that were known at the time, such as silver, cobalt, iron, lead, zinc, bismuth, copper, tin, antimony, arsenic, and platinum, and the results were published in his book. Despite the fact that HEAs encompass a broad spectrum of principal elements, they consistently exhibit simple crystal structures. These include body-centered cubic (BCC) [30], face-centered cubic (FCC) [18,31], and hexagonal close-packed structures [32,33], or a combination of these structures [17].
The intriguing characteristics of HEAs significantly enhance their appeal. For example, FCC-structured CrCoNi HEAs exhibit exceptionally high fracture toughness, rivaling cryogenic steels; superior fatigue resistant, oxidation, and magnetic properties; and remarkable irradiation resistance [34,35,36,37,38,39].
The stabilization of HEAs is anchored to four primary effects [40]: (a) high configurational entropy; (b) severe lattice distortion; (c) sluggish diffusion; and (d) cocktail effect [41,42]. While it is frequently demonstrated that large configurational entropy (ΔSmix ≥ 1.5R, R being the universal gas constant) serves as a stabilizing factor for equiatomic multicomponent HEAs [17,43,44], this might not always be accurate due to the interplay between enthalpy and entropy [45,46].
Nickel-based super alloys have been used as high-temperature load-bearing candidates for the past seven decades because they can withstand temperatures up to 1100 °C. They possess excellent room temperature ductility and good creep and fatigue behavior at relatively high temperatures. In the same direction, Ni-based HEAs are developed to further increase the limits of Ni superalloys. This review outlines the potential applications of nickel-based HEAs in industries such as aerospace, automotive, and energy, and identifies future research directions to address challenges and fully realize the potential of these advanced materials.

2. High Entropy Alloys (HEAs)

HEAs consist of multiple principal elements, often four or more, in near-equimolar ratios, along with minor elements. The core concept of HEAs is that solid solution phases exhibit higher stability due to significantly increased mixing entropy compared with intermetallic compounds, especially at elevated temperatures. This makes HEAs suitable for efficient synthesis, processing, analysis, and property tuning [38].
Do complex concentrated alloys, or multicomponent alloys, come under the umbrella of HEAs? Considering an equimolar solid solution, based on Boltzmann’s hypothesis on the relationship between entropy and permutations of all possible states, R ln(n) can be used to calculate its configurational entropy per mole (Equation (1)):
S c o n f = k ln w = R ln 1 n = R ln n
R is gas constant, 8.314 J/K mol, and n is the number of elements.
Physically, entropy has four contributors, namely, configurational, vibrational, magnetic, and electronic. Configurational entropy dominates the other three at room temperature, as compared with vibration at high temperatures [47]. Vibrational entropy at high temperatures could be calculated from phonon calculations in the framework of density functional theory [48,49]. Excess vibrational entropy is found to takeover configurational entropy and lead to phase stabilization at high temperatures [50]. Vibrational entropy is dependent upon mass disorder in the solid solution, which again originates from the incorporation of different elements in alloys [12,49].
In Table 1, the configurational entropy of a three-component equimolar solution exceeds the threshold of 1R, and that of a five-component solid solution is 61% greater. Consequently, it is plausible to consider that 1.5R (even without considering the other three mixing entropy contributions mentioned above) is sufficient to counterbalance mixing enthalpy and can be used as a threshold between HEAs and medium-entropy alloys [47]. This implies that only a select few conventional alloys possess a high configurational entropy greater than 1.5R. The configurational entropy equation indicates that when an element is present at a concentration of 5%, it contributes a mixing entropy of 0.05Rln0.05 equivalent to 0.15R. However, this amount constitutes only 10% of the minimum threshold of 1.5R. Therefore, an element present in a quantity exceeding 5 at.% could be classified as a primary element. The contributions for 4 at.%, 3 at.%, 2 at.%, and 1 at.% are 0.129R, 0.105R, 0.046R, and 0.078R, respectively, corresponding to percentages of 8.6%, 7%, 5.2%, and 3.1% based on 1.5R. Consequently, we categorize an element in a quantity of 5 at.% as a minor element. The question about HEAs’ metallic principal elements’ upper limit arises due to the data in Table 1 and Equation (1). The total configurational entropies for 6, 9, and 11 element alloys, along with 13, 14, 15, 20, and 40 element equal-mole alloys, are 1.79R, 2.2R, 2.4R, 2.57R, 2.64R, 2.71R, 3.0R, and 3.69R, respectively. The change of 0.07R from a 13 element to a 14 element alloy is small (0.07/2.57 = 2.7%), leading to a suggestion of a practical range of principal elements between 5 and 13 for HEAs.
From a chosen set of 13 metallic elements, one can create 7099 different combinations for designing equal-mole HEA systems. This encompasses systems with 5 to 13 elements, adhering to the practical limit of 13 metallic elements as defined earlier [21].
C 5 13 + C 6 13 + C 7 13 + C 8 13 + C 9 13 + C 10 13 + C 11 13 + C 12 13 + C 13 13 = 7099
We could develop an equimolar AlCoCrCuFeNi alloy. We can also create unequal-mole alloys based on structure property relation, with minor alloying elements like AlCo0.5CrCu Fe1.5Ni1.2B0.1C0.15 for further microstructure and property modification. As a result, there are an infinite number of HEAs [47,51]. Table 2 Shows ∆Sconf of some commonly used alloys.

3. Selection of Elements for HEA

The existing literature and bars can be used to identify candidate major constituent elements for possible use in alloys. Table 3 shows possible candidates drawn from the periodic table, excluding radioactive elements. The selection of candidate elements is based on desired attributes such as temperature, density, and strength, among others. Operating temperatures are classified as Low-T (200 °C), Medium-T (600 °C), and High-T (1000 °C) [40,52]. Microstructural properties determine strength, whereas melting temperature is determined by constituents, atomic sizes, and chemical interactions. The Pauling electronegativity and atomic radius are essential factors in alloying element selection [32,44]. The use temperature must be 50% of the temperature at which the alloy will be used. The Hume–Rothery rule identifies individual elements with atomic radii differences within 8%, which is crucial for the synthesis of single-phase alloys [40,52].
Achieving a trade-off between strength and tensile ductility in single-phased HEAs is challenging [13,15,18,53,54,55]. Studies indicate that single-phased FCC-structured HEAs are ductile but lack sufficient strength [31,56,57,58]. HEAs also suffer from elemental segregation leading to composition heterogeneity, which negatively affects structural properties and limits engineering application [59]. Single-phased BCC-structured HEAs are found to be extremely hard, and hence are not ductile enough to be machined or formed [35].
Figure 1 depicts a set of elements chosen to synthesize HEA components through additive manufacturing (AM) methods. These elements primarily include lightweight Al, 3D transition metals, refractory metals, the metalloid Si, and lanthanides La and Sm [60]. In certain variations of HEAs, additional substitutional or interstitial elements such as carbon and nitrogen are mixed as well.

4. Mechanical Properties and Defects of HEAs Processed by Additive Manufacturing

Hardness, tensile, and compression testing, among other AM-HEAs’ mechanical properties, have been investigated. Post-treatment processes like HIP, solution heat treatment, and aging can enhance performance by eliminating metallurgical defects and residual stress in the fabricated machine components [61,62].

4.1. Microhardness

A summary of the structural properties, including microhardness, and the printing parameters of HEAs processed by various additive manufacturing processes is presented in Table 4 and Table 5. In Figure 2, the hardness values of HEAs range from 100 to 1400 HV, influenced by the processing methods and elemental makeup of alloy. In general, HEAs with a BCC crystal structure (e.g., MoNbTaW) exhibit more complex dislocation motion compared with HEAs with an FCC phase (e.g., CoCrFeNi). The more considerable lattice distortion relative to the FCC phase originates from the enhanced solid solution strengthening of the BCC structure in HEAs [60].

4.2. Tensile Properties

Ultimate tensile strengths (UTS) v/s elongation at fracture for several additively manufactured HEAs are presented in Figure 3. The UTSs of various HEAs range from 250 to 1400 MPa. It is evident that within a single HEA system, such as AlCoCrFeNi, the tensile characteristics varied significantly depending on the microstructure achieved using various AM processes and the selected processing parameters [63]. According to Xiang et al. [64] microstructure modification might lead to tuning the structural properties of CoCrFeMnNi specimens via altering the laser parameters such as power and scanning rates. It was reported that the dominating (001) pattern changed to a random texture as the laser power rose from 1 to 1.4 KW, the anisotropy decreased in the structure, and enhanced structural properties in contrast to the alloys were produced by melting-casting. In terms of the Hall–Petch relation, described in [65,66,67], it was discovered that additively made HEAs obey this relationship. The yield strength of a polycrystalline material is dependent on the statistical distribution size of the grains. The CoCrFeNi HEA deposited via SLM exhibits remarkable strength and ductility due to the lack of brittle intermetallic compounds or secondary phases. According to Guan et al. [68], the tensile deformation of the LMD-deposited CoCrFeMnNi HEA primarily involved dislocation actions aided by deformation twins, resulting in a tensile strength comparable to that of counterparts made from finer-grained wrought-annealed materials [60]. The UTS and fracture elongation of AM-processed HEAs demonstrate high variability even within the same alloy systems. This significant variance is attributed to the different microstructures obtained from the diverse AM technologies and processing parameters employed. For instance, in laser metal deposition (LMD)-fabricated CoCrFeMnNi, the tensile properties are tuned by varying the laser parameters such as power and scanning patter/rate which modifies the crystallographic texture [69].
Table 4. The mechanical properties (tensile properties and hardness) of various Ni-containing HEAs.
Table 4. The mechanical properties (tensile properties and hardness) of various Ni-containing HEAs.
HEAs SystemAM TechniquesPowder Size (μm)Density Relative DensityPrinting ParametersPhase CompositionsYield Stength (MPa)Ultimate Tensile Strength (MPa)Elongation (%)Hardness (HV)Ref.
Al0.3CoCrFeNiSLM20–42 7.7699.9P = 150–170 W, v = 1100–1300 mm s−1, t = 25–30 µm, h = 45 µmFCC~73089629-[70]
Al0.5CoCrFeNiSLM15–45--P = 400 W,
v = 270 mm s−1, t = 40 µm, h = 90 µm
FCC57972122262.5 ± 5[71]
AlCoCrFeNiEBM45–105---FCC + BCC----[72]
Al0.1CoCrFeNiSLM20–63--P = 150 W,
v = 1600 mm s−1, t = 50 µm, h = 100 µm
FCC-520~2-[73]
Al0.5CoCrFeNiSLM37–74-99.92P = 320 W,
v = 800 mm s−1, t = 60 µm,
h = 50 µm
FCC + BCC609878~0.175270[74]
AlCoCrFeNiSLM27–65--P = 120 W,
t = 50–80 µm,
h = 25 µm
BCC---541 ± 18[75]
AlCoCrFeNiSLM---P = 98 W,
v = 2000 mm s−1,
t = 52 µm
BCC + B2----[76]
AlCoCrFeNiEBM~70--Voltage = 60 kV, vacuum = 7 × 10−1 Pa of He, build plate temperature = 1173 KFCC + BCC + B2769 ± 12.71073.5 ± 21.31.2 ± 0.2-[77,78]
AlCoCrFeNiTiSLM22–65--P = 260 W, v = 1000 mm s−1, t = 50 µm, h = 70 µmFCC + Oxides1235155010.69-[78]
Al4(CoCrFeNi)94Ti2SLM22–65--P = 400 W, v = 10 mm s−1, h = 460 µmFCC----[79]
Al0.5Cr0.8CoFeNi2.5V0.2SLM15–75-~98.9P = 250 W, v = 960 mm s−1, t = 40 µm, h = 80 µmFCC530 ± 151842 ± 3540263 ± 10[80]
AlCrFeNiVSLM20–80-99.88P = 140 W, v = 900 mm s−1, t = ~30 µm,
h = 50 µm
FCC651.36~1057.4730.3-[81]
FeCoCrNiSLM---P = 200 W,
v = 300 mm s−1, t = 20 µm
FCC60074532238[65]
---P = 200 W,
v = 300 mm s−1, t = 50 µm
FCC4024808205[65]
14–48-99.71 ± 0.25P = 200 W,
v = 740 mm s−1, t = 40 µm,
h = 40 µm
FCC581.9707.920218[82]
14–48-99.71 ± 0.25Annealed at 1573 K for 2 hFCC221633.245138[82]
FeCoNiCuAlSLM15–65-~96P = 100–400 W, v = 400–1200 mm s−1,
t = 40 µm,
h = 80 µm
BCC matrix + Cu-rich B2---~858[83]
Ni6Cr4WFe9TiSLM15–53--P = 300 W, v = 2500 mm s−1,
t = 100 µm,
h = 80 µm
-74297212.2-[84]
CoCrFeNiMnEBM45–106--Reference current = 2–14, scanning speed = 492–3446 mm/s), layer thickness = 50–70 µm, hatch depth = 50–70 µm-205 ± 3497 ± 263 ± 1157.1[85]
CoCrFeNiW0.2SLM15–45--P = 175 W, v = 150 mm s−1, t = 30 µm, h = 100 µmFCC + W610 ± 15814 ± 917 ± 1314.1[86]
CoCrFeMnNiSLM5–45--P = 400 W,
v = 830 mm s−1,
t = 30 µm,
h = 90 µm
FCC-60135-[87]
SLM---P = 400 W,
v = 830 mm s−1,
t = 30 µm,
h = 90 µm then HIP temp 1150 °C, time 3 h, and pressure 150 MPa
FCC-64918-[87]
SLM--99.2P = 240 W, v = 2000 mm s−1, t = 40 µm, h = 50 µmFCC510 ± 10609 ± 1034 ± 3-[88]
SLM5–45-99.2P = 240 W, v = 2000 mm s−1, t = 40 µm, h = 50 µm then
HT at 900 °C for 1 h in Ar atmosphere
FCC381 ± 8619 ± 547 ± 2-[88]
CrFeNiMnSLM20–45-98.0% ± 0.1P = 70 W,
v = 200 mm s−1, t = 25 µm,
h = 90 µm
FCC---248 ± 8[89]
NbMoTaWSLM---P = 400 W,
v = 250 mm s−1, t = 100 µm, h = 100 µm
----826[90]
Ni6Cr4WFe9TSLM15–53--P = 300 W, v = 2500 mm s−1, t = 100 µm, h = 80 µm-74297212.2-[84]
P = power (W), v = scan speed (mm s−1), t = layer thickness (µm), and h = hatching space (µm).
Table 5. The mechanical properties (compressive strength and hardness) of various Ni-containing high entropy alloys.
Table 5. The mechanical properties (compressive strength and hardness) of various Ni-containing high entropy alloys.
HEAs SystemAM TechniquesPowder Size (μm)Density Relative DensityPrinting ParametersPhase CompositionsYield Compressive Stength (Mpa)Compressive Fracture Stength (Mpa)Elongation (%)Hardness (HV)Ref.
AlCoCrFeNiEBM~70--Voltage = 60 kV, print orientation 0 degreesFCC + BCC1015.0 ± 52.51668.3 ± 71.526.4 ± 6.7-[91]
AlCoCrFeNiEBM~70--Voltage = 60 kV, print orientation 90 degreesFCC + BCC944.0 ± 55.41447.0 ± 135.814.5 ± 5.3-[91]
Al2.1Co0.3Cr0.5FeNi2.1WAAM---Wire feed speed 8 m/min, voltage 17 V, travel speed 0.3 m/min, substrate heating temperature ≈ 250 °CA2 + B2550189918.5463[92]
AlCoCrFeNiBJ~32--50 μm layer thickness, recoating speed of 1 mm/s, binder saturation 60%, heating power of 50%, and a drying time of 25 s, sintering at 1320 °C for 4 h, annealing at 1000 °CFCC + BCC + B21203 ± 221996 ± 4532.25 ± 2.5418.1[93]
AlCoCrFeNiBJ~32--50 μm layer thickness, recoating speed of 1 mm/s, binder saturation 60%, heating power of 50%, and a drying time of 25 s, sintering at 1320 °C for 4 h, annealing at 1200 °CFCC + BCC + B21461 ± 232272 ± 4831.46 ± 2.1499.6[93]
AlCrCuFeNiSLM11–58--P = 300 W, v = 600 mm s−1, t = 40 µm, h = 80 µmBCC-2052.8 ± 123.66.8 ± 1.3-[94]
AlCrCuFeNiW1LMD90–150--P = 1600 W, v = 20 mm s−1, t = 40 µmFCC + BCC998.41274.624.3240 ± 10[95]
AlCrCuFeNiW3LMD90–150--P = 1600 W, v = 20 mm s−1, t = 40 µmFCC + BCC1005.91287.923.4230 ± 10[95]
CoCrFeMnNiSLM18–38--P = 90 W, v = 600 mm s−1, t =25 µm, h = 80 µm in scanning directionFCC778.4---[70]
CoCrFeMnNiSLM18–38--P = 90 W, v = 600 mm s−1, t = 25 µm, h = 80 µm in transverse directionFCC766.4---[70]
CoCrFeMnNiSLM18–38--P = 90 W, v = 600 mm s−1, t = 25 µm, h = 80 µm in building directionFCC703.5---[70]
P = power (W), v = scan speed (mm s−1), t = layer thickness (µm), and h = hatching space (µm).
Figure 3. Different AMed HEAs’ UTSs against fracture elongation. The HEAs manufactured by SLM, LMD, EBM, and WAAM, respectively, are represented by the solid, top half-solid, open, and bottom half-solid symbols (just considering the as-fabricated state) [60]. Adapted with permission from ref. [60]. A 2022 Creative Commons Attribution (CC BY) license.
Figure 3. Different AMed HEAs’ UTSs against fracture elongation. The HEAs manufactured by SLM, LMD, EBM, and WAAM, respectively, are represented by the solid, top half-solid, open, and bottom half-solid symbols (just considering the as-fabricated state) [60]. Adapted with permission from ref. [60]. A 2022 Creative Commons Attribution (CC BY) license.
Materials 17 03826 g003
The majority of AM-processed HEAs demonstrate adherence to the Hall–Petch relationship, showing an inverse correlation between yield strength and average grain size. Achieving remarkable strength alongside required ductility has been realized by avoiding the creation of brittle intermetallic phases via compositional tuning and/or thermomechanical processing. For example, the as-deposited CoCrFeNi HEA via LMD reached an elongation of 92.5% alongside strength surpassing its cast analogue [96]. Alloying HEAs with interstitial elements (C, N, Si) through AM technique power flexibility synergizes strength and ductility. The improvements originate from varied mechanisms including nano-scale solute clustering, dislocation slip, deformation twinning, and strain-prompted phase transformations. Certain metastable HEAs like Fe60Co15Ni15Cr10 showcase high tensile ductility after first necking owing to the steady spread of strain-generated martensitic transformation domains. For example, an as-printed sample of this alloy composition exhibited 69.5% elongation at acceptable strength levels [97]. Eutectic HEAs processed via AM form alternating nano-lamellar microconstituents spacing within colonies, conferring simultaneous high yield strength (~1000 MPa) and reasonable tensile ductility (~20%) [98,99,100,101]. Their impressive strength–ductility balance highlights their candidacy for AM-related structural applications. In summary, AM provisionally enables extensive microstructure and composition tailoring avenues to develop HEAs with excellent strength-ductility synergies for load-bearing components. The careful selection of AM method and tuning fabrication parameters alongside judicious alloying additions can elicit optimized tensile properties.

4.3. Compressive Properties

Compressive strength analysis in HEAs additive manufacturing is essential for determining material appropriateness for parts that could be compressed, including those in the automobile or aerospace sectors. The optimization of HEA composition and printing settings is guided by compressive strength data, guaranteeing that constructions possess the requisite strength for their designated purpose. It is essential for assessing structural integrity, performing failure analysis, and arriving at well-informed design choices. Compressive strength is a statistic used in quality control operations that helps with material certification and standardization in additive manufacturing. Furthermore, these data help with predictive modeling, which enables engineers to estimate how HEA constructions will behave under different compressive stress scenarios. The compressive strength, hardness, phase composition, printing parameters and powder size to produce various Ni-containing alloys are listed in Table 5.
Figure 4 shows a comparison between the different types of Ni-based HEAs obtained from various studies. Fujieda et al. showed electron-beam-melted (EBM) AlCoCrFeNi had much higher deformability (1668 MPa compressive strength) compared with cast AlCoCrFeNi, ascribed to the finer grains and enhanced ductility of the FCC-structured phase from EBM [91]. Meanwhile, increasing Al content in laser-metal-deposited (LMD) AlxCoCrFeNi increased compressive strength but reduced ductility [54,102]. Al0.3CoCrFeNi sustained high work hardening rates exceeding 1300 MPa compressive strength with a strain of 1 due to deformation twinning, while BCC Al0.85CoCrFeNi showed higher power (2250 MPa) but lower elongation (24.5%) due to second phase particles [103]. Significant tension-compression property differences were reported in LMD Al0.3CoCrFeNi as well, related to the absence of deformation twins in tension [103]. Anisotropy in properties was also found in selective laser melted (SLM) CoCrFeMnNi due to differences in dislocation and twinning behavior along different directions [70]. A yield strength prediction model involving grain size, oxides, and orientation-dependent Taylor factors matched these experiments. Similar anisotropy occurred in LMD AlCoCrFeNi2.1 and SLM AlCrCuFeNi during compression testing [94,100]. For refractory HEAs, LMD MoNbTaW (W = 0) exhibited an 874 MPa yield strength and a 1140 MPa compressive strength. Meanwhile, laminated LMD AlCoCrFeNiTi/CoCrFeMnNi showed yield strengths of up to 990 MPa and no fractures at strains as large as 80%, significantly outperforming monolithic HEAs. This demonstrates AM’s capability to create HEAs with enhanced property combinations [68]. It is important to note that AM HEAs can attain high compressive strengths exceeding 2 GPa but anisotropy and a lack of strain causing failure are persistent challenges. Alloy and process optimization to promote twinning and fine, uniform microstructures is critical, as is the exploration of laminated architectures that can achieve synergistic strengthening across interfaces.
Lopes and Oliveira studied the Optimization of Process Parameters in Laser Powder Bed Fusion of SS 316L Parts Using Artificial Neural Networks [104]. To find the optimal process parameters for user-required part properties in laser powder bed fusion, an optimization model based on experimental data was developed using artificial neural networks. An exclusive neural network was developed to optimize the laser power, scan speed, and hatch spacing with the desired relative density, surface roughness, dimensional error, and microhardness. Their predicted values were compared with the experimental values in the literature and indicated a close match. They also demonstrated the ability of an exclusive neural network in modeling the process parameter–property relationships for multiple properties. The developed model can find the optimal processing parameters that satisfy the user requirement for customized part properties. Thus, it reduces the preprocessing time and cost significantly. The following conclusions are drawn based on the current study. Finding the optimal parameters for the laser powder bed fusion process requires an understanding of the combined effect of the process parameters on the part properties. ANN is a powerful tool in modeling the combined relationship and obtaining the optimal process parameters in the given range of data.

4.4. Porosity

One of the leading causes of the uneven porosities inside the AM components is inadequate melting due to incomplete fusion. Still, there are other factors as well that raise serious concerns. When there is not enough fusion, the melt pool does not go far enough into the preceding layer to facilitate the formation of metallurgical bonds with the subsequent layers [62]. Underneath the present layer, un-melted and unconsolidated powder remains trapped. The remaining powdered areas then appear as pores in the finished product. Optimizing processing parameters to obtain sufficient energy density and melt pool penetration for complete consolidation without excess is necessary to reduce the lack of fusion [105].
Islam et al. [106] studied six metallic powders characterized in terms of PSD using a range of techniques and concluded that spherical powder produced more acceptable results over the range of equipment used. Larger differences arose as the morphology deviated away from spherical. Sofia et al. [107] reported that at a fixed scan velocity and power of the laser, the melting of the powders was significantly affected by their size. Yuan et al. [108] studied SLM-manufactured GH3230 nickel-based superalloys and explored the impacts of various post-processing on the microstructural advancement and extensive mechanical properties of the GH3230 superalloy. They reported that after HIP treatment, the liquid pool structure vanished, and the microstructure principally comprised columnar grains.
The size and penetration of the melt pool were influenced by several parameters, including the preheat temperature, scanning speed, focus offset, and beam current. Figure 5 shows images of the as-built EBM CoCrFeNiMn HEA parts. Figure 5a shows the whole part with surrounding powders adhered to the sides, indicating non-optimal contour processing parameters and roughness. Figure 5b shows the rough side surface finish under SEM. Figure 5c shows the characters’ cuboid sample top surfaces exhibiting varying swelling and lack of fusion defects. Swelling occurs when process parameters cause huge melt pools and high temperatures. An absence of fusion defects occurs when there is insufficient energy density to fully melt and fuse each layer.
Porosity is a challenge in additively manufactured parts, which is very important to take care of while printing. Figure 6 illustrates the significant impact of feedstock powder quality on the final porosity levels in additively manufactured CoCrFeNiMn HEA parts built by EBM [85]. Optical microscopy reveals that most pores in the EBM samples match the argon gas pores captured within the gas-atomized powder particles utilized as the powder bed feedstock. More pores in the powder means more pores in the printed part. Furthermore, a lack of fusion defects manifests as larger pores during EBM processing if the energy density is insufficient to fully consolidate each layer, sharply escalating porosity. Wang et al. [85] demonstrated that while the gas-atomized alloy powder itself exhibited suitable morphology and flowability for EBM, the optimization of the EBM process parameters was essential to minimize both types of pores and achieve 99%+ relative density. This extensive parameter tuning would be considerably significantly reduced with high quality, pore-free powder feedstock. In summary, Figure 6 signifies that powder quality fundamentally constrains the final part quality, and eliminating internal gas pores in the atomization process is critical to enable the reliable, efficient printing of high-performance additively manufactured metal alloys [85].

5. Effect of VEC on Ni-Based HEAs

The structural properties of HEAs are influenced by their microstructure, which is significantly determined by valence electron concentrations (VEC) of the HEA composition [109]. The VEC is determined from the statistical average of VECs of constituting elements of the HEA. Guo et al. [109] explored the relationship between VEC and crystal structure in various HEAs as depicted in Figure 7. Their findings suggest that alloys with a VEC greater than 8 tend to stabilize in the FCC phase, those with a VEC between 6.87 and 8 exhibit a mixture of FCC and BCC phases, and those with a VEC less than 6.87 favor the BCC phase. However, this rule has no consideration for HCP-structured alloys and physically it is impossible for all outer shell electrons to take part in bonding process; moreover, the VEC rule provides a crude solution to predict crystal structure of HEAs.
Vida et al. [110] reported two different HEAs, one Ni35Fe30Cr20Mo10W5 which is equiatomic, and Ni35Fe30Cr20Mo10W5, which is non-equiatomic. The VEC values for these alloys were calculated to be 7.2 and 8, respectively. With a VEC value around 8, both alloys exhibited a mixture of phases, but their structures could shift between BCC and FCC, affecting their properties, as illustrated in Figure 8. Thus, when the VEC approaches 8, the FCC phase becomes more dominant.
Synthesizing HEAs requires balancing between strength and ductility. Higher VEC tend to promote FCC structures, which enhance ductility but often at the cost of strength. To optimize the power and flexibility, Chen et al. [111] aimed to enhance strength and ductility by alloying Ni and Mo within two different types of HEAs, namely (AlCoCrFeNi)100−xNix and (CoCrCuFeNi)100−xMox, by utilizing VEC to tailor the phase composition. It is important to note that AlCorFeNi exhibits a BCC structure and CoCrCuFeNi exhibits an FCC structure. It was seen that when increasing the Ni percentage in the alloy (AlCoCrFeNi)100-xNix, the VEC shifts towards the FCC structure, as illustrated in Figure 9. When the Ni content varies from 0% to 16%, the compressive fracture strain improves from 25% to 40%. Likewise, as the Mo percentage increases in the (CoCrCuFeNi)100−xMox, the VEC shifts towards the BCC structure, as illustrated in Figure 10. The compressive yield strength showed a consistent increase from 260 MPa to 928 MPa as the Mo content varied from 0% to 16%.
Figure 9 shows SEM micrographs of as-cast HEAs with varying Ni percentages, demonstrated in the left column. The regions identified in the left column are magnified and displayed in greater detail in the corresponding areas on the right column. Figure 9a has confirmed that the AlCoCrFeNi HEA exhibits a uniform, single phase structure. The enlarged SEM image in Figure 9b (for labeled area A in Figure 9a) resembles honeycomb-like structures with uniformly distributed spherical particles. The reported microstructures matched the microstructures of AlCoCrFeNi HEA synthesized by Bridgman solidification [112]. Additionally, they resembled the as-cast Fe-Ni-Mn-Al alloy microstructure [113]. Figure 9c shows the microstructures for the (AlCoCrFeNi)96Ni4 HEA. It reveals the existence of FCC phases with a deficiency of aluminum, noticeable at the boundaries of the grains. These phases were also observed within the grains themselves, as depicted in Figure 9d. Notably, their small volume fraction made them undetectable using XRD techniques. The microstructures depicted in Figure 9e,f reveal the characteristics of the (AlCoCrFeNi)92Ni8 HEA. In the (AlCoCrFeNi)96Ni4 HEA, an aluminum-deficient composition, FCC phases of notable size were identified, surpassing those observed in the (AlCoCrFeNi)96Ni4 HEA. Figure 9g–j illustrate the microstructures of the (AlCoCrFeNi)100−xNix HEA specifically for x values of 12 and 16. The SEM micrographs indicated a gradual growth of precipitated FCC phases as the percentage of Ni increased, highlighting an Al-deficient composition. This observation suggests a positive influence of Ni on the transformation from BCC to FCC, given its role as an FCC stabilizer. Within the as-cast AlxCoCrFeNi alloy, using varying concentrations of Al as the alloying element in CoCrFeNi, it was found that Al significantly enhances the mechanical properties by stabilizing the BCC structure of the Al-rich alloy as compared with Al-deficient alloys [114,115,116]. In principle, the mechanical properties are dependent on their microstructures and this alloy also exhibits various crystal structures and microstructures depending on the aluminum content, ranging from FCC structures at low Al content levels to BCC phases with spinodal decomposition at higher Al content levels. The microstructure evolves from columnar cellular and dendritic structures to equiaxed nondendritic and dendritic grains, with the Widmanstätten side plates and complete spinodal structures forming at specific Al content levels [114].
The black regions in Figure 9g–i were detected as O with EDS. The origin of oxygen in the system, whether from the initial raw material or during subsequent processing stages, remained unclear. The minuscule sizes of oxide particles had no discernible impact on the structural properties of HEAs. The selection of Mo, with a lower VEC compared with the average VEC value for the CoCrCuFeNi alloy, was selected as an element for incorporation into the CoCrCuFeNi matrix to prove the feasibility of the alloy design. Figure 10 shows the micrographs of the as-cast alloy with varying Mo percentages.
Figure 10a shows single phase CoCrCuFeNi HEA. As seen in the high magnification micrograph in Figure 10b (focused on area A in Figure 10a), Cu-rich phases were discovered along the grain boundaries, as depicted in Figure 10b–d, which display micrographs at varying magnifications for the (CoCrCuFeNi)96Mo4 HEA. Within the grains, Cu-rich phases were observed, highlighting the distribution within the material. Figure 10e,f showcase magnified microstructures from the (CoCrCuFeNi)92Mo8 HEA. In the microstructures, BCC phases containing Mo are evident, with their volume fraction appearing to rise alongside the Mo percentage, as depicted in (Figure 10g–j). This suggests a beneficial impact of Mo on the transition from FCC to BCC.

6. Pressure-Induced FCC to HCP Phase Transition in Ni-Based HEAs

Ni-based high entropy alloys with an FCC structure can undergo a phase transformation to an HCP structure under extremely high pressures. This pressure-induced FCC to HCP phase transition has significant implications for the mechanical properties and potential applications of these multi-principal element alloys. Zhang et al. [117], investigating the transition of FCC to HCP phase via pressure-induced strains, proposed that this has to be a martensitic transformation. These mechanisms require less activation energy. In this study, three HEAs were arc melted and drop cast, namely, NiCoCr, NiCoCrFe, and NiCoCrFePd, from individual metal powders of high purity (>99.9%). All alloys were found to crystalize into a face-centered cubic structure, as presented in Figure 11.
Synchrotron XRD studies revealed that these HEAs undergo a lethargic FCC to HCP transition at high pressure. Figure 12a–c shows the XRD profile of “NiCoCrFe”, “NiCoCr”, and “NiCoCrFePd” alloys, respectively. The NiCoCrFe alloy exhibits an FCC–HCP transition at around 13.5 GPa. Similarly, ternary alloy “NiCoCr” undergoes a transformation from FCC to HCP at 45 GPa while quinary alloy “NiCoCrFePd” shows transformation from FCC to HCP at a much higher pressure of 74 GPa. The diffraction peaks marked with * are from Au pressure marker as shown in Figure 12c. The sluggish nature of this polymorphic transition and the varying transition pressures across different alloy compositions highlight the importance of elemental chemistry while down selecting elements for HEAs.
HCP and FCC structures show densely packed arrangements of atoms, contrasting each other only in stacking order [118]. Studies found that alloys underwent an HCP shift because of defects caused by pressure, resulting in altered stacking fault from ABCABC… to ABAB… This aligns with the lower stacking fault energy observed in HEAs [119,120].

7. Precipitation-Strengthened Ni-Based HEAs

HEAs are known to exhibit remarkable strength while maintaining superior ductility. Precipitation strengthening mechanisms hold promising potential in this domain, offering avenues for tailoring the required properties in these alloys. Tsao et al. [121] developed a cost-effective method to produce Ni-based HEAs exhibiting exceptional high-temperature performance on par with traditional Ni-based superalloys. Their approach aimed to create advanced materials that could withstand extreme thermal conditions while maintaining affordability and practicality. For high temperature applications, the alloy must exhibit exceptional structural integrity and oxidation resistance at elevated temperatures [122,123]. To stabilize the surface in Ni-based HEAs, L12 Ƴ’ precipitate was procured, which is attributed to the coherency of Ƴ’ with the matrix [124]. Moreover, intermetallics/ordered phases like σ, µ, η, etc., tend to make HEAs brittle due to principate-induced higher interfacial energy with the matrix [125]. Tsao et al. [121] synthesized two HEAs, as tabulated in Table 6.
The microstructure in Figure 13 shows that, after solution and aging treatment, the Ƴ’ uniformly dispersed precipitates within the FCC Ƴ matrix. Further, the inset diffraction pattern confirms the presence of an L12 ordered γ’ phase, with no other degrees detected, which is comparable to conventional Ni-based superalloys [124]. Both HE-Superalloys exhibit Ƴ’ precipitate with spherical morphology.
Table 7 shows the composition of Ƴ/Ƴ’ of HESA and some common Ni-based superalloys [121]. Ƴ and Ƴ’ are Ti-rich solid solutions, with Ƴ’ having a higher Ti content. Additionally, Fe promotes partitioning between Ƴ and Ƴ’, causing lattice misfit, which enhances strengthening by increasing the anti-phase boundary energy. Moreover, the incorporation of less refractory elements and high-temperature stability might exceed that of commercial superalloys.

8. Substitute for TRIP and TWIP Steels via Ni-Based HEAs

Ni-based HEAs have emerged as promising candidates to substitute transformation-induced plasticity (TRIP) and twinning-induced plasticity (TWIP) steels, offering superior mechanical properties and corrosion resistance [126,127]. These multi-principal element alloys leverage compositional complexity to achieve exceptional strength-ductility synergy through tailored stacking fault energies and deformation mechanisms. In particular, AlxCoCrFeNi alloys typically exhibit either FCC or BCC structures, contingent to the values of x [114,128]. Notably, Yasuda and colleagues [129] designed the Al0.3CoCrFeNi alloy, synthesized using a plasma arc furnace, and revealed the presence of a NiAl phase with a B2 structure precipitated along the grain boundaries. During cold rolling, these fine precipitates were found to pin the grain boundaries, thereby inhibiting grain growth. Figure 14 presents the SEM backscattered electron micrographs and their pole figures.
Figure 14a–c illustrate that precipitates primarily form along grain boundaries at temperatures up to 1273 K. However, at 1373 K, these precipitates tend to dissolve. Figure 14e presents the SAED pattern of the residue, and on subsequent EBSD analysis in Figure 14f, identifies the B2 structure dispersed within the FCC matrix. A noteworthy observation was made during the annealing of a single crystal Al0.3CoCrFeNi alloy: no B2 precipitates appeared even after 100 h at 1273 K. This indicates that the B2 structure is inclined to form at grain boundaries.
The mechanical properties of the alloy were analyzed and are presented in Figure 15a,b. As illustrated in Figure 15a the alloy annealed at temperature at or below 1273 K displays a high yield stress exceeding 500 MPa. Additionally, an optimal combination of ultimate yield strength (over 1 GPa) and ductility (over 40 percent) is observed at an annealing temperature of 1073 K, which is equivalent to that of transformation-induced plasticity (TRIP) and twinning-induced plasticity (TWIP) [129].
In Figure 15b, yield stress is plotted against 1 d m for the recrystallized alloys annealed between 1073 K and 1373 K. Despite the absence of precipitates at temperatures beyond 1273 K, a linear trend is still observed. This indicates that grain boundary hardening remains the primary factor contributing to the enhanced strength for recrystallized Al0.3CoCrFeNi alloy.

9. Thermal Shock and Oxidation Resistance on Ni-Containing HEA

The mechanical characteristics of alloys made from one or multiple solid solution phases are distinctive. Studies show that BCC-structured HEA systems have excellent hardness and strength under compression [17,130,131], whereas FCC-based HEAs exhibit excellent plasticity.
By arc melting in a tungsten electrode-based melting chamber with inert atmosphere resting on a water-cooled copper plate, elemental powders of 99.99 wt% purity were combined in molar ratios chosen for enabling the production of alloys with the elemental percentages stated in Table 8 [132]. High purity Ti powder is generally employed as a getter. Resultant pallets are then remelted five times to acquire appropriate homogenization. The samples were then cut into discs with a diameter of about 150 mm and 1.5 mm thick, using a precision saw. Following these preparations, the samples underwent cyclic oxidation in a normal environment at 1273 K. Each phase of the thermal shock cycles, involving heating and cooling, lasted for a duration of 2 h. The heating and chilling processes each took about a minute. Throughout the oxidation process, more than 500 cycles (or more than 1000 h) were completed [132]. Figure 16a,b, respectively, exhibit SEM micrographs of the cross-sections for Al25Co25Cr25Ni25 and Al20Co25Cr25Ni25Si5 after cyclic oxidation at 1273 K, and Figure 16c,d show images of Al9.1Co18.2Cr18.2Fe18.2Ni36.3 at two different magnifications. Table 9 contains the EDS map findings from the zones shown in Figure 16a,b,d. These show chemical compositions discovered using EDS point analysis on Al25Co25Cr25Ni25, Al20Co25Cr25Ni25Si5, and Al9.1Co18.2Cr18.2Fe18.2Ni36.3 cross-sections following oxidation under thermal shock state at 1273 K in atmospheric environment.
According to Table 8, the thin intermediate layer of the scale developed on Al25Co25Cr25Ni25 predominantly comprises aluminum and oxygen, with trace amounts of the other constituent elements. In contrast, the outer layer of the scale is primarily composed of chromium and oxygen. Underneath this layer, where considerable levels of Al are also found, the oxide scale has significantly higher concentrations of Cr, Co, and Ni [132]. There is an Al-rich layer at the scale/substrate contact (dark phase zone). This dark zone is in connection with a bright degree consisting primarily of Ni, Co, and Cr. It has been established that the Co-Cr-Fe-Ni rich zone is the FCC structured in high entropy alloys of AlCoCrFeNi [133,134,135,136,137]. This is also consistent with Butler et al.’s [119] interpretation of oxidized Al20Co25Cr25Ni25Si5. Contrary to this, the NiAl-rich B2 phase is distinguished by its darker dendritic regions within the substrate [114,128,129]. The similar procedure produces continuous Cr2O3 production in the Al9.1Co18.2Cr18.2Fe18.2Ni36.3 sample, coupled with a CoCrFeNi spinel structure at the scale surface. In contrast, non-continuous Al2O3 precipitates exhibit significant growth depths within the metallic substrate. Also found close to the scale/substrate interface were some Cr2O3 precipitates. The scales grown on Al9.1Co18.2Cr18.2Fe18.2Ni36.3 were not enough to address this issue, leading to a persistent mass increase after a certain amount of oxidation time. The majority of results indicate that oxide scales formed on Al25Co25Cr25Ni25 and Al20Co25Cr25Ni25Si5 were adequate for shielding those metallic substrates from notable internal oxidation [132]. This suggests that maintaining a continuous production of Al2O3 is essential to establish a robust barrier against internal oxidation when operating for extended periods in high-temperature conditions. Al20Co25Cr25Ni25Si5 appears to have a higher Al contribution, which explains why this material has slightly superior oxidation kinetics [132].
The research of Gawel et al. [132]. found that an AlCoCrFeNi HEA containing aluminum content of less than 9 at.% Al does not always ensure that Al2O3 layers continue to form inside the scale that forms as the substrate oxidizes. Further, the oxide scale forms a continuous layer made entirely of Cr2O3 and a mixed/complex oxide phase near the oxidation region. In contrast, it is found that when Aluminum composition reaches a minimum of 20 atomic %, in the case of Al25Co25Cr25Ni25 and Al20Co25Cr25Ni25Si5 materials, a highly protective continuous Al2O3 layer along with Cr2O3 is found. Interestingly, oxide materials, such as Cu2O, exhibit unique properties that can significantly enhance their functionality [138,139,140]. Whan coupled with HEAs these characteristics make them effective in forming protective layers, contributing to improved applications and material performance.

10. Concluding Remarks

In conclusion, this review comprehensively examines the advancements in Ni-containing high-entropy alloys, spanning from fundamental understanding to alloy design intricacies and manufacturing processes. Discussions encompass mechanical properties, additive manufacturing defects, and the influence of factors like valence electron concentration (VEC) and pressure-induced phase transitions. The review also discusses new uses, like replacing TRIP and TWIP steel, and the impressive performance of Ni-based HEAs in resisting thermal shock and oxidation. As research in this field progresses, the insights presented here will continue to guide future investigations and innovations in advanced materials science. This paper also aims to systematically summarize the knowledge surrounding Ni-based high-entropy alloys. Various manufacturing techniques such as casting, additive manufacturing, and powder metallurgy are employed in creating high-entropy alloys, each affecting the characteristics of the alloy differently. Furthermore, additive manufacturing stands out as an up-and-coming method for processing high-entropy alloys, offering unique advantages in terms of design flexibility, rapid prototyping, and creating complex geometries. This capability is especially beneficial for Nickel-containing HEAs, where additive manufacturing techniques can facilitate precise control over alloy composition and microstructure, ultimately enhancing the structural properties and performance. Given their exceptional qualities, HEAs have garnered significant interest, offering many options for tailoring properties through different manufacturing processes and alloy compositions. Particularly noteworthy is the excellent mechanical characteristics exhibited by Ni-containing high-entropy alloys at extreme temperatures, enabling them to be well-suited candidates in structural components for aerospace engines. The formation of aluminum oxide (Al2O3) and chromium oxides (Cr2O3) layers during oxidation plays a crucial role in reducing the oxidation rate, preserving the material from surface deterioration. In the future, delving deeper into the thermodynamics of phase formation, studying kinetics, exploring various processing methods, and utilizing computer modeling and simulations holds the potential to open up new avenues in alloy research. As research in this field advances, the insights presented here will continue to serve as a guiding beacon for future investigations and innovations in advanced materials science.

Author Contributions

Conceptualization, A.K.G.; methodology, A.K.G.; validation, A.K.G., A.C., A.R., A.T., P.S. (Prince Sharma), A.G., P.S. (Prathamesh Sapale), R.T.A.T., R.M. and A.K.; formal analysis, A.K.G. and A.G.; investigation, A.K.G. and A.G.; data curation, A.K.G.; writing—original draft preparation, A.T.; writing—review and editing, A.K.G., A.C., A.R., P.S. (Prince Sharma), A.G., P.S. (Prathamesh Sapale), R.T.A.T., R.M. and A.K.; visualization, A.C., A.R., P.S. (Prathamesh Sapale), R.T.A.T., R.M. and A.K.; supervision, A.T.; project administration, A.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This is a review study and it deeply analyze the previously reported data to understand the fundamentals for Nickel-Containing High-Entropy Alloys.

Acknowledgments

Ashish Kumar Gupta and Aditya Rane would like to acknowledge the School of Mechanical and Aerospace Engineering, Oklahoma State University, Stillwater, USA. Amit Choudhary would like to acknowledge the Department of Mechanical Engineering, Cleveland State University, Cleveland, OH, USA. Abhishek Tiwari would like to acknowledge the Department of Mechanical and Aerospace Engineering, Indian Institute of Technology Hyderabad, Telangana, India for his post-doctoral fellowship during his tenure from July 2022 to January 2023. Ashutosh Gupta would like to acknowledge the Department of Zoology, Dayanand Vedic College, Orai, UP, India. Prathamesh Sapale, Ravi Teja A. Tirumala would like to acknowledge the School of Chemical Engineering, Oklahoma State University, Stillwater, OK, USA. Rajmohan Muthaiah would like to acknowledge the School of Aerospace and Mechanical Engineering, University of Oklahoma, Norman, OK, USA. Abhishek Kumar would like to acknowledge J. Mike Walker, ’66 Department of Mechanical Engineering, Texas A&M University, College Station, TX, USA.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sharma, P. A Brief Account of Man, Material and Manufacturing: On the Timeline. Mater. Today Proc. 2022, 66, 3572–3577. [Google Scholar] [CrossRef]
  2. ASM International. Properties and Selection: Irons, Steels, and High-Performance Alloys; ASM International: Materials Park, OH, USA, 1990; ISBN 978-1-62708-161-0. [Google Scholar]
  3. Chikumba, S.; Rao, V.V. High Entropy Alloys: Development and Applications. In Proceedings of the 7th International Conference on Latest Trends in Engineering & Technology (ICLTET 2015), Pretoria, South Africa, 25–27 November 2015; pp. 1–5. [Google Scholar]
  4. Sharma, P.; Naushin, N.; Rohila, S.; Tiwari, A. Magnesium Containing High Entropy Alloys. In Magnesium Alloys Structure and Properties; IntechOpen: Rijeka, Croatia, 2022. [Google Scholar]
  5. Gupta, A.K.; Mallik, B.; Roy, D. Structure Property Correlation of In Situ Reinforced Al–Based Metal Matrix Composite via Stir Casting. Mater. Perform. Charact. 2020, 9, 20190038. [Google Scholar] [CrossRef]
  6. Roy, D.; Gupta, A.K.; Alam, S.; Srikanth, S.; Jha, B.K. Enhancement of Properties of Micro-Alloyed Low-Carbon Ni-Added Steel by Thermomechanical Treatment. J. Mater. Eng. Perform. 2020, 29, 7952–7963. [Google Scholar] [CrossRef]
  7. Kang, B.; Lee, J.; Ryu, H.J.; Hong, S.H. Ultra-High Strength WNbMoTaV High-Entropy Alloys with Fine Grain Structure Fabricated by Powder Metallurgical Process. Mater. Sci. Eng. A 2018, 712, 616–624. [Google Scholar] [CrossRef]
  8. Roy, A.; Hussain, A.; Sharma, P.; Balasubramanian, G.; Taufique, M.F.N.; Devanathan, R.; Singh, P.; Johnson, D.D. Rapid Discovery of High Hardness Multi-Principal-Element Alloys Using a Generative Adversarial Network Model. Acta Mater. 2023, 257, 119177. [Google Scholar] [CrossRef]
  9. Sharma, P.; Das, C.; Sreeramagiri, P.; Balasubramanian, G. Additively Manufactured Lightweight and Hard High-Entropy Alloys by Thermally Activated Solvent Extraction. High Entropy Alloys Mater. 2024, 2, 41–47. [Google Scholar] [CrossRef]
  10. Sharma, P.; Singh, P.; Balasubramanian, G. Engineering Phonon Transport through Cation Disorder in Dimensionally Constricted High Entropy MXene. Carbon 2024, 223, 119015. [Google Scholar] [CrossRef]
  11. Sharma, P.; Balasubramanian, G. Electronic and Lattice Distortions Induce Elastic Softening in Refractory Multicomponent Borides. Chem. Mater. 2023, 35, 7511–7520. [Google Scholar] [CrossRef] [PubMed]
  12. Singh, P.; Acemi, C.; Kuchibhotla, A.; Vela, B.; Sharma, P.; Zhang, W.; Mason, P.; Balasubramanian, G.; Karaman, I.; Arroyave, R.; et al. Alloying Effects on the Transport Properties of Refractory High-Entropy Alloys. Acta Mater. 2024, 276, 120032. [Google Scholar] [CrossRef]
  13. Singh, P.; Vela, B.; Ouyang, G.; Argibay, N.; Cui, J.; Arroyave, R.; Johnson, D.D. A Ductility Metric for Refractory-Based Multi-Principal-Element Alloys. Acta Mater. 2023, 257, 119104. [Google Scholar] [CrossRef]
  14. Singh, P.; Sharma, A.; Smirnov, A.V.; Diallo, M.S.; Ray, P.K.; Balasubramanian, G.; Johnson, D.D. Design of High-Strength Refractory Complex Solid-Solution Alloys. NPJ Comput. Mater. 2018, 4, 16. [Google Scholar] [CrossRef]
  15. Singh, P.; Trehern, W.; Vela, B.; Sharma, P.; Kirk, T.; Pei, Z.; Arroyave, R.; Gao, M.C.; Johnson, D.D. Understanding the Effect of Refractory Metal Chemistry on the Stacking Fault Energy and Mechanical Property of Cantor-Based Multi-Principal Element Alloys. Int. J. Plast. 2024, 179, 104020. [Google Scholar] [CrossRef]
  16. Sharma, P. Synthesis of Superhard Multi-Principal Element Alloy Coatings by Spark Hardening. 2021. Available online: https://www.researchgate.net/publication/378182272_Synthesis_of_Superhard_Multi-Principal_Element_Alloy_Coatings_by_Spark_Hardening (accessed on 18 July 2024).
  17. Yeh, J.-W.; Chen, S.-K.; Lin, S.-J.; Gan, J.-Y.; Chin, T.-S.; Shun, T.-T.; Tsau, C.-H.; Chang, S.-Y. Nanostructured High-Entropy Alloys with Multiple Principal Elements: Novel Alloy Design Concepts and Outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  18. Cantor, B.; Chang, I.T.H.; Knight, P.; Vincent, A.J.B. Microstructural Development in Equiatomic Multicomponent Alloys. Mater. Sci. Eng. A 2004, 375–377, 213–218. [Google Scholar] [CrossRef]
  19. Roy, D.; Pal, S.; Tiwary, C.S.; Gupta, A.K.; Babu, P.N.; Mitra, R. Stable Nanocrystalline Structure Attainment and Strength Enhancement of Cu Base Alloy Using Bi-Modal Distributed Tungsten Dispersoids. Philos. Mag. 2022, 102, 189–209. [Google Scholar] [CrossRef]
  20. Chakraborty, S.; Gupta, A.K.; Roy, D.; Basumallick, A. Studies on nano-metal dispersed Cu-Cr matrix composite. Mater. Lett. 2019, 257, 126739. [Google Scholar] [CrossRef]
  21. Roy, D.; Chakraborty, S.; Gupta, A.; Mallick, A.B.; Koch, C.C. Synergistic Effect of Nb and Zr Addition in Thermal Stabilization of Nano-Crystalline Cu Synthesized by Ball Milling. Mater. Lett. 2020, 271, 127780. [Google Scholar] [CrossRef]
  22. Roy, D.; Chakraborty, S.; Gupta, A.K.; BasuMallick, A.; Scattergood, R.O.; Koch, C.C. Synergistic Effect of Nb and Zr Additions on the Structure-Property Relationships of Nanocrystalline Cu Processed by Mechanical Alloying and Hot Pressing. J. Alloys Compd. 2021, 854, 157174. [Google Scholar] [CrossRef]
  23. Chakraborty, S.; Gupta, A.K.; Roy, D.; Mallick, A.B. Nanomechanical Properties of Mechanically Alloyed and Spark Plasma Sintered W-Nanoparticulate Dispersed Cu-Nb Alloys. Mater. Lett. 2020, 274, 128004. [Google Scholar] [CrossRef]
  24. Sharma, P.; Gandhi, P.M.; Chintersingh, K.-L.; Schoenitz, M.; Dreizin, E.L.; Liou, S.-C.; Balasubramanian, G. Accelerated Intermetallic Phase Amorphization in a Mg-Based High-Entropy Alloy Powder. J. Magnes. Alloys 2024, 12, 1792–1798. [Google Scholar] [CrossRef]
  25. Koch, C.C.; Youssef, K.M.; Scattergood, R.O.; Murty, K.L. Breakthroughs in Optimization of Mechanical Properties of Nanostructured Metals and Alloys. Adv. Eng. Mater. 2005, 7, 787–794. [Google Scholar] [CrossRef]
  26. Weissmüller, J.; Löffler, J.; Kleber, M. Atomic Structure of Nanocrystalline Metals Studied by Diffraction Techniques and EXAFS. Nanostruct. Mater. 1995, 6, 105–114. [Google Scholar] [CrossRef]
  27. Günther, B.; Kumpmann, A.; Kunze, H.-D. Secondary Recrystallization Effects in Nanostructured Elemental Metals. Scr. Metall. Et. Mater. 1992, 27, 833–838. [Google Scholar] [CrossRef]
  28. Olevsky, E.A.; Bradbury, W.L.; Haines, C.D.; Martin, D.G.; Kapoor, D. Fundamental Aspects of Spark Plasma Sintering: I. Experimental Analysis of Scalability. J. Am. Ceram. Soc. 2012, 95, 2406–2413. [Google Scholar] [CrossRef]
  29. Achard, F.C. Recherches sur les Propriétés des Alliages Métalliques; Imprimé Chez George Jacques Decker & Fils: Berlin, Germany, 1982. [Google Scholar]
  30. Senkov, O.N.; Wilks, G.B.; Miracle, D.B.; Chuang, C.P.; Liaw, P.K. Refractory high-entropy alloys. Intermetallics 2010, 18, 1758–1765. [Google Scholar] [CrossRef]
  31. Wu, Z.; Bei, H.; Otto, F.; Pharr, G.M.; George, E.P. Recovery, Recrystallization, Grain Growth and Phase Stability of a Family of FCC-Structured Multi-Component Equiatomic Solid Solution Alloys. Intermetallics 2014, 46, 131–140. [Google Scholar] [CrossRef]
  32. Gao, M.C.; Zhang, B.; Guo, S.M.; Qiao, J.W.; Hawk, J.A. High-Entropy Alloys in Hexagonal Close-Packed Structure. Metall. Mater. Trans. A 2016, 47, 3322–3332. [Google Scholar] [CrossRef]
  33. Zhao, Y.J.; Qiao, J.W.; Ma, S.G.; Gao, M.C.; Yang, H.J.; Chen, M.W.; Zhang, Y. A Hexagonal Close-Packed High-Entropy Alloy: The Effect of Entropy. Mater. Des. 2016, 96, 10–15. [Google Scholar] [CrossRef]
  34. Gludovatz, B.; Hohenwarter, A.; Thurston, K.V.S.; Bei, H.; Wu, Z.; George, E.P.; Ritchie, R.O. Exceptional Damage-Tolerance of a Medium-Entropy Alloy CrCoNi at Cryogenic Temperatures. Nat. Commun. 2016, 7, 10602. [Google Scholar] [CrossRef]
  35. Hemphill, M.A.; Yuan, T.; Wang, G.Y.; Yeh, J.W.; Tsai, C.W.; Chuang, A.; Liaw, P.K. Fatigue behavior of Al0.5CoCrCuFeNi high entropy alloys. Acta Mater. 2012, 60, 5723–5734. [Google Scholar] [CrossRef]
  36. Senkov, O.N.; Senkova, S.V.; Dimiduk, D.M.; Woodward, C.; Miracle, D.B. Oxidation Behavior of a Refractory NbCrMo0.5Ta0.5TiZr Alloy. J. Mater. Sci. 2012, 47, 6522–6534. [Google Scholar] [CrossRef]
  37. Kao, Y.-F.; Chen, S.-K.; Chen, T.-J.; Chu, P.-C.; Yeh, J.-W.; Lin, S.-J. Electrical, Magnetic, and Hall Properties of AlxCoCrFeNi High-Entropy Alloys. J. Alloys Compd. 2011, 509, 1607–1614. [Google Scholar] [CrossRef]
  38. Liaw, P.K.; Egami, T.; Zhang, C.; Zhang, F.; Zhang, Y. Radiation Behavior of High-Entropy Alloys for Advanced Reactors; Final Report; U.S. Department of Energy: Idaho Falls, ID, USA, 2015. [Google Scholar]
  39. Lu, C.; Niu, L.; Chen, N.; Jin, K.; Yang, T.; Xiu, P.; Zhang, Y.; Gao, F.; Bei, H.; Shi, S.; et al. Enhancing Radiation Tolerance by Controlling Defect Mobility and Migration Pathways in Multicomponent Single-Phase Alloys. Nat. Commun. 2016, 7, 13564. [Google Scholar] [CrossRef]
  40. Zhao, Z.K.; Li, J.C.; Jiang, Q. Hypereutectic Al-Si Binary Alloys Prepared by Melt Spinning Method. Adv. Eng. Mater. 2004, 6, 303–306. [Google Scholar] [CrossRef]
  41. Senkov, O.N.; Miller, J.D.; Miracle, D.B.; Woodward, C. Accelerated Exploration of Multi-Principal Element Alloys with Solid Solution Phases. Nat. Commun. 2015, 6, 6529. [Google Scholar] [CrossRef]
  42. Troparevsky, M.C.; Morris, J.R.; Daene, M.; Wang, Y.; Lupini, A.R.; Stocks, G.M. Beyond Atomic Sizes and Hume-Rothery Rules: Understanding and Predicting High-Entropy Alloys. JOM 2015, 67, 2350–2363. [Google Scholar] [CrossRef]
  43. Miracle, D.B.; Senkov, O.N. A Critical Review of High Entropy Alloys and Related Concepts. Acta Mater. 2017, 122, 448–511. [Google Scholar] [CrossRef]
  44. Yeh, J.W.; Chen, Y.L.; Lin, S.J.; Chen, S.K. High-Entropy Alloys—A New Era of Exploitation. Mater. Sci. Forum 2007, 560, 1–9. [Google Scholar] [CrossRef]
  45. Gwalani, B.; Gorsse, S.; Choudhuri, D.; Styles, M.; Zheng, Y.; Mishra, R.S.; Banerjee, R. Modifying Transformation Pathways in High Entropy Alloys or Complex Concentrated Alloys via Thermo-Mechanical Processing. Acta Mater. 2018, 153, 169–185. [Google Scholar] [CrossRef]
  46. Singh, S.; Wanderka, N.; Kiefer, K.; Siemensmeyer, K.; Banhart, J. Effect of Decomposition of the Cr–Fe–Co Rich Phase of AlCoCrCuFeNi High Entropy Alloy on Magnetic Properties. Ultramicroscopy 2011, 111, 619–622. [Google Scholar] [CrossRef]
  47. Yeh, J.-W. Alloy Design Strategies and Future Trends in High-Entropy Alloys. JOM 2013, 65, 1759–1771. [Google Scholar] [CrossRef]
  48. Fultz, B. Vibrational Thermodynamics of Materials. Prog. Mater. Sci. 2010, 55, 247–352. [Google Scholar] [CrossRef]
  49. Srivastava, G.P. The Physics of Phonons; Routledge: New York, NY, USA, 2019; ISBN 9780203736241. [Google Scholar]
  50. Jin, W.; Sharma, P.; Singh, P.; Kundu, A.; Balasubramanian, G.; Chan, H.M. Solid State Reduction Driven Synthesis of Mn Containing Multi-Principal Component Alloys. Metall. Mater. Trans. A 2024, 1–10. [Google Scholar] [CrossRef]
  51. Yeh, J.-W. Recent progress in high-entropy alloys. Ann. Chim. Sci. Mater. 2006, 31, 633–648. [Google Scholar] [CrossRef]
  52. Miracle, D.; Miller, J.; Senkov, O.; Woodward, C.; Uchic, M.; Tiley, J. Exploration and Development of High Entropy Alloys for Structural Applications. Entropy 2014, 16, 494–525. [Google Scholar] [CrossRef]
  53. Gao, Y.F.; Zhang, W.; Shi, P.J.; Ren, W.L.; Zhong, Y.B. A Mechanistic Interpretation of the Strength-Ductility Trade-off and Synergy in Lamellar Microstructures. Mater. Today Adv. 2020, 8, 100103. [Google Scholar] [CrossRef]
  54. Joseph, J.; Stanford, N.; Hodgson, P.; Fabijanic, D.M. Understanding the Mechanical Behaviour and the Large Strength/Ductility Differences between FCC and BCC AlxCoCrFeNi High Entropy Alloys. J. Alloys Compd. 2017, 726, 885–895. [Google Scholar] [CrossRef]
  55. Lim, K.R.; Lee, K.S.; Lee, J.S.; Kim, J.Y.; Chang, H.J.; Na, Y.S. Dual-Phase High-Entropy Alloys for High-Temperature Structural Applications. J. Alloys Compd. 2017, 728, 1235–1238. [Google Scholar] [CrossRef]
  56. George, E.P.; Raabe, D.; Ritchie, R.O. High-Entropy Alloys. Nat. Rev. Mater. 2019, 4, 515–534. [Google Scholar] [CrossRef]
  57. Li, W.; Xie, D.; Li, D.; Zhang, Y.; Gao, Y.; Liaw, P.K. Mechanical Behavior of High-Entropy Alloys. Prog. Mater. Sci. 2021, 118, 100777. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Zuo, T.T.; Tang, Z.; Gao, M.C.; Dahmen, K.A.; Liaw, P.K.; Lu, Z.P. Microstructures and Properties of High-Entropy Alloys. Prog. Mater. Sci. 2014, 61, 1–93. [Google Scholar] [CrossRef]
  59. Hsu, C.-Y.; Juan, C.-C.; Wang, W.-R.; Sheu, T.-S.; Yeh, J.-W.; Chen, S.-K. On the Superior Hot Hardness and Softening Resistance of AlCoCrxFeMo0.5Ni High-Entropy Alloys. Mater. Sci. Eng. A 2011, 528, 3581–3588. [Google Scholar] [CrossRef]
  60. Zhang, W.; Chabok, A.; Kooi, B.J.; Pei, Y. Additive Manufactured High Entropy Alloys: A Review of the Microstructure and Properties. Mater. Des. 2022, 220, 110875. [Google Scholar] [CrossRef]
  61. Lewandowski, J.J.; Seifi, M. Metal Additive Manufacturing: A Review of Mechanical Properties. Annu. Rev. Mater. Res. 2016, 46, 151–186. [Google Scholar] [CrossRef]
  62. Moghaddam, A.O.; Shaburova, N.A.; Samodurova, M.N.; Abdollahzadeh, A.; Trofimov, E.A. Additive Manufacturing of High Entropy Alloys: A Practical Review. J. Mater. Sci. Technol. 2021, 77, 131–162. [Google Scholar] [CrossRef]
  63. Herzog, D.; Seyda, V.; Wycisk, E.; Emmelmann, C. Additive manufacturing of metals. Acta Mater. 2016, 117, 371–392. [Google Scholar] [CrossRef]
  64. Xiang, S.; Luan, H.; Wu, J.; Yao, K.-F.; Li, J.; Liu, X.; Tian, Y.; Mao, W.; Bai, H.; Le, G.; et al. Microstructures and Mechanical Properties of CrMnFeCoNi High Entropy Alloys Fabricated Using Laser Metal Deposition Technique. J. Alloys Compd. 2019, 773, 387–392. [Google Scholar] [CrossRef]
  65. Brif, Y.; Thomas, M.; Todd, I. The Use of High-Entropy Alloys in Additive Manufacturing. Scr. Mater. 2015, 99, 93–96. [Google Scholar] [CrossRef]
  66. Nartu, M.S.K.K.Y.; Alam, T.; Dasari, S.; Mantri, S.A.; Gorsse, S.; Siller, H.; Dahotre, N.; Banerjee, R. Enhanced Tensile Yield Strength in Laser Additively Manufactured Al0.3CoCrFeNi High Entropy Alloy. Materialia 2020, 9, 100522. [Google Scholar] [CrossRef]
  67. Tong, Z.; Ren, X.; Jiao, J.; Zhou, W.; Ren, Y.; Ye, Y.; Larson, E.A.; Gu, J. Laser Additive Manufacturing of FeCrCoMnNi High-Entropy Alloy: Effect of Heat Treatment on Microstructure, Residual Stress and Mechanical Property. J. Alloys Compd. 2019, 785, 1144–1159. [Google Scholar] [CrossRef]
  68. Guan, S.; Wan, D.; Solberg, K.; Berto, F.; Welo, T.; Yue, T.M.; Chan, K.C. Additively Manufactured CrMnFeCoNi/AlCoCrFeNiTi0.5 Laminated High-Entropy Alloy with Enhanced Strength-Plasticity Synergy. Scr. Mater. 2020, 183, 133–138. [Google Scholar] [CrossRef]
  69. Kim, J.G.; Park, J.M.; Seol, J.B.; Choe, J.; Yu, J.-H.; Yang, S.; Kim, H.S. Nano-Scale Solute Heterogeneities in the Ultrastrong Selectively Laser Melted Carbon-Doped CoCrFeMnNi Alloy. Mater. Sci. Eng. A 2020, 773, 138726. [Google Scholar] [CrossRef]
  70. Kim, Y.-K.; Choe, J.; Lee, K.-A. Selective Laser Melted Equiatomic CoCrFeMnNi High-Entropy Alloy: Microstructure, Anisotropic Mechanical Response, and Multiple Strengthening Mechanism. J. Alloys Compd. 2019, 805, 680–691. [Google Scholar] [CrossRef]
  71. Zhou, P.F.; Xiao, D.H.; Wu, Z.; Ou, X.Q. Al0.5FeCoCrNi High Entropy Alloy Prepared by Selective Laser Melting with Gas-Atomized Pre-Alloy Powders. Mater. Sci. Eng. A 2019, 739, 86–89. [Google Scholar] [CrossRef]
  72. Yamanaka, K.; Shiratori, H.; Mori, M.; Omura, K.; Fujieda, T.; Kuwabara, K.; Chiba, A. Corrosion Mechanism of an Equimolar AlCoCrFeNi High-Entropy Alloy Additively Manufactured by Electron Beam Melting. NPJ Mater. Degrad. 2020, 4, 24. [Google Scholar] [CrossRef]
  73. Sun, Z.; Tan, X.; Wang, C.; Descoins, M.; Mangelinck, D.; Tor, S.B.; Jägle, E.A.; Zaefferer, S.; Raabe, D. Reducing Hot Tearing by Grain Boundary Segregation Engineering in Additive Manufacturing: Example of an AlxCoCrFeNi High-Entropy Alloy. Acta Mater. 2021, 204, 116505. [Google Scholar] [CrossRef]
  74. Sun, K.; Peng, W.; Yang, L.; Fang, L. Effect of SLM Processing Parameters on Microstructures and Mechanical Properties of Al0.5CoCrFeNi High Entropy Alloys. Metals 2020, 10, 292. [Google Scholar] [CrossRef]
  75. Karimi, J.; Ma, P.; Jia, Y.D.; Prashanth, K.G. Linear Patterning of High Entropy Alloy by Additive Manufacturing. Manuf. Lett. 2020, 24, 9–13. [Google Scholar] [CrossRef]
  76. Karlsson, D.; Marshal, A.; Johansson, F.; Schuisky, M.; Sahlberg, M.; Schneider, J.M.; Jansson, U. Elemental Segregation in an AlCoCrFeNi High-Entropy Alloy—A Comparison between Selective Laser Melting and Induction Melting. J. Alloys Compd. 2019, 784, 195–203. [Google Scholar] [CrossRef]
  77. Kuwabara, K.; Shiratori, H.; Fujieda, T.; Yamanaka, K.; Koizumi, Y.; Chiba, A. Mechanical and Corrosion Properties of AlCoCrFeNi High-Entropy Alloy Fabricated with Selective Electron Beam Melting. Addit. Manuf. 2018, 23, 264–271. [Google Scholar] [CrossRef]
  78. Lin, W.-C.; Chang, Y.-J.; Hsu, T.-H.; Gorsse, S.; Sun, F.; Furuhara, T.; Yeh, A.-C. Microstructure and Tensile Property of a Precipitation Strengthened High Entropy Alloy Processed by Selective Laser Melting and Post Heat Treatment. Addit. Manuf. 2020, 36, 101601. [Google Scholar] [CrossRef]
  79. Liu, S.; Wan, D.; Guan, S.; Fu, Y.; Ren, X.; Zhang, Z.; He, J. Microstructure and Nanomechanical Behavior of an Additively Manufactured (CrCoNiFe)94Ti2Al4 High-Entropy Alloy. Mater. Sci. Eng. A 2021, 823, 141737. [Google Scholar] [CrossRef]
  80. Yan, Y.; Song, W.-D.; Li, K.-F.; Zhao, K.; Sun, T.-T.; Song, K.-K.; Gong, J.-H.; Hu, L.-N. Microstructural Features and Mechanical Behaviors of Al0.5Cr0.8CoFeNi2.5V0.2 High-Entropy Alloys Fabricated by Selective Laser Melting Technique. Acta Met. Sin. 2022, 35, 1591–1606. [Google Scholar] [CrossRef]
  81. Yao, H.; Tan, Z.; He, D.; Zhou, Z.; Zhou, Z.; Xue, Y.; Cui, L.; Chen, L.; Wang, G.; Yang, Y. High Strength and Ductility AlCrFeNiV High Entropy Alloy with Hierarchically Heterogeneous Microstructure Prepared by Selective Laser Melting. J. Alloys Compd. 2020, 813, 152196. [Google Scholar] [CrossRef]
  82. Lin, D.; Xu, L.; Jing, H.; Han, Y.; Zhao, L.; Minami, F. Effects of Annealing on the Structure and Mechanical Properties of FeCoCrNi High-Entropy Alloy Fabricated via Selective Laser Melting. Addit. Manuf. 2020, 32, 101058. [Google Scholar] [CrossRef]
  83. Ren, Y.; Liang, L.; Shan, Q.; Cai, A.; Du, J.; Huang, Q.; Liu, S.; Yang, X.; Tian, Y.; Wu, H. Effect of Volumetric Energy Density on Microstructure and Tribological Properties of FeCoNiCuAl High-Entropy Alloy Produced by Laser Powder Bed Fusion. Virtual Phys. Prototyp. 2020, 15, 543–554. [Google Scholar] [CrossRef]
  84. Yang, X.; Zhou, Y.; Xi, S.; Chen, Z.; Wei, P.; He, C.; Li, T.; Gao, Y.; Wu, H. Additively Manufactured Fine Grained Ni6Cr4WFe9Ti High Entropy Alloys with High Strength and Ductility. Mater. Sci. Eng. A 2019, 767, 138394. [Google Scholar] [CrossRef]
  85. Wang, P.; Huang, P.; Ng, F.L.; Sin, W.J.; Lu, S.; Nai, M.L.S.; Dong, Z.; Wei, J. Additively Manufactured CoCrFeNiMn High-Entropy Alloy via Pre-Alloyed Powder. Mater. Des. 2019, 168, 107576. [Google Scholar] [CrossRef]
  86. Ng, C.K.; Bai, K.; Wuu, D.; Lau, K.B.; Lee, J.J.; Cheong, A.K.H.; Wei, F.; Cheng, B.; Wang, P.; Tan, D.C.C.; et al. Additive Manufacturing of High-Strength and Ductile High Entropy Alloy CoCrFeNiW0.2 Composites via Laser Powder Bed Fusion and Post-Annealing. J. Alloys Compd. 2022, 906, 164288. [Google Scholar] [CrossRef]
  87. Li, R.; Niu, P.; Yuan, T.; Cao, P.; Chen, C.; Zhou, K. Selective Laser Melting of an Equiatomic CoCrFeMnNi High-Entropy Alloy: Processability, Non-Equilibrium Microstructure and Mechanical Property. J. Alloys Compd. 2018, 746, 125–134. [Google Scholar] [CrossRef]
  88. Zhu, Z.G.; Nguyen, Q.B.; Ng, F.L.; An, X.H.; Liao, X.Z.; Liaw, P.K.; Nai, S.M.L.; Wei, J. Hierarchical Microstructure and Strengthening Mechanisms of a CoCrFeNiMn High Entropy Alloy Additively Manufactured by Selective Laser Melting. Scr. Mater. 2018, 154, 20–24. [Google Scholar] [CrossRef]
  89. Yang, X.; Ge, Y.; Lehtonen, J.; Hannula, S.-P. Hierarchical Microstructure of Laser Powder Bed Fusion Produced Face-Centered-Cubic-Structured Equiatomic CrFeNiMn Multicomponent Alloy. Materials 2020, 13, 4498. [Google Scholar] [CrossRef] [PubMed]
  90. Zhang, H.; Zhao, Y.; Huang, S.; Zhu, S.; Wang, F.; Li, D. Manufacturing and Analysis of High-Performance Refractory High-Entropy Alloy via Selective Laser Melting (SLM). Materials 2019, 12, 720. [Google Scholar] [CrossRef] [PubMed]
  91. Fujieda, T.; Shiratori, H.; Kuwabara, K.; Kato, T.; Yamanaka, K.; Koizumi, Y.; Chiba, A. First Demonstration of Promising Selective Electron Beam Melting Method for Utilizing High-Entropy Alloys as Engineering Materials. Mater. Lett. 2015, 159, 12–15. [Google Scholar] [CrossRef]
  92. Osintsev, K.; Konovalov, S.; Gromov, V.; Panchenko, I.; Ivanov, Y. Microstructural and Mechanical Characterisation of Non-Equiatomic Al2.1Co0.3Cr0.5FeNi2.1 High-Entropy Alloy Fabricated via Wire-Arc Additive Manufacturing. Philos. Mag. Lett. 2021, 101, 353–359. [Google Scholar] [CrossRef]
  93. Karlsson, D.; Lindwall, G.; Lundbäck, A.; Amnebrink, M.; Boström, M.; Riekehr, L.; Schuisky, M.; Sahlberg, M.; Jansson, U. Binder Jetting of the AlCoCrFeNi Alloy. Addit. Manuf. 2019, 27, 72–79. [Google Scholar] [CrossRef]
  94. Luo, S.; Gao, P.; Yu, H.; Yang, J.; Wang, Z.; Zeng, X. Selective Laser Melting of an Equiatomic AlCrCuFeNi High-Entropy Alloy: Processability, Non-Equilibrium Microstructure and Mechanical Behavior. J. Alloys Compd. 2019, 771, 387–397. [Google Scholar] [CrossRef]
  95. Malatji, N.; Lengopeng, T.; Pityana, S.; Popoola, A. Microstructural, Mechanical and Electrochemical Properties of AlCrFeCuNiWx High Entropy Alloys. J. Mater. Res. Technol. 2021, 11, 1594–1603. [Google Scholar] [CrossRef]
  96. Zhou, K.; Li, J.; Wang, L.; Yang, H.; Wang, Z.; Wang, J. Direct laser deposited bulk CoCrFeNiNbx high entropy alloys. Intermetallics 2019, 114, 106592. [Google Scholar] [CrossRef]
  97. Park, J.M.; Asghari-Rad, P.; Zargaran, A.; Bae, J.W.; Moon, J.; Kwon, H.; Choe, J.; Yang, S.; Yu, J.-H.; Kim, H.S. Nano-Scale Heterogeneity-Driven Metastability Engineering in Ferrous Medium-Entropy Alloy Induced by Additive Manufacturing. Acta Mater. 2021, 221, 117426. [Google Scholar] [CrossRef]
  98. Yang, F.; Wang, L.; Wang, Z.; Wu, Q.; Zhou, K.; Lin, X.; Huang, W. Ultra Strong and Ductile Eutectic High Entropy Alloy Fabricated by Selective Laser Melting. J. Mater. Sci. Technol. 2022, 106, 128–132. [Google Scholar] [CrossRef]
  99. Zhu, Y.; Zhou, S.; Xiong, Z.; Liang, Y.-J.; Xue, Y.; Wang, L. Enabling Stronger Eutectic High-Entropy Alloys with Larger Ductility by 3D Printed Directional Lamellae. Addit. Manuf. 2021, 39, 101901. [Google Scholar] [CrossRef]
  100. Vikram, R.J.; Murty, B.S.; Fabijanic, D.; Suwas, S. Insights into Micro-Mechanical Response and Texture of the Additively Manufactured Eutectic High Entropy Alloy AlCoCrFeNi2.1. J. Alloys Compd. 2020, 827, 154034. [Google Scholar] [CrossRef]
  101. Zhou, K.; Li, J.; Wu, Q.; Zhang, Z.; Wang, Z.; Wang, J. Remelting Induced Fully-Equiaxed Microstructures with Anomalous Eutectics in the Additive Manufactured Ni32Co30Cr10Fe10Al18 Eutectic High-Entropy Alloy. Scr. Mater. 2021, 201, 113952. [Google Scholar] [CrossRef]
  102. Joseph, J.; Hodgson, P.; Jarvis, T.; Wu, X.; Stanford, N.; Fabijanic, D.M. Effect of Hot Isostatic Pressing on the Microstructure and Mechanical Properties of Additive Manufactured AlxCoCrFeNi High Entropy Alloys. Mater. Sci. Eng. A 2018, 733, 59–70. [Google Scholar] [CrossRef]
  103. Joseph, J.; Stanford, N.; Hodgson, P.; Fabijanic, D.M. Tension/Compression Asymmetry in Additive Manufactured Face Centered Cubic High Entropy Alloy. Scr. Mater. 2017, 129, 30–34. [Google Scholar] [CrossRef]
  104. Theeda, S.; Jagdale, S.H.; Ravichander, B.B.; Kumar, G. Optimization of Process Parameters in Laser Powder Bed Fusion of SS 316L Parts Using Artificial Neural Networks. Metals 2023, 13, 842. [Google Scholar] [CrossRef]
  105. Sola, A.; Nouri, A. Microstructural Porosity in Additive Manufacturing: The Formation and Detection of Pores in Metal Parts Fabricated by Powder Bed Fusion. J. Adv. Manuf. Proc. 2019, 1, e10021. [Google Scholar] [CrossRef]
  106. Islam, S.F.; Hawkins, S.M.; Meyer, J.L.L.; Sharman, A.R.C. Evaluation of Different Particle Size Distribution and Morphology Characterization Techniques. Addit. Manuf. Lett. 2022, 3, 100077. [Google Scholar] [CrossRef]
  107. Sofia, D.; Macrì, D.; Barletta, D.; Lettieri, P.; Poletto, M. Use of Titania Powders in the Laser Sintering Process: Link between Process Conditions and Product Mechanical Properties. Powder Technol. 2021, 381, 181–188. [Google Scholar] [CrossRef]
  108. Yuan, Z.; Chen, A.; Xue, L.; Ma, R.; Bai, J.; Zheng, J.; Bai, X.; Zheng, J.; Mu, S.; Zhang, L. Effect of Hot Isostatic Pressing and Heat. Treatment on the Evolution of Precipitated Phase and Mechanical Properties of GH3230 Superalloy Fabricated via Selective Laser Melting. J. Mater. Res. Technol. 2024, 30, 507–519. [Google Scholar] [CrossRef]
  109. Guo, S.; Ng, C.; Lu, J.; Liu, C.T. Effect of Valence Electron Concentration on Stability of Fcc or Bcc Phase in High Entropy Alloys. J. Appl. Phys. 2011, 109, 103505. [Google Scholar] [CrossRef]
  110. Vida, Á.; Chinh, N.Q.; Lendvai, J.; Heczel, A.; Varga, L.K. Microstructures and Transition from Brittle to Ductile Behavior of NiFeCrMoW High Entropy Alloys. Mater. Lett. 2017, 195, 14–17. [Google Scholar] [CrossRef]
  111. Chen, R.; Qin, G.; Zheng, H.; Wang, L.; Su, Y.; Chiu, Y.; Ding, H.; Guo, J.; Fu, H. Composition Design of High Entropy Alloys Using the Valence Electron Concentration to Balance Strength and Ductility. Acta Mater. 2018, 144, 129–137. [Google Scholar] [CrossRef]
  112. Zhang, Y.; Ma, S.G.; Qiao, J.W. Morphology Transition from Dendrites to Equiaxed Grains for AlCoCrFeNi High-Entropy Alloys by Copper Mold Casting and Bridgman Solidification. Metall. Mater. Trans. A 2012, 43, 2625–2630. [Google Scholar] [CrossRef]
  113. Hanna, J.A.; Baker, I.; Wittmann, M.W.; Munroe, P.R. A New High-Strength Spinodal Alloy. J. Mater. Res. 2005, 20, 791–795. [Google Scholar] [CrossRef]
  114. Wang, W.-R.; Wang, W.-L.; Wang, S.-C.; Tsai, Y.-C.; Lai, C.-H.; Yeh, J.-W. Effects of Al Addition on the Microstructure and Mechanical Property of AlxCoCrFeNi High-Entropy Alloys. Intermetallics 2012, 26, 44–51. [Google Scholar] [CrossRef]
  115. Wang, W.-R.; Wang, W.-L.; Yeh, J.-W. Phases, Microstructure and Mechanical Properties of AlxCoCrFeNi High-Entropy Alloys at Elevated Temperatures. J. Alloys Compd. 2014, 589, 143–152. [Google Scholar] [CrossRef]
  116. Yang, T.; Xia, S.; Liu, S.; Wang, C.; Liu, S.; Zhang, Y.; Xue, J.; Yan, S.; Wang, Y. Effects of AL Addition on Microstructure and Mechanical Properties of Al CoCrFeNi High-Entropy Alloy. Mater. Sci. Eng. A 2015, 648, 15–22. [Google Scholar] [CrossRef]
  117. Zhang, F.X.; Zhao, S.; Jin, K.; Bei, H.; Popov, D.; Park, C.; Neuefeind, J.C.; Weber, W.J.; Zhang, Y. Pressure-Induced Fcc to Hcp Phase Transition in Ni-Based High Entropy Solid Solution Alloys. Appl. Phys. Lett. 2017, 110, 011902. [Google Scholar] [CrossRef]
  118. Yang, J.X.; Zhao, H.L.; Gong, H.R.; Song, M.; Ren, Q.Q. Proposed Mechanism of HCP → FCC Phase Transition in Titianium through First Principles Calculation and Experiments. Sci. Rep. 2018, 8, 1992. [Google Scholar] [CrossRef]
  119. Ma, D.; Grabowski, B.; Körmann, F.; Neugebauer, J.; Raabe, D. Ab Initio Thermodynamics of the CoCrFeMnNi High Entropy Alloy: Importance of Entropy Contributions beyond the Configurational One. Acta Mater. 2015, 100, 90–97. [Google Scholar] [CrossRef]
  120. Weng, S.; Han, W.; Chen, G.; Fu, T. Spatially varied stacking fault energy induced low twinning ability in high entropy alloys. J. Appl. Phys. 2024, 135, 205103. [Google Scholar] [CrossRef]
  121. Tsao, T.; Yeh, A.; Kuo, C.; Murakami, H. On The Superior High Temperature Hardness of Precipitation Strengthened High Entropy Ni-Based Alloys. Adv. Eng. Mater. 2017, 19, 1600475. [Google Scholar] [CrossRef]
  122. Reed, R.C.; Tao, T.; Warnken, N. Alloys-By-Design: Application to Nickel-Based Single Crystal Superalloys. Acta Mater. 2009, 57, 5898–5913. [Google Scholar] [CrossRef]
  123. Kawagishi, K.; Yeh, A.; Yokokawa, T.; Kobayashi, T.; Koizumi, Y.; Harada, H. Development of an Oxidation-Resistant High-Strength Sixth-Generation Single-Crystal Superalloy TMS-238. In Superalloys; Wiley: Michigan, MI, USA, 2012; pp. 189–195. [Google Scholar]
  124. Reed, R.C. The Superalloys; Cambridge University Press: Cambridge, UK, 2006; ISBN 9780521859042. [Google Scholar]
  125. Sugui, T.; Minggang, W.; Tang, L.; Benjiang, Q.; Jun, X. Influence of TCP Phase and Its Morphology on Creep Properties of Single Crystal Nickel-Based Superalloys. Mater. Sci. Eng. A 2010, 527, 5444–5451. [Google Scholar] [CrossRef]
  126. López Ríos, M.; Socorro Perdomo, P.P.; Voiculescu, I.; Geanta, V.; Crăciun, V.; Boerasu, I.; Mirza Rosca, J.C. Effects of Nickel Content on the Microstructure, Microhardness and Corrosion Behavior of High-Entropy AlCoCrFeNix Alloys. Sci. Rep. 2020, 10, 21119. [Google Scholar] [CrossRef]
  127. Li, Z.; Pradeep, K.G.; Deng, Y.; Raabe, D.; Tasan, C.C. Metastable High-Entropy Dual-Phase Alloys Overcome the Strength–Ductility Trade-Off. Nature 2016, 534, 227–230. [Google Scholar] [CrossRef]
  128. Zhang, C.; Zhang, F.; Diao, H.; Gao, M.C.; Tang, Z.; Poplawsky, J.D.; Liaw, P.K. Understanding Phase Stability of Al-Co-Cr-Fe-Ni High Entropy Alloys. Mater. Des. 2016, 109, 425–433. [Google Scholar] [CrossRef]
  129. Yasuda, H.Y.; Miyamoto, H.; Cho, K.; Nagase, T. Formation of Ultrafine-Grained Microstructure in Al0.3CoCrFeNi High Entropy Alloys with Grain Boundary Precipitates. Mater. Lett. 2017, 199, 120–123. [Google Scholar] [CrossRef]
  130. He, J.Y.; Liu, W.H.; Wang, H.; Wu, Y.; Liu, X.J.; Nieh, T.G.; Lu, Z.P. Effects of Al Addition on Structural Evolution and Tensile Properties of the FeCoNiCrMn High-Entropy Alloy System. Acta Mater. 2014, 62, 105–113. [Google Scholar] [CrossRef]
  131. Gorsse, S.; Nguyen, M.H.; Senkov, O.N.; Miracle, D.B. Database on the Mechanical Properties of High Entropy Alloys and Complex Concentrated Alloys. Data Brief 2018, 21, 2664–2678. [Google Scholar] [CrossRef]
  132. Gawel, R.; Rogal, Ł.; Grzesik, Z. Behaviour of Al, Co, Cr, Ni-Based High Entropy Alloys under High-Temperature Thermal Shock Oxidising Conditions. Corros. Sci. 2022, 198, 110116. [Google Scholar] [CrossRef]
  133. Ng, C.; Guo, S.; Luan, J.; Shi, S.; Liu, C.T. Entropy-Driven Phase Stability and Slow Diffusion Kinetics in an Al0.5CoCrCuFeNi High Entropy Alloy. Intermetallics 2012, 31, 165–172. [Google Scholar] [CrossRef]
  134. Butler, T.M.; Alfano, J.P.; Martens, R.L.; Weaver, M.L. High-Temperature Oxidation Behavior of Al-Co-Cr-Ni-(Fe or Si) Multicomponent High-Entropy Alloys. JOM 2015, 67, 246–259. [Google Scholar] [CrossRef]
  135. Talkar, S.; Choudhari, A.; Rayar, P. Building Envelope Optimization and Cost-Effective Approach in HVAC to Support Smart Manufacturing. In Proceedings of the International Conference on Intelligent Manufacturing and Automation, Mostar, Bosnia and Herzegovina, 21–24 October 2020; Springer: Singapore, 2020. [Google Scholar]
  136. Sharma, P. Electro-Spark Coating of High Entropy Alloy on Steel and Inconel Superalloy for Surface Hardening and Oxidation Protection. Master’s Dissertation, Indian Institute of Technology Hyderabad, Kandi, India, 2020. [Google Scholar]
  137. Jumaev, E.; Hong, S.H.; Kim, J.T.; Park, H.J.; Kim, Y.S.; Mun, S.C.; Park, J.-Y.; Song, G.; Lee, J.K.; Min, B.H.; et al. Chemical Evolution-Induced Strengthening on AlCoCrNi Dual-Phase High-Entropy Alloy with High Specific Strength. J. Alloys Compd. 2019, 777, 828–834. [Google Scholar] [CrossRef]
  138. Gyawali, S.; Tirumala, R.T.A.; Loh, H.; Andiappan, M.; Bristow, A.D. Photocarrier Recombination Dynamics in Highly Scattering Cu2O Nanocatalyst Clusters. J. Phys. Chem. C 2024, 128, 2003–2011. [Google Scholar] [CrossRef]
  139. Gyawali, S.; Tirumala, R.T.A.; Andiappan, M.; Bristow, A.D. Size- and Shape-Dependent Charge-Carrier Dynamics in Sub-Micron Cuprous Oxide Nanoparticles. In Proceedings of the Frontiers in Optics + Laser Science 2022 (FIO, LS), Rochester, NT, USA, 18–19 October 2022; Optica Publishing Group: Washington, DC, USA, 2022; p. JTu4A.86. [Google Scholar] [CrossRef]
  140. Gyawali, S.; Tirumala, R.T.A.; Andiappan, M.; Bristow, A.D. Carrier Dynamics in Cuprous Oxide-Based Nanoparticles and Heterojunctions. In Proceedings of the Ultrafast Phenomena and Nanophotonics XXVIII, San Francisco, CA, USA, 29–31 January 2024; Betz, M., Elezzabi, A.Y., Eds.; SPIE: New York, NY, USA, 2024; Volume 12884, p. 128840F. [Google Scholar] [CrossRef]
Figure 1. The primary components utilized for AM HEAs are highlighted by clusters of alloying elements [60]. Adapted with permission from ref. [60]. A 2022 Creative Commons Attribution (CC BY) license.
Figure 1. The primary components utilized for AM HEAs are highlighted by clusters of alloying elements [60]. Adapted with permission from ref. [60]. A 2022 Creative Commons Attribution (CC BY) license.
Materials 17 03826 g001
Figure 2. Data on the microhardness of AM HEAs were gathered from Zhang et al. [60]. (The HEAs produced by SLM, LMD, EBM, and WAAM are represented by the solid, top half-solid, open, and bottom half-solid symbols, respectively. Only considering the condition of fabrication). Adapted with permission from ref. [60]. A 2022 Creative Commons Attribution (CC BY) license.
Figure 2. Data on the microhardness of AM HEAs were gathered from Zhang et al. [60]. (The HEAs produced by SLM, LMD, EBM, and WAAM are represented by the solid, top half-solid, open, and bottom half-solid symbols, respectively. Only considering the condition of fabrication). Adapted with permission from ref. [60]. A 2022 Creative Commons Attribution (CC BY) license.
Materials 17 03826 g002
Figure 4. The fracture elongation and compressive strength of several AMed high-energy alloys. The HEAs manufactured by SLM, LMD, EBM, and WAAM are represented by the solid, top half-solid, open, and bottom half-solid symbols, respectively (only considering the as-fabricated state) [60]. Adapted with permission from ref. [60]. A 2022 Creative Commons Attribution (CC BY) license.
Figure 4. The fracture elongation and compressive strength of several AMed high-energy alloys. The HEAs manufactured by SLM, LMD, EBM, and WAAM are represented by the solid, top half-solid, open, and bottom half-solid symbols, respectively (only considering the as-fabricated state) [60]. Adapted with permission from ref. [60]. A 2022 Creative Commons Attribution (CC BY) license.
Materials 17 03826 g004
Figure 5. (a) shows a part made using EBM, (b) an SEM image showcasing the rough surface texture from a side view, and (c) a representation of the top surface of cuboid samples exhibiting swelling and lack of fusion. Swelling is denoted by yellow arrows, while areas of incomplete fusion are highlighted by red arrows [85]. Adapted with permission from ref. [85]. A 2019 Creative Commons Attribution (CC BY) license.
Figure 5. (a) shows a part made using EBM, (b) an SEM image showcasing the rough surface texture from a side view, and (c) a representation of the top surface of cuboid samples exhibiting swelling and lack of fusion. Swelling is denoted by yellow arrows, while areas of incomplete fusion are highlighted by red arrows [85]. Adapted with permission from ref. [85]. A 2019 Creative Commons Attribution (CC BY) license.
Materials 17 03826 g005
Figure 6. (af) OM images of EBM-built specimens with corresponding (a′d′) pore size and (a″d″) pore area distributions. (a,a′,a″) stand for specimens 1 (S1), 2 (S2), 3, and 4, respectively. (c,c′,c″) stand for specimens 3 (S3) and 4 (S4), respectively. Enlarged pictures from (a,d) in (e,f) correspondingly reveal the presence of gas pores and the absence of fusion faults [85]. Adapted with permission from ref. [85]. A 2019 Creative Commons Attribution (CC BY) license.
Figure 6. (af) OM images of EBM-built specimens with corresponding (a′d′) pore size and (a″d″) pore area distributions. (a,a′,a″) stand for specimens 1 (S1), 2 (S2), 3, and 4, respectively. (c,c′,c″) stand for specimens 3 (S3) and 4 (S4), respectively. Enlarged pictures from (a,d) in (e,f) correspondingly reveal the presence of gas pores and the absence of fusion faults [85]. Adapted with permission from ref. [85]. A 2019 Creative Commons Attribution (CC BY) license.
Materials 17 03826 g006
Figure 7. Co-relation of VEC with the crystal structure (BCC or FCC) [109]. Reprinted with permission from ref. [109]. © 2016 AIP Publishing.
Figure 7. Co-relation of VEC with the crystal structure (BCC or FCC) [109]. Reprinted with permission from ref. [109]. © 2016 AIP Publishing.
Materials 17 03826 g007
Figure 8. Variation of tensile properties due to change in crystal structure i.e., BCC to FCC phase [110]. Reprinted with permission from ref. [110]. © 2017 Elsevier.
Figure 8. Variation of tensile properties due to change in crystal structure i.e., BCC to FCC phase [110]. Reprinted with permission from ref. [110]. © 2017 Elsevier.
Materials 17 03826 g008
Figure 9. SEM of as-cast (AlCoCrFeNi)100−xNix, where x is (1/4, 0, 4, 8, 12, and 16) at corresponding low and high magnification [111]. Reprinted with permission from ref. [111]. © 2018 Elsevier.
Figure 9. SEM of as-cast (AlCoCrFeNi)100−xNix, where x is (1/4, 0, 4, 8, 12, and 16) at corresponding low and high magnification [111]. Reprinted with permission from ref. [111]. © 2018 Elsevier.
Materials 17 03826 g009
Figure 10. Scanning Electron Micrograph (SEM) of as-cast (CoCrCuFeNi)100−XMox, where x is (1/4, 0, 4, 8, 12, and 16) at corresponding low and high magnification [111]. Reprinted with permission from ref. [111]. © 2018 Elsevier.
Figure 10. Scanning Electron Micrograph (SEM) of as-cast (CoCrCuFeNi)100−XMox, where x is (1/4, 0, 4, 8, 12, and 16) at corresponding low and high magnification [111]. Reprinted with permission from ref. [111]. © 2018 Elsevier.
Materials 17 03826 g010
Figure 11. Neutron diffraction profiles depicting single phase FCC solid solution for alloys “NiCoCr”, “NiCoCrFe”, and “NiCoCrFePd” [117]. Reprinted with permission from ref. [117]. © 2017 AIP Publishing.
Figure 11. Neutron diffraction profiles depicting single phase FCC solid solution for alloys “NiCoCr”, “NiCoCrFe”, and “NiCoCrFePd” [117]. Reprinted with permission from ref. [117]. © 2017 AIP Publishing.
Materials 17 03826 g011
Figure 12. (ac) shows XRD patterns for NiCoCrFe, NiCoCr, and NiCoCrFePd alloys under high pressure, illustrating the transformation from FCC to HCP structure at pressures of 13.5, 45, and 74 GPa, respectively [117]. The diffraction peaks marked with * are from Au pressure marker. Reprinted with permission from ref. [117]. © 2017 AIP Publishing.
Figure 12. (ac) shows XRD patterns for NiCoCrFe, NiCoCr, and NiCoCrFePd alloys under high pressure, illustrating the transformation from FCC to HCP structure at pressures of 13.5, 45, and 74 GPa, respectively [117]. The diffraction peaks marked with * are from Au pressure marker. Reprinted with permission from ref. [117]. © 2017 AIP Publishing.
Materials 17 03826 g012
Figure 13. (a,b) Displays TEM bright field images with insets showing diffraction spots at the [001] zone axis for the alloys HESA-1 and HESA-2, respectively [121]. Reprinted with permission from ref. [121]. © 2017 John Wiley and Sons.
Figure 13. (a,b) Displays TEM bright field images with insets showing diffraction spots at the [001] zone axis for the alloys HESA-1 and HESA-2, respectively [121]. Reprinted with permission from ref. [121]. © 2017 John Wiley and Sons.
Materials 17 03826 g013
Figure 14. (ad) show SEM-BSE micrographs of samples that underwent recrystallization for 1 h at temperatures of 1073 K, 1173 K, 1273 K, and 1373 K, respectively. These micrographs provide insights into the microstructural evolution during the recrystallization process at different temperatures; (e) displays the SAED pattern obtained from a B2 precipitate phase present in the material. SAED is a technique used to analyze the crystal structure and orientation of localized regions within a sample; (f) presents an EBSD image illustrating the distribution of face-centered cubic (FCC) and B2 phases within the material. Accompanying the EBSD image are pole figures, which represent the crystallographic orientations of the respective phases; (g) shows a BSE micrograph of a single crystal sample that was annealed at 1273 K for 100 h. This micrograph provides information about the microstructural changes induced by the prolonged heat treatment [129]. Reprinted with permission from ref. [129]. © 2017 Elsevier.
Figure 14. (ad) show SEM-BSE micrographs of samples that underwent recrystallization for 1 h at temperatures of 1073 K, 1173 K, 1273 K, and 1373 K, respectively. These micrographs provide insights into the microstructural evolution during the recrystallization process at different temperatures; (e) displays the SAED pattern obtained from a B2 precipitate phase present in the material. SAED is a technique used to analyze the crystal structure and orientation of localized regions within a sample; (f) presents an EBSD image illustrating the distribution of face-centered cubic (FCC) and B2 phases within the material. Accompanying the EBSD image are pole figures, which represent the crystallographic orientations of the respective phases; (g) shows a BSE micrograph of a single crystal sample that was annealed at 1273 K for 100 h. This micrograph provides information about the microstructural changes induced by the prolonged heat treatment [129]. Reprinted with permission from ref. [129]. © 2017 Elsevier.
Materials 17 03826 g014
Figure 15. Presents (a) the stress−strain curve and; (b) the Hall−Petch relation for Al0.3CoCrFeNi HEA at different annealing temperatures [129]. Reprinted with permission from ref. [129]. © 2017 Elsevier.
Figure 15. Presents (a) the stress−strain curve and; (b) the Hall−Petch relation for Al0.3CoCrFeNi HEA at different annealing temperatures [129]. Reprinted with permission from ref. [129]. © 2017 Elsevier.
Materials 17 03826 g015
Figure 16. SEM cross-section images of (a) Al25Co25Cr25Ni25, (b) Al20Co25Cr25Ni25Si5, and (c,d) Al9.1Co18.2Cr18.2Fe18.2Ni36.3 prior to oxidation under thermal shock state at 1273 K in air [132]. Reprinted with permission from ref. [132]. © 2022 Elsevier.
Figure 16. SEM cross-section images of (a) Al25Co25Cr25Ni25, (b) Al20Co25Cr25Ni25Si5, and (c,d) Al9.1Co18.2Cr18.2Fe18.2Ni36.3 prior to oxidation under thermal shock state at 1273 K in air [132]. Reprinted with permission from ref. [132]. © 2022 Elsevier.
Materials 17 03826 g016
Table 1. Variation in configurational entropy of mixing with number of components mixed in equimolar ratios [47].
Table 1. Variation in configurational entropy of mixing with number of components mixed in equimolar ratios [47].
nΔSconf
10
20.69R
31.10R
41.39R
51.61R
61.79R
71.95R
82.08R
92.20R
102.30R
112.40R
122.49R
132.57R
Table 2. ∆Sconf of some commonly used alloys [47].
Table 2. ∆Sconf of some commonly used alloys [47].
SystemAlloys∆Sconf at Liquid State
Low-Alloy Steel43400.22R low
Stainless Steel3040.96R low
3161.15R medium
High speed SteelM20.73R low
Mg AlloyAZ91D0.35R low
Al Alloy20240.29R low
70750.43R low
Cu-alloybrass0.61R low
Ni-Base SuperalloyInconel 7181.31 medium
Hastealloy1.37 medium
Co-Base SuperalloyStellite 61.13 medium
BMGCu47Zr11Ti34Ni81.17 medium
Zr53Ti5Cu16Ni10Al161.3 medium
Table 3. Major constituent elements found for use in HEAs [2].
Table 3. Major constituent elements found for use in HEAs [2].
Major Metallic ElementsMinor Metallic ElementsMinor Non-Metallic Elements
Li, Be, MgLi, Be, MgC, B, Si, P
Al, Sc, Ti, V,Sc, Ti, V, Cr, FeS, O, N
Cr, Fe, Co, Ni,Co, Ni, Cu, Zn,
Cu, Zn, Y, ZrGa, Ge, Sy, Cd
Nb, Mo, SmLn, Sn, Sb, Y
Eu, Au, Gd,Zr, Nb, Mo, Ru
Tb, Rh, Pb,Rh, Pb, Bi, Pd,
Pd, Ag, Hf, TaAg, Hf, Ta, W, Pt,
W, Pt, NdAu, La, Ce,
Pr, Nd, Sm, Fu.
Gd, Tb
Table 6. Depicts composition, mixing entropy (ΔSmix), and alloy densities for alloys: HESA-1 and HESA-2, respectively [121,124].
Table 6. Depicts composition, mixing entropy (ΔSmix), and alloy densities for alloys: HESA-1 and HESA-2, respectively [121,124].
SamplesNiAlCoCrFeTiTaMoW∆Smix [-R]
HESA-140.77.820.612.211.57.2---1.58
HESA-248.610.3177.595.80.60.80.41.56
Table 7. The Ƴ/Ƴ’ compositions of HESA-1, HESA-2, and some commercial Ni-based superalloys [121].
Table 7. The Ƴ/Ƴ’ compositions of HESA-1, HESA-2, and some commercial Ni-based superalloys [121].
SamplesNiAlCoCrFeTiTaMoWReRu
ƳHESA-132.34.824.718.216.13.9-----
HESA-240.16.222.51214.32.40.51.40.6--
CM247LC64.98.71110.5-0.70.90.23.1--
ME1553.73.614.820.80.1-0.11.33.521-
Rene’ N557.85.312.817.9--0.11.81.825-
RR210049.13.126.89.5--0.3-3.77.5-
RR2101463.126.79.5--0.3-3.67.92.9
ƳHESA-154.49.913.74.35.712-----
HESA-256.210.912.33.85.49.21.20.60.4--
CM247LC69.415.55.83.4-1.42.30.12.1--
ME156917.75.32.50-1.90.52.90.2-
Rene’ N574.515.74.32.6--10.51.20.2-
RR210067.116.98.61.4--2.6-2.90.5-
RR210166.116.68.81.4--2.7-30.50.9
Table 8. Designation and nominal chemical composition of samples investigated by Gawel et al. [132].
Table 8. Designation and nominal chemical composition of samples investigated by Gawel et al. [132].
DesignationNominal Chemical Composition [at.%]
AlCoCrFeNiSi
Al25Co25Cr25Ni252525250250
Al20Co25Cr25Ni25Si52025250255
Al9.1Co18.2Cr18.2Fe18.2Ni36.39.118.218.218.236.30
Table 9. Chemical compositions discovered using EDS point analysis on Al25Co25Cr25Ni25, Al20Co25Cr25Ni25Si5, and Al9.1Co18.2Cr18.2Fe18.2Ni36.3 cross-sections following oxidation under thermal shock conditions at 1273 K in ambient conditions [132].
Table 9. Chemical compositions discovered using EDS point analysis on Al25Co25Cr25Ni25, Al20Co25Cr25Ni25Si5, and Al9.1Co18.2Cr18.2Fe18.2Ni36.3 cross-sections following oxidation under thermal shock conditions at 1273 K in ambient conditions [132].
SamplePointElement [at.%]
AlCoCrFeNiSiO
Al25Co25Cr25Ni25 (Figure 16a)10.61.255.40.01.40.041.4
235.23.57.30.05.30.043.7
320.313.719.30.011.00.035.2
436.71.01.40.00.00.010.9
55.038.230.40.020.30.05.6
626.222.28.10.041.30.02.2
Al20Co25Cr25Ni25Si5 (Figure 16b)12.20.951.30.01.50.044.1
284.70.41.90.00.00.013.0
31.826.446.30.09.27.83.5
498.50.00.70.00.20.60.0
54.235.523.20.021.35.15.7
624.622.13.90.037.93.33.2
Al9.1Co18.2Cr18.2Fe18.2Ni36.3 (Figure 16d)10.40.654.51.31.00.042.2
21.718.615.919.434.60.09.8
350.30.74.50.30.40.043.3
415.31.441.31.91.70.038.3
599.60.00.40.00.00.00.0
63.018.111.220.136.20.011.4
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gupta, A.K.; Choudhari, A.; Rane, A.; Tiwari, A.; Sharma, P.; Gupta, A.; Sapale, P.; Tirumala, R.T.A.; Muthaiah, R.; Kumar, A. Advances in Nickel-Containing High-Entropy Alloys: From Fundamentals to Additive Manufacturing. Materials 2024, 17, 3826. https://doi.org/10.3390/ma17153826

AMA Style

Gupta AK, Choudhari A, Rane A, Tiwari A, Sharma P, Gupta A, Sapale P, Tirumala RTA, Muthaiah R, Kumar A. Advances in Nickel-Containing High-Entropy Alloys: From Fundamentals to Additive Manufacturing. Materials. 2024; 17(15):3826. https://doi.org/10.3390/ma17153826

Chicago/Turabian Style

Gupta, Ashish Kumar, Amit Choudhari, Aditya Rane, Abhishek Tiwari, Prince Sharma, Ashutosh Gupta, Prathamesh Sapale, Ravi Teja A. Tirumala, Rajmohan Muthaiah, and Abhishek Kumar. 2024. "Advances in Nickel-Containing High-Entropy Alloys: From Fundamentals to Additive Manufacturing" Materials 17, no. 15: 3826. https://doi.org/10.3390/ma17153826

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop