Open access peer-reviewed chapter

Microstructures and Deformation Mechanisms of FCC-Phase High-Entropy Alloys

Written By

Kaisheng Ming, Shijian Zheng and Jian Wang

Submitted: 04 April 2022 Reviewed: 05 April 2022 Published: 06 June 2022

DOI: 10.5772/intechopen.104822

From the Edited Volume

High Entropy Materials - Microstructures and Properties

Edited by Yong Zhang

Chapter metrics overview

336 Chapter Downloads

View Full Metrics

Abstract

Strength and ductility are the most fundamental mechanical properties of structural materials. Most metallurgical mechanisms for enhancing strength often sacrifice ductility, referred to as the strength–ductility trade-off. Over the past few decades, a new family of alloys—high-entropy alloys (HEAs) with multi-principal elements, has appeared great potential to overcome the strength–ductility trade-off. Among various HEAs systems, CrFeCoNi-based HEAs with a face-centered cubic (fcc) structure exhibit a great combination of strength, ductility, and toughness via tailoring microstructures. This chapter summarizes recent works on realizing strength–ductility combinations of fcc CrFeCoNi-based HEAs by incorporating multiple strengthening mechanisms, including solid solution strengthening, dislocation strengthening, grain boundary strengthening, and precipitation strengthening, through compositional and microstructural engineering. The abundant plastic deformation mechanisms of fcc HEAs, including slips associated with Shockley partial dislocation and full dislocations, nanotwinning, martensitic phase transformation, deformation-induced amorphization, and dynamically reversible shear transformation, are reviewed. The design strategies of advanced HEAs are also discussed in this chapter, which provides a helpful guideline to explore the enormous number of HEA compositions and their microstructures to realize exceptional strength–ductility combinations.

Keywords

  • high-entropy alloys
  • strength
  • ductility
  • twinning
  • phase transformation
  • amorphization
  • reversible shear transformation

1. Introduction

Developing structural materials with both high-strength and ductility could mitigate the ecological and economical concerns for decreasing weight and improving energy efficiency. Unfortunately, these properties are generally mutually exclusive, i.e., increasing strength will inevitably lead to ductility loss, an effect referred to as the strength–ductility trade-off [1, 2, 3, 4]. Various strengthening mechanisms, including solid solution strengthening, dislocation strengthening, grain boundary strengthening, precipitation strengthening, twinning, and phase transformation-induced hardening, have been widely utilized to produce high-strength and high-ductility alloys that are based on one principal element [5]. Recently, a new idea—high-entropy alloys (HEAs), has appeared that shows great potential to overcome the strength–ductility trade-off, and thus break through the mechanical property limits [6, 7, 8, 9, 10, 11, 12, 13]. HEAs were founded independently by two different research groups in 2004 [6, 7], which emphasize the unexplored regions in the center of multi-element phase diagrams, in which all elements are concentrated and there is no base element. HEAs are generally defined as an alloy that is composed of four or more elements in an equiatomic or near-equiatomic composition [7]. Each HEA can be considered as a new alloy base because its properties can be further optimized by minor elemental additions, similar to alloying in convention alloys. HEAs provide near-infinite new alloy bases for designing structural materials with excellent performance. Depending on the composition and microstructure, HEAs display attractive mechanical properties, and possible combinations of some properties, including high strength/hardness, outstanding wear resistance, excellent fatigue resistance, exceptional high-temperature strength, good structural stability, good corrosion and oxidation resistance, and high radiation tolerance [8, 10, 12, 14]. Particularly, single-phase face-centered cubic (fcc) HEAs and medium-entropy alloys (MEAs) based on the transition metal elements Cr, Mn, Fe, Co, or Ni generally display some of the best mechanical properties reported to date [9, 10, 11, 12, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25]. For example, equiatomic CrMnFeCoNi HEA and CoCrNi MEA possess exceptional combinations of tensile strength and ductility (tensile strength of ~1 GPa as well as ductility exceeding 60%) at 77 K, and ultra-high fracture toughness at both room temperature and 77 K (KJ1C > 200 MPa√m), making them one class of the toughest metallic materials reported so far [17, 23, 26]. Such exceptional mechanical properties are attributed to continuous steady strain-hardening, resulting from extensive dislocation activities and deformation-induced nanotwinning [18, 21, 22, 27]. These fcc-structured HEAs can be used as ideal alloy bases to design high-strength and high-ductility structural materials through further compositional and microstructural engineering.

Intensive studies have been invested in overcoming the strength–ductility trade-off of the single-phase fcc CrFeCoNi-based HEAs through tailoring the chemical composition and microstructure [9, 10, 11, 12, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24]. It has been reported that metastable high-entropy dual-phase alloys can overcome the strength–ductility trade-off by interface hardening and transformation-induced hardening, realized by reducing the stacking fault energy (SFE) via tailoring chemical composition [15, 19, 24, 28, 29, 30, 31, 32, 33]. The tensile strength and ductility are simultaneously enhanced due to heterogeneous microstructures, such as gradient nanotwins, gradient nano-grains, or recrystallized and non-recrystallized grains arranged in hierarchical structures with characteristic dimensions spanning from submicron scale to micro-scale, that are obtained by cold-rolling and annealing [9, 16, 27, 34, 35]. Note that single-phase fcc CrFeCoNi-based HEAs often have a very low SFE, which promotes deformation-induced nanotwinning and martensitic phase transformation [36, 37, 38, 39, 40, 41, 42]. The simultaneous increase in tensile strength and ductility is related to the enhanced strain-hardening capability enabled by nanotwinning and/or phase transformation. However, they generally possess very low yield strength since the nanotwinning and phase transformation cannot be activated at the early stages of plastic deformation [12].

Multiple strengthening mechanisms (such as solid solution strengthening, dislocation strengthening, grain boundary strengthening, and precipitation strengthening, et al.) show great potential to enhance the yield strength and tensile strength of fcc CrFeCoNi-based HEAs while maintaining their excellent ductility and strain-hardening capability [43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61, 62]. In addition, some new strengthening mechanisms are also proposed to improve the mechanical properties, such as magnetic hardening [60]. This chapter summarizes recent works on realizing strength–ductility combinations of fcc CrFeCoNi-based HEAs by incorporating multiple strengthening mechanisms, achieved through compositional and microstructural engineering. The abundant plastic deformation modes of fcc HEAs, including slips associated with Shockley partial dislocation and full dislocations, twinning, martensitic phase transformation, deformation-induced amorphization, and dynamically reversible shear transformation, are also summarized. We also demonstrate some HEA design strategies to provide guidelines for exploring the enormous number of HEA compositions and their microstructures to realize exceptional strength–ductility combinations.

Advertisement

2. Strengthening via tailoring composition and microstructure

2.1 Ductile nanoprecipitates

Single-phase fcc HEAs generally exhibit excellent ductility and strain-hardening capability but low yield strength at room temperature. Extensive studies demonstrate that nanoprecipitation strengthening is one of the most effective approaches to strengthen HEAs without apparent ductility loss. For example, Ming et al. [61] introduced highly dispersed nano-sized, coherent precipitates in the grain interior of coarse-grained fcc CrFeCoNi-based HEAs (grain size ~ 1 mm) via alloying a small addition of Al and Ti elements, which realizes exceptional combinations of strength and ductility (Figure 1). Yang et al. [47] produce a CrFeCoNi-based HEA with a superb yield strength of above 1.0 GPa while maintaining 50% ductility in tension at room temperature via introducing high-density ductile multicomponent intermetallic nanoparticles with coherent phase boundaries, enabled by alloying a small addition of Ti and Al elements. The size and distribution of nanoprecipitates can be tailored by the content of alloying elements (Ti and Al), aging temperature, and aging time [61]. As shown in Figure 1ag, the strength and ductility can be optimized corresponding to the aging time-dependent size and spacing of nanoprecipitates in an Al0.2Co1.5CrFeNi1.5Ti0.3 HEA. When aged at 800°C for 1–5 h, numerous uniformly dispersed, nano-sized, spherical L12 precipitates are formed in coarse grains (grain size ~ 1 mm), which results in a significant increase in both yield strength and ultimate tensile strength without apparent sacrificing of ductility (Figure 1a, b, f). With increasing aging time, the average diameter of precipitates increases from 6 nm at 1 h of aging to 51 nm at 100 h of aging (Figure 1e). Figure 1f presents that aging time does not show a significant effect on the yield strength and ultimate tensile strength, but increasing aging time results in an apparent decrease in ductility. As shown in Figure 1g, the strain-hardening rate obviously increases with increasing aging time, in particular, for the aging time equal and longer than 5 h. The samples aged for 1–5 h exhibit superior combinations of strength and ductility (yield strength of ~760 MPa, ultimate tensile strength of ~1160 MPa, elongation of ~40%).

Figure 1.

TEM images of the Al0.2Co1.5CrFeNi1.5Ti0.3 HEA after aging at 800°C for: (a) 1 h (b) 5 h, (c) 50 h, and (d) 100 h, with the corresponding selected area electron diffraction patterns inset. (e) Influence of aging time on the diameter of the precipitate (D), separation distance (L) between the centers of neighboring precipitates, and edge-to-edge inter-precipitate distance (l). (f) Engineering stress–strain curves of the solution-annealed and aged samples. (g) Variation of strain-hardening rates with plastic strain [61].

The nature of dislocation interaction with the nanoprecipitates in Al0.2Co1.5CrFeNi1.5Ti0.3 HEA samples is revealed by using transmission electron microscopy (TEM) analysis, as shown in Figure 2. The TEM bright-field image of the sample aging at 800°C for 1 h (Figure 2a) shows high-density dislocations tangled with stacking faults after tensile deformation to fracture. As shown in Figure 2b, original spherical precipitates have changed to an irregular shape after tensile deformation, which indicates the dislocations cutting through precipitates. In contrast to dislocation cutting through nano-sized precipitates, dislocations bypass relatively large precipitates by looping them in the sample aging at 800°C for 50 h, as demonstrated in Figure 2c. Figure 2dschematically illustrates that dislocation cutting through precipitates is the primary strengthening mechanism for the 1 h and 5 h aged HEA with small precipitates (diameter < 10 nm), while the dominant strengthening mechanism is dislocation bypassing precipitates by looping them in HEAs aged for >25 h with large precipitates (diameter > 30 nm). Based on the tensile testing and TEM characterizations, it can be concluded that nanoscale coherent precipitation strengthening is a very effective approach for strengthening coarse-grained fcc-structured HEAs without loss of ductility. The exceptional combinations of strength and ductility can be achieved via introducing uniformly distributed, nano-sized, coherent precipitates in the coarse-grained HEAs, where gliding dislocations cut through nano-sized coherent precipitates. In contrast, with increasing the size of precipitates, the ductility is reduced due to the strong barrier of large precipitates to dislocation motion where gliding dislocations bypass a precipitate by looping it, resulting in dislocations pileups at precipitate-matrix interfaces which generate high stress/strain concentration and micro-crack initiation.

Figure 2.

(a) TEM image of the 1 h aged Al0.2Co1.5CrFeNi1.5Ti0.3 HEA at a tensile strain of 36%, (b) HRTEM image of a precipitate with an irregular shape. (c) TEM image of the 50 h aged sample at a tensile strain of 3%. (d) Variation of precipitate shearing stress (σsh) and Orowan dislocation looping stress (σOr) with aging time [61].

It should be noted that the sluggish diffusion effect in HEAs enables easily tailoring the size of precipitates by adjusting aging processing and optimizing compositions. Recently, the coarsening kinetics and thermal stability of nanoscale precipitates in HEAs at elevated temperatures have also been widely studied to explore their applications at high temperatures [63, 64, 65, 66]. In a word, precipitation strengthening is one of the most promising approaches for enhancing the mechanical properties of fcc HEAs to meet the requirements for engineering applications both at room temperatures and high temperatures.

2.2 Brittle intermetallic compounds

The original concept employed in the HEAs design is to suppress the formation of brittle intermetallic compounds which can lead to poor ductility [6, 7]. Unfortunately, most HEAs reported to date contain various brittle intermetallic compounds [63, 64, 65, 66]. Therefore, it is required to manipulate the intermetallic compounds to reconcile the strength and ductility of HEAs. It has been shown that the brittle but hard intermetallic compound μ phase can be effectively used as a strengthening unit in CrFeCoNiMo HEAs while relieving its harmful effect on ductility by manipulating its dimension and distribution via tailoring Mo contents [62, 67, 68, 69]. Moreover, by further coupling solid solution hardening and nanotwinning induced hardening, a superb yield strength-tensile strength–ductility combination is realized [62]. As shown in Figure 3, the dimension and distribution of the μ phase can be tuned through thermal-mechanical processing and annealing. The μ phases grow mainly at boundaries (grain boundaries, triple junctions, and annealing twin boundaries), and secondarily in the interior of grains. The corresponding engineering stress–strain curves in Figure 4a,b demonstrate that dual-phase (fcc matrix + nanoscale μ phase) Cr15Fe20Co35Ni20Mo10 (Mo10) HEA displays high strength (yield strength of 0.8–1.3 GPa and ultimate tensile strength of 1.1–1.4 GPa) but moderate ductility (elongation to fracture is approximately 13–28%). The single-phase fcc Cr12.5Fe20Co42.5Ni20Mo5 (Mo5) HEAs have superb ductility (elongation to fracture of 45–75%) but moderate strength (yield strength of 0.3–0.8 GPa and ultimate tensile strength of 0.7–1.0 GPa). Figure 4c demonstrates that dual-phase Mo10 HEAs annealed at 850–1000°C exhibit a higher strain-hardening rate than single-phase Mo5 HEAs, and Mo10 HEA annealed at 1150°C as the true strain is less than half the maximum elongation. The high strain-hardening rate in Mo10 HEAs annealed at 850–1000°C is ascribed to the formation of distributed μ phase precipitates while the low strain-hardening rate in Mo10 HEAs annealed at 1150°C is due to the decomposition and disappearance of clustered μ phase precipitates. Figure 4d compares the mechanical properties of the dual-phase Mo10 HEA and single-phase Mo5 HEA with various other fcc HEAs [70, 71, 72, 73, 74, 75]. An exceptional combination of strength and ductility of Mo10 and Mo5 HEAs is attributed to the synergetic effect of solid solution strengthening, precipitation hardening, and twinning-induced hardening. The solution strengthening effect of the Mo addition in single-phase Mo10 alloys is evidenced by the higher strain-hardening rate of single-phase Mo10 alloy after true strain exceeds 13% (Figure 4c). Figure 5a,b demonstrates dislocations are bowing out or piled up at the μ phase interfaces. It is noted that μ precipitates are plastically non-shearable by gliding dislocations, which accounts for the high strength of Mo10 alloy according to the precipitation strengthening mechanism and the high strain-hardening rate at the early stage of plastic deformation. As demonstrated in Figure 5c, numerous deformation nano-twins are observed in single-phase Mo10 alloy at a strain of 25%, which provides a steady source of strain-hardening by blocking the motion of dislocations.

Figure 3.

TEM images of cold-rolled Mo10 HEAs after annealing at 800°C for (a1) 5 min, (a2) 1 h; 900°C for (b1) 5 min, (b2) 1 h; 1000°C for (c1) 5 min, (c2) 1 h [62].

Figure 4.

Engineering stress–strain curves for the (a) homogenized and recrystallized Mo10 alloys and (b) Mo5 alloys. (c) the strain-hardening rates of Mo10 and Mo5 alloys as a function of plastic strain. (d) Comparing the yield strength and elongation of the Mo10 and Mo5 alloys with various HEAs [62].

Figure 5.

TEM bright-field images of annealed Mo10 alloy at tension strains of 15–25%, (a) bow-out of gliding dislocations, (b) pile-up of dislocations at interfaces, and (c) high-density deformation nano-twins [62].

Other brittle intermetallic compounds, such as the sigma phase, Laves phases, have been also frequently observed in HEAs, which generally deteriorate the ductility [76, 77, 78, 79, 80]. Similarly, if the size, distribution, and morphology of intermetallic compounds can be tailored via controlling the chemical composition and thermomechanical treatment, an excellent combination of strength and ductility can be achieved in multi-component HEAs.

2.3 Hierarchically heterogeneous microstructures

Hierarchically heterogeneous microstructures at scales from a few nanometers to hundreds of micrometers can be easily produced in fcc HEAs by conventional cold work and annealing treatment, which could generate superior mechanical properties to those with simple microstructures due to the hetero-deformation induced strengthening [81]. As shown in Figure 6, single-phase fcc Cr20Fe6Co34Ni34Mo6 HEAs with hierarchical microstructures that comprise high-density annealing nanotwins in recrystallized fine grains (grain size ~ 1 μm), and dislocation walls and stacking faults in non-fully recrystallized fine grains are produced by cold-rolling and annealing [82]. The formation of numerous annealing nanotwins and stacking faults is obviously attributed to the very low SFE of the Cr20Fe6Co34Ni34Mo6 alloy. The addition of Mo in the Cr-Fe-Co-Ni system is found to be very effective in retarding the recrystallization and grain growth, promoting the formation of recrystallized fine grains. Such hierarchical microstructures can be generated in various single-phase fcc HEAs with low SFE by cold-rolling and annealing.

Figure 6.

TEM images of the cold-rolled Cr20Fe6Co34Ni34Mo6 HEAs after annealing at (a) 675°C for 1 h; (b) 700°C for 0.5 h; (c) 700°C for 1 h; (d) 800°C for 1 h. (e) Higher magnification micrograph of the interface between fully recrystallized grain and non-fully recrystallized grain, showing high density of dislocations, (f) selected area electron diffraction pattern patterns corresponding to non-fully recrystallized grain in (e) [82].

Comparing the mechanical properties of the samples with hierarchical microstructures and fully recrystallized coarse-grained microstructures, the former exhibit exceptional combinations of yield strength-ultimate tensile strength–ductility. As shown in Figure 7a, the single-phase HEAs with hierarchical microstructure (annealed at 675–800°C for 3 min, 0.5 h, 1 h, 5 h, or 10 h) exhibit a very high strength (YS = 0.95–1.1 GPa and UTS = 1.2–1.3 GPa, Figure 7a1) as well as a high ductility (EL is around 30–40%), which is much superior to that of the HEAs (annealed at 800–1150°C for 1 h) with simple fully recrystallized microstructures (Figure 7a2). Such high yield strengths exceeding 1.0 GPa at enhanced ductility (30%) in single-phase HEAs are comparable to the nanoprecipitation strengthened HEAs (Figure 7b1,b2) [17, 23, 70, 71, 83, 84, 85, 86, 87, 88, 89]. This indicates that enhancing yield strength while retaining good ductility of single-phase fcc HEAs can be achieved by developing hierarchical microstructures. The grain boundaries, annealing nanotwins, and dislocation walls play important roles in enhancing the strength, ductility, and strain-hardening capability of the annealed sample with hierarchical microstructure. Figure 7c plots the yield strength (σy) as a function of the grain size (d), which follows the well-known Hall–Petch relationship, σy = σ0 + k·d−1/2. The σ0 is 246 MPa. The Hall–Petch coefficient k = 743 MPa·μm1/2 is higher than that of most fcc metals (600 MPa·μm1/2) [90], which indicates that the grain-boundary strengthening mechanism is very effective for improving yield strength. Microstructure characterizations reveal that both twin boundaries and dislocation walls act as strong barriers for dislocation motion, strengthening the alloy, as shown in Figure 8. Figure 8a shows a large number of stacking faults (marked by red arrows) around twin boundaries. The enlarged HRTEM image in Figure 8b shows high-density stacking faults in the twin and matrix, indicating slip transmission associated with dislocations crossing through the coherent twin boundaries (CTBs). Consequently, atomically flat CTBs develop stepped or serrated CTBs that contain interface defects. These interfacial defects along stepped or serrated CTBs then provide sources for nucleating and blocking dislocations. Figure 8c,d shows that dislocation walls act as sources for dislocations and deformation twins at large deformation stages, contributing to strain-hardening by continually introducing new interfaces and blocking the motion of dislocations. The enhanced strength enables the production of dislocations at serrated/stepped CTBs, preventing the early onset of necking instability. Therefore, the strength is enhanced without apparent ductility loss.

Figure 7.

(a1, a2) engineering stress–strain curves of the cold-rolled and annealed Cr20Fe6Co34Ni34Mo6 HEAs at various temperatures of 675–1150°C. (b1, b2) yield strength and ultimate tensile strength versus uniform elongation of annealed alloys compared with various HEAs. (c) the variation of yield strength with average grain size [82].

Figure 8.

(a, b) TEM micrographs of the recrystallized grains in annealed Cr20Fe6Co34Ni34Mo6 HEAs (700°C for 1 h) at a strain of ∼30%. (c, d) TEM micrographs of non-fully recrystallized grains at strains of 15–30% [82].

Other hierarchically heterogeneous microstructures, such as gradient grain structure, gradients in twin and dislocation densities, hierarchical nano-twins, could be introduced into fcc HEAs via surface mechanical grinding treatment, which contributes to the strain hardening capability and thus exceptional strength–ductility combinations, owing to the hetero-deformation induced strengthening. These results demonstrate that engineering the hierarchical microstructure should be an efficient strategy for enhancing the strength and ductility of single-phase fcc HEAs with low and medium SFE.

2.4 Shear transformation bands

Depending on chemical composition and deformation temperature, fcc HEAs with relatively low SFE generate various types of nanoscale shear transformation bands during plastic deformation, including stacking fault bands, nanotwin bands, phase transformation bands, deformation bands with ultrahigh density dislocation, and amorphous bands [10]. The shear transformation banding plays an important role in enhancing strain-hardening capability and thus contributes to enhanced strength and ductility simultaneously. The typical examples are twinning-induced plasticity (TWIP) and transformation-induced plasticity (TRIP) effects in metastable HEAs, which could overcome the long-standing strength–ductility dilemma [18, 21, 22, 27]. For example, the strength, ductility, and toughness of single-phase fcc CrMnFeCoNi HEAs can be improved simultaneously with decreasing temperature from room temperature to 77 K, owing to the transition of deformation mechanisms from dislocation slip to nanotwinning [17, 20, 21, 23]. As shown in Figure 9, the formation of stacking faults, nanotwinning, and martensitic phase transformation could be generated sequentially in a non-equiatomic CrMnFeCoNi HEA with increasing tensile strains at low temperatures, leading to the formation of various nanoscale bands [91]. Such shear transformation banding, by continually introducing new interfaces and decreasing the mean free path of dislocations during tensile testing (“dynamic Hall–Petch”), produces a high degree of work-hardening and a significant increase in the ultimate tensile strength [92, 93, 94].

Figure 9.

TEM images of the single-phase fcc CrMnFeCoNi HEA after tension to different strains at 93 K, show the sequential formation of stacking faults, nanotwinning, and martensitic phase transformation [91].

Recently, amorphous bands (one type of shear transformation bands) have been observed in a non-equiatomic CrMnFeCoNi HEA under tensile deformation at 93 K [91]. These amorphous bands are generated by deformation-induced solid-state amorphization, as summarized in Figure 10. The nanoscale amorphous bands, stacking fault bands, nanotwin bands, phase transformation bands can be generated simultaneously in HEAs, which contributes to enhanced strength, ductility, and strain-hardening rate synergistically. In particular, the nanoscale amorphous bands exhibit much higher thermal stability than nanotwins and phase transformation bands. Amorphous bands ensure the enhanced strength and good ductility of high-temperature tempered samples. Amorphous bands can plastically co-deform with the matrix. The interfaces between the amorphous band and fcc matrix provide not only strong barriers for dislocation motion, strengthening materials, but also natural sinks of dislocations, disrupting stress concentrations and delaying decohesion and fracture initiation. These results demonstrate that engineering amorphous bands could be an efficient strategy in remaining enhanced mechanical properties of fcc HEAs at high temperatures.

Figure 10.

(a, b) TEM images of the non-equiatomic CrMnFeCoNi HEA after tension to a strain of 36% at 93 K, (c, d) the corresponding high-resolution TEM images showing the amorphous bands. (e–g) TEM images show the interactions of stacking faults, nano-twins, and amorphous bands [91].

In addition, Ming et al. design nano-laminated dual-phase structures in a non-equiatomic CrMnFeCoNi HEA via dynamically reversible shear transformations associated with reversible martensitic phase transformation and nanotwinning. The detailed mechanism for dynamically reversible shear transformations can be seen in the following sections. The nano-laminated dual-phase structures could evade the strength–ductility dilemma due to the synergistic operations of the TRIP effect, TWIP effect, and associated with “dynamic Hall–Petch” effect. In a word, the abundant deformation mechanisms of fcc HEAs, as described below, enable us to design various types of microstructures to achieve an exceptional combination of strength and ductility.

Advertisement

3. Deformation mechanisms

Similar to conventional fcc metals and alloys, fcc HEAs plastically deform through three deformation mechanisms: dislocation slip, twinning, and phase transformation. Recently, some unique deformation pathways of deformation-induced amorphization and dynamically reversible shear transformations which rarely occur in conventional materials are found in fcc HEAs.

3.1 Deformation-induced amorphization

Various mechanical processes can partially or fully amorphized crystalline materials, including high-pressure treatments, severe plastic deformation, or mechanical alloying [95, 96, 97, 98, 99, 100, 101, 102, 103]. From the thermodynamic driving force perspective, amorphization starts from the massive displacement of atoms into metastable positions and occurs when the free energy of the crystalline phase is higher than that of the amorphous phase [104, 105, 106]. From the kinetic hindrance perspective, the formation of a metastable amorphous phase requires the kinetic hindrances which block the formation of a more stable equilibrium crystalline phase [104]. It should be noted that deformation-induced soild-state amorphization under stresses can take place in the elastic regime at an extremely high strain rate, or in the plastic stage at severe strain [95]. At an extremely high strain rate, there is no enough time to activate plastic deformation modes such as dislocation slips or even the faster twinning mode. Thus, the large elastic strains promote the crystalline phase mechanically unstable due to the loss of shear rigidity, which leads to amorphization [107, 108, 109].

Interestingly, Ming et al. [91] for the first time observed deformation-induced amorphization in a non-equiatomic Cr26Mn20Fe20Co20Ni14 HEA at cryogenic temperature. Such deformation-induced amorphization was later observed by several groups [110, 111, 112, 113]. The deformation-induced amorphization generates extensive nanoscale amorphous bands, as shown in Figure 10ad. The formation of extensive nanoscale amorphous bands is attributed to the significant dislocation accumulation in a constrained region inside shear bands, which raises the free energy of the original fcc phase to a point higher than that of the amorphous phase, then the energy difference drives amorphization. Subsequently, the deformation-induced amorphization was also observed in equiatomic CrMnFeCoNi HEA under severe plastic deformation through swaging followed by either quasi-static compression or dynamic deformation in shear [110]. In addition, the deformation-induced localized amorphization can also occur at the tip of cracks, which enhances the toughness significantly by blunting cracks and impeding the expansion of cracks [111]. Subsequently, the nanoscale origin of the mechanism of amorphization is revealed by using molecular dynamics simulation [112, 113]. The amorphization originates from the formation of multi-dislocation junctions due to the low SFE, which results in high lattice resistance to dislocation glide and facilitates nucleation of amorphous nuclei. The deformation mechanisms in the amorphous/crystalline dual-phase regions include high-density Shockley partial dislocations, multi-dislocation junctions, and nanotwinning in the crystalline region (TEM observations in Figure 10eg), as well as radiation-shaped shear bands and amorphous bridges in the amorphous region.

Based on the mechanism of deformation-induced solid-state amorphization in HEAs, nanoscale amorphization bands can be introduced into HEAs to optimize mechanical properties. It is demonstrated that the yield strength of the HEAs could be enhanced without apparent loss of ductility by introducing high-density nanoscale amorphous bands. TEM characterizations (Figure 10eg) reveal that introducing nanoscale amorphous bands can achieve three key benefits: (i) amorphous-crystalline interface (ACI) hardening, i.e., ACIs not only act as high-capacity sources for dislocations nucleation but also barriers for dislocation motion; (ii) large stress concentrations at the ACIs can be disrupted and relieved by the amorphous bands since ACIs act as natural sinks of dislocations, averting dislocation pileups at ACIs, which delay decohesion and fracture initiation at the ACIs; (iii) high thermal stability of amorphous bands enables to increase strain-hardening capability through the tempering at relatively higher temperatures without sacrificing the high yield strength. Wang et al. [111] observed the formation of the amorphous area ahead of the crack tip after amorphization in equiatomic CrMnFeCoNi HEA by using in situ straining TEM experiments and found that the amorphous bridges in the crack wake provided effective toughening of the HEA. Ji et al. [112] provided atomistic insights, via molecular dynamics simulations, into the origin of the solid-state amorphization ahead of a crack tip and report the deformation mechanisms contributing to cryogenic damage-tolerance. It is believed that a much better combination of yield strength, ductility, and toughness can be achieved through optimizing the amorphization. These results demonstrate that engineering amorphous bands could be an efficient strategy in remaining enhanced mechanical properties of fcc-structured HEAs.

3.2 Dynamically reversible shear transformations

Traditional shear transformation banding, such as deformation-induced {111} twinning and martensitic phase transformation (fcc-γ → hcp-ε), has been widely observed in fcc HEAs with low SFE. Recently, Ming et al. [114] found two dynamically reversible shear transformation mechanisms in a CrMnFeCoNi HEA under uniaxial tension at 4.2 K, featured by γ → ε → {101¯1} twin → γ/γtw and γ → ε → γ/γtw. When deformed at cryogenic temperature, the lower SFE promotes γ → ε shear transformation, forming hcp grains which gradually deplete the original fcc grains. Meanwhile, high-density {0001} stacking faults and {101¯1} nanotwinning are activated to accommodate plastic deformation, as shown in Figure 11a,b. More intriguingly, reverse hcp → fcc shear transformations are stimulated within {101¯1} twin and surrounding hcp matrix by deformation-induced local dissipative heating (Figure 11c). When the {101¯1} twins transform into fcc structure, Shockley partial dislocations are activated on {111} planes, leading to formation of {111} SFs and {111} nanotwins in the newly formed fcc domain. Figure 11c shows two fcc domains with {111} twin orientation inside an {101¯1} twin. This shear transformation mechanism is described by γ → ε → {101¯1} twin → γ/γtw.

Figure 11.

(a–c) TEM images a CrMnFeCoNi HEA after uniaxial tension to fracture at 4.2 K, showing the reversible shear transformation mechanism featured by γ → ε → {11} twin → γ/γtw. (d, e) the reversible shear transformation mechanism of γ → ε → γ/γtw [114].

In addition, for high-density basal stacking faults in the hcp-ε phase, the faulted regions can transform back into fcc structure through correcting SFs via nucleation and glide of Shockley partial dislocations, which is energetically favorable since the process will reduce the density of stacking faults. It is noted that fcc laminates frequently have two orientations, forming a {111} twin orientation relationship (Figure 11d,e). This shear transformation mechanism is described by γ → ε → γ/γtw. The reversible fcchcp shear transformations and both {101¯1} and {111} nanotwinning lead to dynamic nano-laminated dual-phase structures, which advance the monotonic “dynamic Hall–Petch” effect in enhancing strength, strain-hardening ability, and ductility by dynamically tailoring the type and width of shear transformation bands.

3.3 Nano-segregation of multi-principal elements

The unique multiple principal elements endow HEAs with adjustable microstructures and corresponding excellent mechanical properties. Extensive efforts invest in tailoring the chemical compositions of fcc-structured HEAs to optimize the strength and ductility simultaneously. When designing the compositions of HEAs, one should realize two important facts. The first one is that the stronger HEAs are not necessarily the ones with the most elements. The nature of the constituent elements is also important, with the Cr-containing alloys, in general, being the strongest in a family of equiatomic binary, ternary and quaternary alloys based on the elements Fe, Ni, Co, Cr, and Mn [71]. Secondly, when designing the compositions of HEAs, nano-segregation of principal elements at grain boundaries (GBs) should be cause for concern since it could deteriorate the ductility significantly. As shown in Figure 12, nanoclustering Cr, Ni, and Mn separately at GBs, as detected by atom probe tomography, reduce GB cohesion, and promotes crack initiation along GBs, leading to ductility loss in the CrMnFeCoNi HEA [115]. The GB segregation engineering strategy is then proposed to avoid ductility loss by shifting the fast segregation of principal elements from GBs into preexisting Cr-rich secondary phases. Such GB decohesion by nanoclustering multi-principal elements is a common phenomenon in HEAs. Linlin Li et al. also found that Ni and Mn co-segregate to some regions of the GBs along with the depletion of Fe, Co, and Cr, while Cr is enriched in other regions of the GBs where Ni and Mn are depleted [116, 117, 118].

Figure 12.

Atom probe tomography of the GB in a CrMnFeCoNi HEA after uniaxial tension to fracture at 700°C: (a) atom map of the tip showing Cr (24 at %) and Ni (23 at %) isocomposition surfaces viewed with the GB edge-on. (b) 1D compositional profiles along the cyan arrow E and the orange arrow F are indicated in (a). (c) Fracture lateral surface of the tensile sample after tensile tests at 700 °C, showing many cracks along the GBs [115].

Generally, only nano-segregation of several elements occurs in fully annealed HEAs with coarse and clean grains after short-time annealing treatment, which can lead to GB decohesion and thus ductility loss. However, with very long-time annealing, such as annealing at intermediate temperatures for tens to hundreds of days, intermetallic phases are precipitated at GBs in coarse-grained HEAs [119, 120]. When the grain size of HEAs is decreased to the nanoscale, annealing at intermediate temperatures for mere minutes can lead to precipitation of nanoscale intermetallic phases at GBs, such as Cr-rich phase, NiMn phase, and FeCo phase in equiatomic CrMnFeCoNi HEA [121]. The formation of intermetallic phases at GBs will reduce the ductility significantly. Therefore, the nature of the constituent elements and their nano-segregation behavior should be considered when we design high-strength and high-ductility HEAs.

Advertisement

4. Conclusion

This chapter summarizes recent works on realizing strength–ductility combinations in fcc CrFeCoNi-based HEAs via composition and microstructure engineering: (I) Nanoprecipitation strengthening associated with tailoring the size, distribution, and morphology of second phases via alloying a small addition of Ti and Al elements; and (II) The synergistic operations of multiple strengthening mechanisms, such as solid solution strengthening, dislocation strengthening, grain boundary strengthening, precipitation strengthening, TWIP/TRIP effect, and amorphization-induced strengthening. The abundant deformation mechanisms, including slips associated with Shockley partial dislocation and full dislocations, nanotwinning, martensitic phase transformation, deformation-induced amorphization, and dynamically reversible shear transformation, are also summarized. Among them, the recently reported deformation-induced amorphization and dynamically reversible shear transformation are highlighted in terms of their nanoscale origins and strengthening effects for overcoming the strength–ductility trade-off. Finally, this chapter points out that the nature of the constituent elements and their nano-segregation behavior should be considered when we design high-strength and high-ductility HEAs via engineering the compositions.

Advertisement

Acknowledgments

The authors acknowledge the financial support from the National Natural Science Foundation of China (No. 52002109), the Natural Science Foundation of the Hebei province (No. E2020202088), and the Overseas Scientists Sponsorship Program by Hebei Province (C20210331).

Advertisement

Conflict of interest

The authors declare no competing financial interests.

References

  1. 1. Ritchie RO. The conflicts between strength and toughness. Nature Materials. 2011;10:817-822. DOI: 10.1038/nmat3115
  2. 2. Launey ME, Ritchie RO. On the fracture toughness of advanced materials. Advanced Materials. 2009;21:2103-2110. DOI: 10.1002/adma.200803322
  3. 3. Han SZ, Choi E-A, Lim SH, Kim S, Lee J. Alloy design strategies to increase strength and its trade-offs together. Progress in Materials Science. 2020;117:100720. DOI: 10.1016/j.pmatsci.2020.100720
  4. 4. Wu H, Fan G. An overview of tailoring strain delocalization for strength-ductility synergy. Progress in Materials Science. 2020;113:100675. DOI: 10.1016/j.pmatsci.2020.100675
  5. 5. Argon A. Strengthening Mechanisms in Crystal Plasticity. 1st ed. Vol. 4. Oxford: OUP; 2007. DOI: 10.1093/acprof:oso/9780198516002.001.0001
  6. 6. Zhang Y, Zuo TT, Tang Z, Gao MC, Dahmen KA, Liaw PK, et al. Microstructures and properties of high-entropy alloys. Progress in Materials Science. 2014;61:1-93. DOI: 10.1016/j.pmatsci.2013.10.001
  7. 7. Yeh JW, Chen SK, Lin SJ, Gan JY, Chin TS, Shun TT, et al. Nanostructured high-entropy alloys with multiple principal elements: Novel alloy design concepts and outcomes. Advanced Engineering Materials. 2004;6:299-303. DOI: 10.1002/adem.200300567
  8. 8. Li W, Xie D, Li D, Zhang Y, Gao Y, Liaw PK. Mechanical behavior of high-entropy alloys. Progress in Materials Science. 2021;118:100777. DOI: 10.1016/j.pmatsci.2021.100777
  9. 9. Sathiyamoorthi P, Kim HS. High-entropy alloys with heterogeneous microstructure: Processing and mechanical properties. Progress in Materials Science. 2020;123:100709. DOI: 10.1016/j.pmatsci.2020.100709
  10. 10. George EP, Curtin WA, Tasan CC. High entropy alloys: A focused review of mechanical properties and deformation mechanisms. Acta Materialia. 2020;188:435-474. DOI: 10.1016/j.actamat.2019.12.015
  11. 11. Cantor B. Multicomponent high-entropy Cantor alloys. Progress in Materials Science. 2020;120:100754. DOI: 10.1016/j.pmatsci.2020.100754
  12. 12. Li Z, Zhao S, Ritchie RO, Meyers MA. Mechanical properties of high-entropy alloys with emphasis on face-centered cubic alloys. Progress in Materials Science. 2019;102:296-345. DOI: 10.1016/j.pmatsci.2018.12.003
  13. 13. Miracle DB, Senkov ON. A critical review of high entropy alloys and related concepts. Acta Materialia. 2017;122:448-511. DOI: 10.1016/j.actamat.2016.08.081
  14. 14. Zhang Z, Armstrong DEJ, Grant PS. The effects of irradiation on CrMnFeCoNi high-entropy alloy and its derivatives. Progress in Materials Science. 2021;123:100807. DOI: 10.1016/j.pmatsci.2021.100807
  15. 15. Li Z, Pradeep KG, Deng Y, Raabe D, Tasan CC. Metastable high-entropy dual-phase alloys overcome the strength-ductility trade-off. Nature. 2016;534:227-230. DOI: 10.1038/nature17981
  16. 16. Ma E, Wu X. Tailoring heterogeneities in high-entropy alloys to promote strength–ductility synergy. Nature Communications. 2019;10:1-10. DOI: 10.1038/s41467-019-13311-1
  17. 17. Gludovatz B, Hohenwarter A, Catoor D, Chang EH, George EP, Ritchie RO. A fracture-resistant high-entropy alloy for cryogenic applications. Science. 2014;345:1153-1158. DOI: 10.1126/science.1254581
  18. 18. Chen S, Oh HS, Gludovatz B, Kim SJ, Park ES, Zhang Z, et al. Real-time observations of TRIP-induced ultrahigh strain hardening in a dual-phase CrMnFeCoNi high-entropy alloy. Nature Communications. 2020;11:1-8. DOI: 10.1038/s41467-020-14641-1
  19. 19. Lu W, Liebscher CH, Dehm G, Raabe D, Li Z. Bidirectional transformation enables hierarchical Nanolaminate dual-phase high-entropy alloys. Advanced Materials. 2018;30:e1804727. DOI: 10.1002/adma.201804727
  20. 20. Naeem M, He H, Zhang F, Huang H, Harjo S, Kawasaki T, et al. Cooperative deformation in high-entropy alloys at ultralow temperatures. Science Advances. 2020;6:eaax4002. DOI: 10.1126/sciadv.aax4002
  21. 21. Zhang Z, Mao MM, Wang J, Gludovatz B, Zhang Z, Mao SX, et al. Nanoscale origins of the damage tolerance of the high-entropy alloy CrMnFeCoNi. Nature Communications. 2015;6:10143. DOI: 10.1038/ncomms10143
  22. 22. Zhang Z, Sheng H, Wang Z, Gludovatz B, Zhang Z, George EP, et al. Dislocation mechanisms and 3D twin architectures generate exceptional strength-ductility-toughness combination in CrCoNi medium-entropy alloy. Nature Communications. 2017;8:14390. DOI: 10.1038/ncomms14390
  23. 23. Otto F, Dlouhý A, Somsen C, Bei H, Eggeler G, George EP. The influences of temperature and microstructure on the tensile properties of a CoCrFeMnNi high-entropy alloy. Acta Materialia. 2013;61:5743-5755. DOI: 10.1016/j.actamat.2013.06.018
  24. 24. Su J, Wu X, Raabe D, Li Z. Deformation-driven bidirectional transformation promotes bulk nanostructure formation in a metastable interstitial high entropy alloy. Acta Materialia. 2019;167:23-39. DOI: 10.1016/j.actamat.2019.01.030
  25. 25. Moon J, Tabachnikova E, Shumilin S, Hryhorova T, Estrin Y, Brechtl J, et al. Deformation behavior of a Co-Cr-Fe-Ni-Mo medium-entropy alloy at extremely low temperatures. Materials Today. 2021;50:55-68. DOI: 10.1016/j.mattod.2021.08.001
  26. 26. Gludovatz B, Hohenwarter A, Thurston KV, Bei H, Wu Z, George EP, et al. Exceptional damage-tolerance of a medium-entropy alloy CrCoNi at cryogenic temperatures. Nature Communications. 2016;7:10602. DOI: 10.1038/ncomms10602
  27. 27. He H, Naeem M, Zhang F, Zhao Y, Harjo S, Kawasaki T, et al. Stacking fault driven phase transformation in CrCoNi medium entropy alloy. Nano Letters. 2021;21:1419-1426. DOI: 10.1021/acs.nanolett.0c04244
  28. 28. Li Z, Tasan CC, Pradeep KG, Raabe D. A TRIP-assisted dual-phase high-entropy alloy: Grain size and phase fraction effects on deformation behavior. Acta Materialia. 2017;131:323-335. DOI: 10.1016/j.actamat.2017.03.069
  29. 29. Li Z, Tasan CC, Springer H, Gault B, Raabe D. Interstitial atoms enable joint twinning and transformation induced plasticity in strong and ductile high-entropy alloys. Scientific Reports. 2017;7:40704. DOI: 10.1038/srep40704
  30. 30. Su J, Raabe D, Li Z. Hierarchical microstructure design to tune the mechanical behavior of an interstitial TRIP-TWIP high-entropy alloy. Acta Materialia. 2019;163:40-54. DOI: 10.1016/j.actamat.2018.10.017
  31. 31. Wang M, Li Z, Raabe D. In-situ SEM observation of phase transformation and twinning mechanisms in an interstitial high-entropy alloy. Acta Materialia. 2018;147:236-246. DOI: 10.1016/j.actamat.2018.01.036
  32. 32. Li Z, CemTasan C, Pradeep KG, Raabe D. A TRIP-assisted dual-phase high- entropy alloy:Grain size and phase fraction eff ects on de formation behavior. Acta Materialia. 2017;131:323-335. DOI: 10.1016/j.actamat.2017.03.069
  33. 33. Li Z, Körmann F, Grabowski B, Neugebauer J, Raabe D. Ab initio assisted design of quinary dual-phase high-entropy alloys with transformation-induced plasticity. Acta Materialia. 2017;136:262-270. DOI: 10.1016/j.actamat.2017.07.023
  34. 34. Jo YH, Jung S, Choi WM, Sohn SS, Kim HS, Lee BJ, et al. Cryogenic strength improvement by utilizing room-temperature deformation twinning in a partially recrystallized VCrMnFeCoNi high-entropy alloy. Nature Communications. 2017;8:15719. DOI: 10.1038/ncomms15719
  35. 35. Pan Q , Zhang L, Feng R, Lu Q , An K, Chuang AC, et al. Gradient-cell–structured high-entropy alloy with exceptional strength and ductility. Science. 2021;374:eabj8114. DOI: 10.1126/science.abj8114
  36. 36. Ding J, Yu Q , Asta M, Ritchie RO. Tunable stacking fault energies by tailoring local chemical order in CrCoNi medium-entropy alloys. Proceedings of the National Academy of Sciences. 2018;115:8919-8924. DOI: 10.1073/pnas.1808660115
  37. 37. Huang S, Huang H, Li W, Kim D, Lu S, Li X, et al. Twinning in metastable high-entropy alloys. Nature Communications. 2018;9:1-7. DOI: 10.1038/s41467-018-04780-x
  38. 38. Shih M, Miao J, Mills M, Ghazisaeidi M. Stacking fault energy in concentrated alloys. Nature Communications. 2021;12:3590. DOI: 10.1038/s41467-021-23860-z
  39. 39. Khan TZ, Kirk T, Vazquez G, Singh P, Smirnov AV, Johnson DD, et al. Towards stacking fault energy engineering in FCC high entropy alloys. Acta Materialia. 2022;224:117472. DOI: 10.1016/j.actamat.2021.117472
  40. 40. Wagner C, Ferrari A, Schreuer J, Couzinié J-P, Ikeda Y, Körmann F, et al. Effects of Cr/Ni ratio on physical properties of Cr-Mn-Fe-Co-Ni high-entropy alloys. Acta Materialia. 2022;227:117693. DOI: 10.1016/j.actamat.2022.117693
  41. 41. Zhao S, Stocks GM, Zhang Y. Stacking fault energies of face-centered cubic concentrated solid solution alloys. Acta Materialia. 2017;134:334-345. DOI: 10.1016/j.actamat.2017.05.001
  42. 42. Huang S, Li W, Lu S, Tian F, Shen J, Holmström E, et al. Temperature dependent stacking fault energy of FeCrCoNiMn high entropy alloy. Scripta Materialia. 2015;108:44-47. DOI: 10.1016/j.scriptamat.2015.05.041
  43. 43. Yang Y, Chen T, Tan L, Poplawsky JD, An K, Wang Y, et al. Bifunctional nanoprecipitates strengthen and ductilize a medium-entropy alloy. Nature. 2021;595:245-249. DOI: 10.1038/s41586-021-03607-y
  44. 44. Jang TJ, Choi WS, Kim DW, Choi G, Jun H, Ferrari A, et al. Shear band-driven precipitate dispersion for ultrastrong ductile medium-entropy alloys. Nature Communications. 2021;12:4703. DOI: 10.1038/s41467-021-25031-6
  45. 45. Han L, Rao Z, Souza Filho IR, Maccari F, Wei Y, Wu G, et al. Ultrastrong and ductile soft magnetic high-entropy alloys via coherent ordered Nanoprecipitates. Advanced Materials. 2021;33:e2102139. DOI: 10.1002/adma.202102139
  46. 46. Feng R, Rao Y, Liu C, Xie X, Yu D, Chen Y, et al. Enhancing fatigue life by ductile-transformable multicomponent B2 precipitates in a high-entropy alloy. Nature Communications. 2021;12:3588. DOI: 10.1038/s41467-021-23689-6
  47. 47. Yang T, Zhao Y, Tong Y, Jiao Z, Wei J, Cai J, et al. Multicomponent intermetallic nanoparticles and superb mechanical behaviors of complex alloys. Science. 2018;362:933-937. DOI: 10.1126/science.aas8815
  48. 48. Liang YJ, Wang L, Wen Y, Cheng B, Wu Q , Cao T, et al. High-content ductile coherent nanoprecipitates achieve ultrastrong high-entropy alloys. Nature Communications. 2018;9:4063. DOI: 10.1038/s41467-018-06600-8
  49. 49. Oh HS, Kim SJ, Odbadrakh K, Ryu WH, Yoon KN, Mu S, et al. Engineering atomic-level complexity in high-entropy and complex concentrated alloys. Nature Communications. 2019;10:2090. DOI: 10.1038/s41467-019-10012-7
  50. 50. Sohn SS, Kwiatkowski da Silva A, Ikeda Y, Kormann F, Lu W, Choi WS, et al. Ultrastrong medium-entropy single-phase alloys designed via severe lattice distortion. Advanced Materials. 2019;31:e1807142. DOI: 10.1002/adma.201807142
  51. 51. Lei Z, Liu X, Wu Y, Wang H, Jiang S, Wang S, et al. Enhanced strength and ductility in a high-entropy alloy via ordered oxygen complexes. Nature. 2018;563:546-550. DOI: 10.1038/s41586-018-0685-y
  52. 52. Ding Q , Zhang Y, Chen X, Fu X, Chen D, Chen S, et al. Tuning element distribution, structure and properties by composition in high-entropy alloys. Nature. 2019;574:223-227. DOI: 10.1038/s41586-019-1617-1
  53. 53. Li QJ, Sheng H, Ma E. Strengthening in multi-principal element alloys with local-chemical-order roughened dislocation pathways. Nature Communications. 2019;10:3563. DOI: 10.1038/s41467-019-11464-7
  54. 54. Zhang R, Zhao S, Ding J, Chong Y, Jia T, Ophus C, et al. Short-range order and its impact on the CrCoNi medium-entropy alloy. Nature. 2020;581:283-287. DOI: 10.1038/s41586-020-2275-z
  55. 55. Chen S, Aitken ZH, Pattamatta S, Wu Z, Yu ZG, Srolovitz DJ, et al. Simultaneously enhancing the ultimate strength and ductility of high-entropy alloys via short-range ordering. Nature Communications. 2021;12:4953. DOI: 10.1038/s41467-021-25264-5
  56. 56. Wu X, Wang B, Rehm C, He H, Naeem M, Lan S, et al. Ultra-small-angle neutron scattering study on temperature-dependent precipitate evolution in CoCrFeNiMo0.3 high entropy alloy. Acta Materialia. 2022;222:117446. DOI: 10.1016/j.actamat.2021.117446
  57. 57. Wang L, Wang L, Zhou S, Xiao Q , Xiao Y, Wang X, et al. Precipitation and micromechanical behavior of the coherent ordered nanoprecipitation strengthened Al-Cr-Fe-Ni-V high entropy alloy. Acta Materialia. 2021;216:117121. DOI: 10.1016/j.actamat.2021.117121
  58. 58. Huang X, Huang L, Peng H, Liu Y, Liu B, Li S. Enhancing strength-ductility synergy in a casting non-equiatomic NiCoCr-based high-entropy alloy by Al and Ti combination addition. Scripta Materialia. 2021;200:113898. DOI: 10.1016/j.scriptamat.2021.113898
  59. 59. He F, Yang Z, Liu S, Chen D, Lin W, Yang T, et al. Strain partitioning enables excellent tensile ductility in precipitated heterogenous high-entropy alloys with gigapascal yield strength. International Journal of Plasticity. 2021;144:103022. DOI: 10.1016/j.ijplas.2021.103022
  60. 60. Niu C, LaRosa CR, Miao J, Mills J, Mills M, Ghazisaeidi M. Magnetically-driven phase transformation strengthening in high entropy alloys. Nature Communications. 2018;9:1363. DOI: 10.1038/s41467-018-03846-0
  61. 61. Ming K, Bi X, Wang J. Realizing strength-ductility combination of coarse-grained Al0.2Co1.5CrFeNi1.5Ti0.3 alloy via nano-sized, coherent precipitates. International Journal of Plasticity. 2018;100:177-191. DOI: 10.1016/j.ijplas.2017.10.005
  62. 62. Ming K, Bi X, Wang J. Precipitation strengthening of ductile Cr15Fe20Co35Ni20Mo10 alloys. Scripta Materialia. 2017;137:88-93. DOI: 10.1016/j.scriptamat.2017.05.019
  63. 63. Cao B, Kong H, Ding Z, Wu S, Luan J, Jiao Z, et al. A novel L12-strengthened multicomponent Co-rich high-entropy alloy with both high γ′-solvus temperature and superior high-temperature strength. Scripta Materialia. 2021;199:113826. DOI: 10.1016/j.scriptamat.2021.113826
  64. 64. Zhao Y, Yang T, Li Y, Fan L, Han B, Jiao Z, et al. Superior high-temperature properties and deformation-induced planar faults in a novel L12-strengthened high-entropy alloy. Acta Materialia. 2020;188:517-527. DOI: 10.1016/j.actamat.2020.02.028
  65. 65. Cao B, Yang T, W-h L, Liu C. Precipitation-hardened high-entropy alloys for high-temperature applications: A critical review. MRS Bulletin. 2019;44:854-859. DOI: 10.1557/mrs.2019.255
  66. 66. Zhao Y, Yang T, Han B, Luan J, Chen D, Kai W, et al. Exceptional nanostructure stability and its origins in the CoCrNi-based precipitation-strengthened medium-entropy alloy. Materials Research Letters. 2019;7:152-158. DOI: 10.1080/21663831.2019.1568315
  67. 67. Liu WH, Lu ZP, He JY, Luan JH, Wang ZJ, Liu B, et al. Ductile CoCrFeNiMo x high entropy alloys strengthened by hard intermetallic phases. Acta Materialia. 2016;116:332-342. DOI: 10.1016/j.actamat.2016.06.063
  68. 68. Bae JW, Park JM, Moon J, Choi WM, Lee B-J, Kim HS. Effect of μ-precipitates on the microstructure and mechanical properties of non-equiatomic CoCrFeNiMo medium-entropy alloys. Journal of Alloys and Compounds. 2019;781:75-83. DOI: 10.1016/j.jallcom.2018.12.040
  69. 69. Kwon H, Asghari-Rad P, Park JM, Sathiyamoorthi P, Bae JW, Moon J, et al. Synergetic strengthening from grain refinement and nano-scale precipitates in non-equiatomic CoCrFeNiMo medium-entropy alloy. Intermetallics. 2021;135:107212. DOI: 10.1016/j.intermet.2021.107212
  70. 70. He JY, Liu WH, Wang H, Wu Y, Liu XJ, Nieh TG, et al. Effects of Al addition on structural evolution and tensile properties of the FeCoNiCrMn high-entropy alloy system. Acta Materialia. 2014;62:105-113. DOI: 10.1016/j.actamat.2013.09.037
  71. 71. Wu Z, Bei H, Pharr GM, George EP. Temperature dependence of the mechanical properties of equiatomic solid solution alloys with face-centered cubic crystal structures. Acta Materialia. 2014;81:428-441. DOI: 10.1016/j.actamat.2014.08.026
  72. 72. Ma S, Zhang S, Qiao J, Wang Z, Gao M, Jiao Z, et al. Superior high tensile elongation of a single-crystal CoCrFeNiAl0. 3 high-entropy alloy by Bridgman solidification. Intermetallics. 2014;54:104-109. DOI: 10.1016/j.intermet.2014.05.018
  73. 73. Shun T-T, Du Y-C. Microstructure and tensile behaviors of FCC Al0. 3CoCrFeNi high entropy alloy. Journal of Alloys and Compounds. 2009;479:157-160. DOI: 10.1016/j.jallcom.2008.12.088
  74. 74. Hemphill MA, Yuan T, Wang G, Yeh J, Tsai C, Chuang A, et al. Fatigue behavior of Al0. 5CoCrCuFeNi high entropy alloys. Acta Materialia. 2012;60:5723-5734. DOI: 10.1016/j.actamat.2012.06.046
  75. 75. Kuznetsov AV, Shaysultanov DG, Stepanov ND, Salishchev GA, Senkov ON. Tensile properties of an AlCrCuNiFeCo high-entropy alloy in as-cast and wrought conditions. Materials Science and Engineering A. 2012;533:107-118. DOI: 10.1016/j.msea.2011.11.045
  76. 76. Ming K, Bi X, Wang J. Microstructures and deformation mechanisms of Cr 26 Mn 20 Fe 20 Co 20 Ni 14 alloys. Materials Characterization. 2017;134:194-201. DOI: 10.1016/j.matchar.2017.10.022
  77. 77. Tsai M-H, Yuan H, Cheng G, Xu W, Jian WW, Chuang M-H, et al. Significant hardening due to the formation of a sigma phase matrix in a high entropy alloy. Intermetallics. 2013;33:81-86. DOI: 10.1016/j.intermet.2012.09.022
  78. 78. Tsai MH, Tsai KY, Tsai CW, Chi L, Juan CC, Yeh JW. Criterion for sigma phase formation in Cr- and V-containing high-entropy alloys. Materials Research Letters. 2013;1:207-212. DOI: 10.1080/21663831.2013.831382
  79. 79. Zhang L, Huo X, Wang A, Du X, Zhang L, Li W, et al. A ductile high entropy alloy strengthened by nano sigma phase. Intermetallics. 2020;122:106813. DOI: 10.1016/j.intermet.2020.106813
  80. 80. Jo YH, Choi W-M, Sohn SS, Kim HS, Lee B-J, Lee S. Role of brittle sigma phase in cryogenic-temperature-strength improvement of non-equi-atomic Fe-rich VCrMnFeCoNi high entropy alloys. Materials Science and Engineering A. 2018;724:403-410. DOI: 10.1016/j.msea.2018.03.115
  81. 81. Zhu Y, Ameyama K, Anderson PM, Beyerlein IJ, Gao H, Kim HS, et al. Heterostructured materials: Superior properties from hetero-zone interaction. Materials Research Letters. 2020;9:1-31. DOI: 10.1080/21663831.2020.1796836
  82. 82. Ming K, Bi X, Wang J. Strength and ductility of CrFeCoNiMo alloy with hierarchical microstructures. International Journal of Plasticity. 2019;113:255-268. DOI: 10.1016/j.ijplas.2018.10.005
  83. 83. Gali A, George EP. Tensile properties of high- and medium-entropy alloys. Intermetallics. 2013;39:74-78. DOI: 10.1016/j.intermet.2013.03.018
  84. 84. Daoud H, Manzoni A, Völkl R, Wanderka N, Glatzel U. Microstructure and tensile behavior of Al8Co17Cr17Cu8Fe17Ni33 (at.%) high-entropy alloy. JOM journal of the minerals metals and materials. Society. 2013;65:1805-1814. DOI: 10.1007/s11837-013-0756-3
  85. 85. He JY, Wang H, Huang HL, Xu XD, Chen MW, Wu Y, et al. A precipitation-hardened high-entropy alloy with outstanding tensile properties. Acta Materialia. 2016;102:187-196. DOI: 10.1016/j.actamat.2015.08.076
  86. 86. Liu L, Zhang Y, Han J, Wang X, Jiang W, Liu CT, et al. Nanoprecipitate-strengthened high-entropy alloys. Advanced Science. 2021;8:2100870. DOI: 10.1002/advs.202100870
  87. 87. Li D, Li C, Feng T, Zhang Y, Sha G, Lewandowski JJ, et al. High-entropy Al 0.3 CoCrFeNi alloy fibers with high tensile strength and ductility at ambient and cryogenic temperatures. Acta Materialia. 2017;123:285-294. DOI: 10.1016/j.actamat.2016.10.038
  88. 88. Ma SG, Zhang SF, Qiao JW, Wang ZH, Gao MC, Jiao ZM, et al. Superior high tensile elongation of a single-crystal CoCrFeNiAl 0.3 high-entropy alloy by Bridgman solidification. Intermetallics. 2014;54:104-109. DOI: 10.1016/j.intermet.2014.05.018
  89. 89. Fu Z, Chen W, Wen H, Zhang D, Chen Z, Zheng B, et al. Microstructure and strengthening mechanisms in an FCC structured single-phase nanocrystalline Co 25 Ni 25 Fe 25 Al 7.5 Cu 17.5 high-entropy alloy. Acta Materialia. 2016;107:59-71. DOI: 10.1016/j.actamat.2016.01.050
  90. 90. Wu D, Zhang J, Huang J, Bei H, Nieh T-G. Grain-boundary strengthening in nanocrystalline chromium and the hall–Petch coefficient of body-centered cubic metals. Scripta Materialia. 2013;68:118-121. DOI: 10.1016/j.scriptamat.2012.09.025
  91. 91. Ming K, Lu W, Li Z, Bi X, Wang J. Amorphous bands induced by low temperature tension in a non-equiatomic CrMnFeCoNi alloy. Acta Materialia. 2020;188:354-365. DOI: 10.1016/j.actamat.2020.02.024
  92. 92. Beladi H, Timokhina I, Estrin Y, Kim J, De Cooman B, Kim S. Orientation dependence of twinning and strain hardening behaviour of a high manganese twinning induced plasticity steel with polycrystalline structure. Acta Materialia. 2011;59:7787-7799. DOI: 10.1016/j.actamat.2011.08.031
  93. 93. Gutierrez-Urrutia I, Raabe D. Dislocation and twin substructure evolution during strain hardening of an Fe–22 wt.% Mn–0.6 wt.% C TWIP steel observed by electron channeling contrast imaging. Acta Materialia. 2011;59:6449-6462. DOI: 10.1016/j.actamat.2011.07.009
  94. 94. Karaman I, Sehitoglu H, Beaudoin A, Chumlyakov YI, Maier H, Tome C. Modeling the deformation behavior of Hadfield steel single and polycrystals due to twinning and slip. Acta Materialia. 2000;48:2031-2047. DOI: 10.1016/S1359-6454(00)00051-3
  95. 95. Ma E. Amorphization in mechanically driven material systems. Scripta Materialia. 2003;49:941-946. DOI: 10.1016/s1359-6462(03)00477-9
  96. 96. Suryanarayana C. Mechanical alloying and milling. Progress in Materials Science. 2001;46:1-184. DOI: 0.1016/S0079-6425(99)00010-9
  97. 97. Sharma SM, Sikka S. Pressure induced amorphization of materials. Progress in Materials Science. 1996;40:1-77. DOI: 10.1016/0079-6425(95)00006-2
  98. 98. Sundeev R, Glezer A, Shalimova A. Phase transformations «Amorphization↔ crystallization» In metallic materials induced by severe plastic deformation. Reviews on Advanced Materials Science. 2018;54:93-105. DOI: 10.1515/rams-2018-0021
  99. 99. Lin ZJ, Zhuo MJ, Sun ZQ , Veyssière P, Zhou YC. Amorphization by dislocation accumulation in shear bands. Acta Materialia. 2009;57:2851-2857. DOI: 10.1016/j.actamat.2009.02.040
  100. 100. Li N, Wang YD, Lin Peng R, Sun X, Liaw PK, Wu GL, et al. Localized amorphism after high-strain-rate deformation in TWIP steel. Acta Materialia. 2011;59:6369-6377. DOI: 10.1016/j.actamat.2011.06.048
  101. 101. Zhao S, Hahn EN, Kad B, Remington BA, Wehrenberg CE, Bringa EM, et al. Amorphization and nanocrystallization of silicon under shock compression. Acta Materialia. 2016;103:519-533. DOI: 10.1016/j.actamat.2015.09.022
  102. 102. Zhao S, Kad B, Remington BA, LaSalvia JC, Wehrenberg CE, Behler KD, et al. Directional amorphization of boron carbide subjected to laser shock compression. Proceedings of the National Academy of Sciences. 2016;113:12088-12093. DOI: 10.1073/pnas.1604613113
  103. 103. Zhao S, Flanagan R, Hahn EN, Kad B, Remington BA, Wehrenberg CE, et al. Shock-induced amorphization in silicon carbide. Acta Materialia. 2018;158:206-213. DOI: 10.1016/j.actamat.2018.07.047
  104. 104. Johnson WL. Thermodynamic and kinetic aspects of the crystal to glass transformation in metallic materials. Progress in Materials Science. 1986;30:81-134. DOI: 10.1016/0079-6425(86)90005-8
  105. 105. Fecht HJ, Desré PJ, Johnson WL. Thermodynamic aspects of solid-state amorphization: Polymorphous melting Clapeyron diagram. Philosophical Magazine B. 2006;59:577-585. DOI: 10.1080/13642818908211178
  106. 106. Wolf D, Okamoto PR, Yip S, Lutsko JF, Kluge M. Thermodynamic parallels between solid-state amorphization and melting. Journal of Materials Research. 2011;5:286-301. DOI: 10.1557/jmr.1990.0286
  107. 107. Koike J. Elastic instability of crystals caused by static atom displacement: A mechanism for solid-state amorphization. Physical Review B. 1993;47:7700. DOI: 10.1103/PhysRevB.47.7700
  108. 108. Li M, Johnson WL. Instability of metastable solid solutions and the crystal to glass transition. Physical Review Letters. 1993;70:1120-1123. DOI: 10.1103/PhysRevLett.70.1120
  109. 109. Delogu F. Connection between shear instability and amorphisation. Materials Science and Engineering A. 2004;367:162-165. DOI: 10.1016/j.msea.2003.10.248
  110. 110. Zhao S, Li Z, Zhu C, Yang W, Zhang Z, Armstrong D, et al. Amorphization in extreme deformation of the CrMnFeCoNi high-entropy alloy. Science Advances. 2021;7:eabb3108. DOI: 10.1126/sciadv.abb3108
  111. 111. Wang H, Chen D, An X, Zhang Y, Sun S, Tian Y, et al. Deformation-induced crystalline-to-amorphous phase transformation in a CrMnFeCoNi high-entropy alloy. Science Advances. 2021;7:eabe3105. DOI: 10.1126/sciadv.abe3105
  112. 112. Ji W, Wu MS. Nanoscale origin of the crystalline-to-amorphous phase transformation and damage tolerance of Cantor alloys at cryogenic temperatures. Acta Materialia. 2022;226:117639. DOI: 10.1016/j.actamat.2022.117639
  113. 113. Li J, Chen H, He Q, Fang Q, Liu B, Jiang C, et al. Unveiling the atomic-scale origins of high damage tolerance of single-crystal high entropy alloys. Physical Review Materials. 2020;4:103612. DOI: 10.1103/PhysRevMaterials.4.103612
  114. 114. Ming K, Li B, Bai L, Jiang P, Wu X, Zheng S, et al. Dynamically reversible shear transformations in a CrMnFeCoNi high-entropy alloy at cryogenic temperature. Acta Materialia. 2022;232:117937. DOI: 10.1016/j.actamat.2022.117937
  115. 115. Ming K, Li L, Li Z, Bi X, Wang J. Grain boundary decohesion by nanoclustering Ni and Cr separately in CrMnFeCoNi high-entropy alloys. Science Advances. 2019;5:eaay0639. DOI: 10.1126/sciadv.aay0639
  116. 116. Li L, Kamachali RD, Li Z, Zhang Z. Grain boundary energy effect on grain boundary segregation in an equiatomic high-entropy alloy. Physical Review Materials. 2020;4:053603. DOI: 10.1103/PhysRevMaterials.4.053603
  117. 117. Li L, Li Z. Aging induced segregation and nanoprecipitation in a severely deformed equiatomic high-entropy alloy. Materials Characterization. 2020;165:110369. DOI: 10.1016/j.matchar.2020.110369
  118. 118. Li L, Li Z, Kwiatkowski da Silva A, Peng Z, Zhao H, Gault B, et al. Segregation-driven grain boundary spinodal decomposition as a pathway for phase nucleation in a high-entropy alloy. Acta Materialia. 2019;178:1-9. DOI: 10.1016/j.actamat.2019.07.052
  119. 119. Otto F, Dlouhý A, Pradeep KG, Kuběnová M, Raabe D, Eggeler G, et al. Decomposition of the single-phase high-entropy alloy CrMnFeCoNi after prolonged anneals at intermediate temperatures. Acta Materialia. 2016;112:40-52. DOI: 10.1016/j.actamat.2016.04.005
  120. 120. Pickering EJ, Muñoz-Moreno R, Stone HJ, Jones NG. Precipitation in the equiatomic high-entropy alloy CrMnFeCoNi. Scripta Materialia. 2016;113:106-109. DOI: 10.1016/j.scriptamat.2015.10.025
  121. 121. Schuh B, Mendez-Martin F, Völker B, George EP, Clemens H, Pippan R, et al. Mechanical properties, microstructure and thermal stability of a nanocrystalline CoCrFeMnNi high-entropy alloy after severe plastic deformation. Acta Materialia. 2015;96:258-268. DOI: 10.1016/j.actamat.2015.06.025

Written By

Kaisheng Ming, Shijian Zheng and Jian Wang

Submitted: 04 April 2022 Reviewed: 05 April 2022 Published: 06 June 2022