Next Article in Journal
Research on Cement-Free Grouting Material for Shield Tunneling in Water-Rich Karst Regions
Next Article in Special Issue
Synergistic Effect of Cu Addition and Pre-Straining on the Natural Aging and Artificial Age-Hardening Behavior of AA6111 Alloy
Previous Article in Journal
Review of Layered Double Hydroxide (LDH) Nanosheets in Corrosion Mitigation: Recent Developments, Challenges, and Prospects
Previous Article in Special Issue
Spontaneous Infiltration Behavior of Al Matrix into Carbon Fiber Bundles Induced by Al-Ni Binary Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure Evolution and Mechanical Properties of Dual-Phase AlCrFe2Ni2 High-Entropy Alloy Under High-Strain-Rate Compression

1
Innovation & Interdisciplinary Institute of Low Carbon Metallurgical Engineering, School of Materials and Energy, Guangdong University of Technology, Guangzhou 510006, China
2
School of Materials Science and Engineering, North University of China, Taiyuan 030051, China
3
School of Iron and Steel, Soochow University, Suzhou 215021, China
4
School of Materials Science and Engineering, South China University of Technology, Guangzhou 510640, China
*
Author to whom correspondence should be addressed.
Materials 2025, 18(6), 1191; https://doi.org/10.3390/ma18061191
Submission received: 23 January 2025 / Revised: 27 February 2025 / Accepted: 4 March 2025 / Published: 7 March 2025

Highlights

  1. Showed the excellent dynamic compression properties in AlCrFe2Ni2 alloy.
  2. Observing significant strain rate dependence of yield strength at high strain rates.
  3. The adiabatic shear bands elucidate thermal softening effects in AlCrFe2Ni2 alloy.

Abstract

:
This paper investigates the effect of strain rate on the mechanical deformation and microstructural development of dual-phase AlCrFe2Ni2 high-entropy alloy during quasi-static and dynamic compression processes. It is revealed that the as-cast AlCrFe2Ni2 alloy is composed of a mixture of FCC, disordered BCC, and ordered B2 crystal structure phases. The alloy shows excellent compressive properties under quasi-static and dynamic deformation. The yield strength exceeds 600 MPa while the compressive strength is more than 3000 MPa at the compression rates of 30% under quasi-static conditions. Under dynamic compression conditions, the ultimate compression stresses are 1522 MPa, 1816 MPa, and 1925 MPa with compression strains about 12.8%, 14.7%, and 18.2% at strain rates of 1300 s−1, 1700 s−1 and 2100 s−1, respectively. The dynamic yield strength is approximately linear with strain rate within the specified range and exhibit great sensitivity. The strong localized deformation regions (i.e., adiabatic shear bands (ASBs)) appear in dynamically deformed samples by dynamic recrystallization due to the conflicting processes of strain rate hardening and heat softening.

1. Introduction

For conventional alloys, such as titanium alloys, magnesium alloys, and aluminum alloys, they are traditionally designed as one or two principal elements with minor additions of the other elements. Recently, a new alloy system, i.e., high-entropy alloys (HEAs), was designed differently from traditional counterparts. For HEAs, four or more principal elements are present in proportions of 5% to 35%, with the mixing entropy more than 1.5R [1,2]. HEAs usually form a simple solid solution phase [1], for instance, BCC, FCC, and HCP phases. These phases can either exist alone or as a dual-phase coexistence [2]. Based on previous research, single-phase HEAs with a BCC structure exhibit low plasticity properties but high tensile strength, whereas HEAs with a single-phase FCC structure typically have good plasticity but low tensile strength [3,4,5]. In order to balance the properties of HEAs, novel eutectic high-entropy alloys (EHEAs) consisting of a composite microstructure with FCC and BCC phases were designed and proposed by Lu et al. [6]. Until now, great potential has been recognized in many applications for these EHEAs, owing to their excellent fracture toughness [7], good cryogenic properties [7,8], exceptional wear, and fatigue resistances [9,10]. In addition, it is simple to adapt EHEAs with good flowability, castability, and high strength and ductility to large-scale industrial manufacturing [11,12].
Currently, several EHEA systems have been developed, including AlCoCrFeNi2.1 [6], Al17Ni34Ti17V32 [13], AlCr1.3TiNi2 [14], and AlCrCuFeNi2 [15]. Owing to better practicality and low cost, the AlCrFe2Ni2 alloy, with similar microstructure and properties to the AlCoCrFeNi2.1 alloy, was investigated. It is indicated that the AlCrFe2Ni2 alloy consists of the FCC and A2/B2 spinodal decomposition phase and presents better fracture strength and ductility [16]. In addition, the dynamic mechanical response of HEAs has also received a lot of attention. To use HEAs in extreme industrial environments, there is an urgent need for a comprehensive understanding of their performance and mechanisms during high-speed compression [17,18,19,20,21].
According to the available literature, in addition to the dislocation-related mechanism, twinning mechanism, grain boundary-related mechanism, diffusion creep mechanism, and dynamic recrystallization mechanism, the shear bands are formed in HEAs at high strain rates or under large deformation conditions [17,18,19,20,21]. The adiabatic shear band (ASB) is a localized region formed by a very large shear deformation in a very short time. In this region, along with the adiabatic temperature rise, the microstructure and properties of the material undergo significant changes, such as grain refinement and an increase in dislocation density. The ASB is an ideal mode of deformation, and the presence of the ASB is conducive to the enhancement of the penetration properties of the material. These traits differ significantly from those that occur when quasi-static characteristics are present. The dynamic behavior of Al0.6CoCrFeNi biphasic (FCC and BCC) alloys was examined by Wang et al. [22]. The findings demonstrated that the yield strength could be clearly increased by increasing the strain rate. Tian et al. found that the AlCoCrFeNi alloy exhibited extremely high strain rate sensitivity by investigating the alloy at different strain rates [23]. The dynamic deformation mechanism of the CrCoFeNi high-entropy alloy was investigated by Cao et al. [24]. It was found that massive dislocation nucleation and drag increased the strain rate sensitivity of the alloy at high strain rates, and the alloy also exhibited significant strain-hardening capability. Yuan et al. prepared Ni32Co30Cr10Fe10Al18 EHEAs by laser metal deposition, and it was found that the presence of the FCC phase caused the alloy to exhibit high ductility and strain-hardening ability at high temperatures and high strain rates [25]. Jeong Min Park et al. [26] found that the CoCrFeMnNi alloy showed strong strain rate sensitivity. Compression experiments on the AlCrCuFeNi2 alloy at a strain rate of 10−3 s−1~3000 s−1 were conducted by Ma et al. [15]. They discovered that over a range of strain rates there was a roughly linear connection between dynamic yield strength and strain rate. N. Kumar et al. [27] conducted compression tests on the Al0.1CrCoFeNi alloy using split-Hopkinson pressure bar (SHPB) and noted notable work hardening and strong strain rate sensitivity, as well as a large number of twinned grains [28].
Herein, in the present study, the primary emphasis is to provide more detailed insights into the compression behavior of the dual-phase AlCrFe2Ni2 alloy under both static compression and dynamic impact. In this context, the impact of strain rates on the microstructural development and mechanical characteristics of the AlCrFe2Ni2 alloy is assessed using the SHPB technique. Correspondingly, these fundamental processes for the development of microstructural features and mechanical characteristics are also examined and explained.

2. Experimental

2.1. Alloy Fabrication

For the present study, the experimental AlCrFe2Ni2 alloys were fabricated by using a vacuum high-frequency induction melting furnace, and all raw elemental metals have purity levels exceeding 99.9 wt.%. Melting and casting were conducted by backfilling with argon (Ar) under vacuum pressure up to 3 × 10−3 Pa at 0.01 atm.

2.2. Property Testing

For the static tests, cylindrical specimens that were about 10 mm tall and 5 mm in diameter were wire-cut and then ground. The deformation behavior of AlCrFe2Ni2 alloy at the strain rate in the range of 10−4 s−1~100 s−1 was examined using a universal testing machine (AGS-X Plus100 KN, Japan), three samples being tested at each strain rate.
Following compression tests, samples were sliced along the stress loading direction, and silicon carbide paper with mechanical thinning was used to mechanically polish the surface parallel to the compression direction. EBSD analysis was used for microstructural characterization and orientation imaging microscope analysis. Samples measuring around 8 mm in diameter and 5 mm in height were prepared for dynamic compression. An SHPB with strain rates ranging from 1300 s−1 to 2100 s−1 was used for dynamic compression experiments at ambient temperature. The force applied to the deformation zone when the cylindrical sample is dynamically loaded by the SHPB was computed using the information gathered by the strain gauges on the incident and transmitted bars. The function of time t was used to calculate the strain rate ε , true strain ε and true stress σ . The following is how the equations are expressed: [29]:
ε = 2 C 0 L 0 ε r t
ε = 2 C 0 L 0 0 l ε r t d t
σ = A 0 A Ε ε t t
where Ε is the elastic modulus of the incident and transmission bars, and A 0 . And C 0 are the wave propagation velocities in the bars. A is the cross-section area of the incident and transmission bars, and L 0 is the height and cross-section area of sample. ε r t and ε t t are the experimentally measured strains of the incident and transmitted stress pulses on the SHPB. You can find more information about the SHPB and the stress and strain calculations elsewhere.

2.3. Microstructure Characterization

An X-ray diffractometer (XRD, EMPYREAN with Cu radiation target) was used to identify the phase. It has a scanning rate of 4° min−1 and a 2θ scanning range of 20° to 100°. Energy dispersive spectrometry (EDS) in conjunction with scanning electron microscopy (SEM, JEOL-JSM-IT800) and transmission electron microscopy (TEM, Talos F200S) were utilized to describe the microstructural evolution and deformation process of the AlCrFe2Ni2 alloy. Metallographic polishing and twin-jet electropolishing (electrolyte: 90 vol.% ethanol + 10 vol.% perchloric acid) were used to produce the thin-foil sample for TEM analysis. Samples were prepared by mechanically thinning after normal metallographic polishing for electron backscatter diffraction (EBSD) examination.

3. Results and Discussion

3.1. Microstructural Analysis

Figure 1 shows the X-ray diffraction pattern of the as-cast AlCrFe2Ni2 alloy. The XRD result indicates that the alloy is composed of a mixture of FCC, disordered BCC, and ordered B2 crystal structures. In addition, due to the same crystal structure and similar lattice parameter, the pattern presents the overlapping disordered BCC and B2 Bragg reflection, which was also found in the other HEAs [30].
Figure 2 shows the SEM images of the as-cast AlCrFe2Ni2 alloy. From Figure 2a, it can be seen that the alloy mainly consists of noodle-like phases and triangular phases. Further observed by the high-magnified SEM image in Figure 2b, the length and size of the noodle-like phases are not uniform, the other phases are not always triangular in shape, and there are various other shapes. Figure 2c,d shows enlarged images of Figure 2b. The area of the triangle that can be seen in Figure 2c is composed of interconnected light and dark microstructures, which is characteristic of the microstructure of the spinodal decomposition microstructure. The noodle-like phases are clearly visible in Figure 2d, and there are fine dark phases present between the noodle-like phases [16,31,32].
Figure 3 shows the EBSD phase diagram and IPF diagram of the as-cast AlCrFe2Ni2 alloy. From Figure 3a, it can be seen that the alloy is mainly formed by the FCC phase and BCC/B2 phase, which is consistent with the XRD analysis results, where the FCC phase accounts for 63.7% and the BCC phase accounts for 36.3%. Combined with Figure 2, the noodle-like phase is the FCC phase, and the other phases are the BCC and B2 phases. As can be seen from Figure 3b, the grain orientations of the noodle-like phases are disordered, whereas the orientations of the alternating phases are partly uniform and partly disordered, which suggests that the remaining phases are composed of the BCC phase and the B2 phase.
Further detailed analysis on the microstructure of the AlCrFe2Ni2 alloy was carried out by TEM, as seen in Figure 4. Figure 4a,c shows bright-field images of the AlCrFe2Ni2 alloy. It is evident that the alloy is composed of noodle-like phases and a spinodal decomposition microstructure, which is an alternating mixture of the two phases. Figure 4b,d shows the selected area electron diffraction (SAED) pattern of the alloy AlCrFe2Ni2. From Figure 4b and the dark-field image (Figure 4c), it can be obtained that the noodle-like phase is the FCC phases, which is also in agreement with the results of the EBSD phase map. From Figure 4d, it can be seen that the modulated structure is a mixture consisting of BCC and B2 phases, and combined with the dark-field image (Figure 4f), in the modulated structure, the matrix has a BCC structure and the particle-shaped phase has a B2 structure. Figure 5b–f shows the element distribution of the AlCrFe2Ni2 alloy, from which it can be seen that the AlCrFe2Ni2 alloy forms BCC (Cr, Fe)-rich inter-dendritic matrices, and B2 (Al, Ni)-rich particles (similar to those in Ref. [33]).

3.2. Mechanical Characteristics Under Compression Testing

Figure 6 shows the true stress–strain curve of the AlCrFe2Ni2 alloy at different strain rates. The true compression stress–strain curves of the AlCrFe2Ni2 alloy under quasi-static deformation at room temperature are shown in Figure 6a. From Figure 6a, it can be seen that when the strain rate is 10−4 s−1–100 s−1, the alloy shows significant work hardening, but there is no significant change in the yield strength value with increasing strain rate. The inset in Figure 6a shows the sample after compression, and the sample did not fracture, which indicates that the AlCrFe2Ni2 alloy has good plasticity. Thus, the maximum strain is taken to be 30%, and the corresponding ultimate compression stress are more than 3000 MPa. Correspondingly, the compression properties of the AlCrFe2Ni2 alloy are shown in Table 1. Figure 6b presents the compression performance of the AlCrFe2Ni2 alloy at high strain rates. It can be seen from Figure 6b that the yield strength changes significantly with the increase in strain rate. Furthermore, a positive strain rate effect is indicated by the dynamic yield stress, and the flow stress shows an upward trend with the rise of strain rate. The compression properties of the AlCrFe2Ni2 alloy are shown in Table 2. In conjunction with the accompanying graph from Figure 6b, because of the transient loading time, the surface of the AlCrFe2Ni2 alloy has not changed in any way, the alloy has failed, and therefore severe changes should have occurred within the alloy at high strain rates [34].
The strain rate sensitivity (m) is used to describe the dependence of strength/stress on strain rate. It can be defined by the slope of the log strength σ versus log strain rate ε at constant strain :
m = ln σ ln ε
The yield strengths are roughly 606 MPa and 658 MPa at strain rates of 10−4 s−1 and 100 s−1, respectively. The AlCrFe2Ni2 alloy m under quasi-static deformation is estimated to be around 0.0089. It is worth noting that it is of the same order of magnitude as the conventional metals: Cu (~0.006) and Ni (~0.0028) [34], but an order of magnitude lower than that of partially eutectic high-entropy alloys (~0.028) [35]. The yield strengths are roughly 892 MPa and 1132 MPa at strain rates of 1300 s−1 and 2100 s−1, respectively, and the m of the AlCrFe2Ni2 alloy is estimated to be about 0.42 at high strain rates, which is much higher than the value under quasi-static deformation, and is also consistent with the image pattern of the true stress–strain curve. The present reason for the sudden increase in the strain rate sensitivity coefficient may be due to phonon drag effects [24,36]. Elastic lattice vibrations that go through crystals are called phonons. The phonon drag effect is the result of the interaction between phonons and dislocations, which creates viscous resistance to dislocation sliding. The phonon drag effect is known to be caused by a number of ways [18].
The yield strength for the AlCrFe2Ni2 alloy is plotted against the logarithmic strain rate in Figure 7. It can be seen from Figure 7 that the trend of yield strength tends to a straight line when the strain rate is 10−4 s−1–103 s−1, and when the strain rate is greater than 103 s−1, the yield strength of the alloy has a clear upward trend with the increase in strain rate. It is evident that under dynamic conditions, the effect of strain rate on yield strength is significantly higher. They are therefore divided into two regions where the strength fluctuates with the strain rate: the area where phonon drag influences dislocation motions ( ε ˙ > 103 s−1) and the area of thermally induced dislocation glide ( ε ˙ < 103 s−1) [20,27,37,38]. In the region of ε ˙ < 103 s−1, the dislocations engage with the surrounding regions’ isotropic phonon flux, and the dislocation velocity is so low that the phonon drag effect barely affects the dislocation motion [18,27]. In the region of ε ˙ >103 s−1, viscous phonon drag significantly affects dislocation motion [18,27,39]. Fast moving dislocations would cause phonon scattering through the strain field. Moreover, dislocations could absorb phonon energy, which leads to phonon scattering, and the kinematic viscosity of dislocations increases due to the phonon dispersion.

3.3. Microstructure Evolution

The IPF and IQ + misorientation border maps (BMs) at a strain rate of 1300 s−1 and 2100 s−1 are displayed in Figure 8, respectively. From Figure 8a, it can be seen that there is a distinct black area in the center, which is an ASB region, while in Figure 8c, only smaller discontinuous ASB zones can be observed. In addition, a large number of highly refined submicron grains can be seen in the red-boxed portion of Figure 8a,c. In Figure 8c, although many ASB regions are observed, the number of fine grains with HAGBs is significantly less than that in the regions with a strain rate of 1300 s−1. The dynamic recrystallization (DRX) is the cause of the grain refinement observed in the dynamically deformed materials, such as the ASBs frequently seen in a variety of deformation materials, which may be the source of DRX [38,40,41]. The ASB at 1300−1 s−1 is a narrow region of high-concentration fine grains. These ASBs cause cracks to form, and eventually shear damage, which explains the poor performance of strain rate at 1300−1 s−1. The emergence of ASBs may be due to the conflicting processes of strain rate hardening and heat softening [18]. These two processes, strain-hardening and heat-softening effects brought on by dislocation proliferation/stacking or twin evolution, would compete with one another during dynamic deformation. The strain-hardening effect would eventually be overcome by the contribution of thermal softening to the flow stresses, which would result in the production of ASBs. As can be seen from Figure 8b,d, there are a large number of blue regions at high strain rates. The blue region indicates high-angle grain boundaries (HAGBs) with dislocation angles more than 15°, whereas the red and green lines correspond to low-angle grain boundaries (LAGBs) less than 15°. LAGBs provide favorable conditions for dislocation proliferation due to their special structure and higher energy states, which explains the increase in yield strength at high strain rates.
The work-hardening rate (WHR, d σ /d ε , where σ is the flow stress and ε is the plastic strain) behavior as a function of the plastic strain at a strain rate of 1300 s−1 and 2100 s−1 is displayed in Figure 9. It has been recognized that higher strain rates produce a larger WHR, especially in the initial plasticity stage. This characterization can be explained by the expression [42]:
ν ¯ = A τ m
where A and m are the constants for specific materials and ν ¯ is the dislocation movement velocity. The initial shear stress for dislocation movement is denoted by τ . The WHR and the strain rates are positively correlated, as seen in Figure 9. The higher strain rates result in the larger shear stress and the higher dislocation motion rate. Two distinct zones can be seen inside the curve. In one region (true plastic strain < 2%), the strain-hardening rate (SHR) drops significantly as the strain increases. In this region, the strain rate is proportional to the work-hardening rate, which is related to the formation of adiabatic shear bands. At a strain rate of 2100 s−1, although there is also the formation of ASBs, the distribution of the shear bands of the fine scattering does not cause the alloy to fail quickly, but rather has a certain strengthening effect. In another region (true plastic strain > 2%), the strain-hardening rate drops less steeply; the plastic deformation in this region is dominated by dislocation slip and the rising dislocation annihilation rate, but the SHR does not rise due to the effect of thermal softening [43].

4. Conclusions

In this work, the effect of strain rate on the microstructural evolution and compression behavior of the dual-phase AlCrFe2Ni2 alloy was studied. The main conclusions are as follows:
  • The as-cast AlCrFe2Ni2 alloy is composed of a mixture of Fe,Ni-enriched FCC phases, disordered Fe,Cr-enriched BCC phases, and ordered Al,Ni-enriched B2 phases.
  • The AlCrFe2Ni2 alloy shows excellent compressive properties under quasi-static deformation. The yield strength exceeds 600 MPa while the compressive strength is more than 3000 MPa at the compression rates of 30%.
  • The ultimate compression stresses are 1522 MPa, 1816 MPa, and 1925 MPa with compression strains about 12.8%, 14.7%, and 18.2% at strain rates of 1300 s−1, 1700 s−1, and 2100 s−1, respectively. Strain rate sensitivity ranges from 0.0089 at low strain (10−4 s−1–100 s−1) to 0.42 at high strain rates (1300 s−1–2100 s−1), which is due to the phonon drag effect at high strains.
  • At high-strain-rate compression, a adiabatic shear phenomenon occurs. The adiabatic shear bands formed by dynamic recrystallization due to the conflicting processes of strain rate hardening and heat softening.

Author Contributions

Methodology, H.Y., Y.W. and S.D.; Validation, Y.W., S.L. and S.D.; Formal analysis, S.D.; Investigation, H.Y., X.G. and S.D.; Resources, Y.D., S.L. and L.M.; Data curation, S.L.; Writing—original draft, H.Y.; Writing—review & editing, X.G. and S.L.; Visualization, Y.W., X.G., S.D. and L.M.; Supervision, L.M.; Project administration, Y.D.; Funding acquisition, Y.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Natural Science Foundation of Guangdong Province (Nos. 2022A0505050052 and 2022A1515012591).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. 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 Elements: Novel Alloy Design Concepts and Outcomes. Adv. Eng. Mater. 2004, 6, 299–303. [Google Scholar] [CrossRef]
  2. Li, W.D.; Liaw, P.K.; Gao, Y. Fracture resistance of high entropy alloys: A review. Intermetallics 2018, 99, 69–83. [Google Scholar] [CrossRef]
  3. Liu, W.H.; He, J.Y.; Huang, H.L.; Wang, H.; Lu, Z.P.; Liu, C.T. Effects of Nb additions on the microstructure and mechanical property of CoCrFeNi high-entropy alloys. Intermetallics 2015, 60, 1–8. [Google Scholar] [CrossRef]
  4. 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]
  5. Senkov, O.N.; Wilks, G.B.; Scott, J.M.; Miracle, D.B. Mechanical properties of NbMoTaW and VNbMoTaW refractory high entropy alloys. Intermetallics 2011, 19, 698–706. [Google Scholar] [CrossRef]
  6. Lu, Y.P.; Dong, Y.; Guo, S.; Jiang, L.; Kang, H.J.; Wang, T.M.; Wen, B.; Wang, Z.J.; Jie, J.C.; Ruan, H.H.; et al. A Promising New Class of High-Temperature Alloys: Eutectic High-Entropy Alloys. Sci. Rep. 2014, 4, 6200. [Google Scholar] [CrossRef]
  7. Gludovatz, B.; Hohenwarter, A.; Catoor, D.; Chang, E.H.; George, E.P.; Ritchie, R.O. A fracture-resistant high-entropy alloy for cryogenic applications. Science 2014, 345, 1153–1158. [Google Scholar] [CrossRef]
  8. Otto, F.; Dlouh, A.; Somsen, C.H.; Bei, H.; Eggeler, G.; George, E.P. The influences of temperature and microstructure on the tensile properties of a CoCrFeMnNi high-entropy alloy. Acta Mater. 2013, 61, 5743–5755. [Google Scholar] [CrossRef]
  9. Chuang, M.H.; Tsai, M.H.; Wang, W.R.; Lin, S.J.; Yeh, J.W. Microstructure and wear behavior of AlxCo1.5CrFeNi1.5Tiy high-entropy alloys. Acta Mater. 2011, 59, 6308–6317. [Google Scholar] [CrossRef]
  10. 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]
  11. Lu, Y.P.; Dong, Y.; Jiang, H.; Wang, Z.J.; Cao, Z.Q.; Guo, S.; Wang, T.M.; Li, T.J.; Liaw, P.K. Promising properties and future trend of eutectic high entropy alloys. Scr. Mater. 2020, 187, 202–209. [Google Scholar] [CrossRef]
  12. Lu, Y.P.; Gao, X.Z.; Jiang, L.; Chen, Z.N.; Wang, T.M.; Jie, J.C.; Kang, H.J.; Zhang, Y.B.; Guo, S.; Ruan, H.H. Directly cast bulk eutectic and near-eutectic high entropy alloys with balanced strength and ductility in a wide temperature range. Acta Mater. 2017, 124, 143–150. [Google Scholar] [CrossRef]
  13. Wang, M.L.; Lu, Y.P.; Lan, J.G.; Wang, T.M.; Zhang, C.; Cao, Z.Q.; Li, T.J.; Liaw, P.K. Lightweight, ultrastrong and high thermal-stable eutectic high-entropy alloys for elevated-temperature applications. Acta Mater. 2023, 248, 118806. [Google Scholar] [CrossRef]
  14. Wang, M.L.; Lu, Y.P.; Lan, J.G.; Wang, T.M.; Zhang, C.A.; Cao, Z.Q.; Li, T.J.; Liaw, P.K. A novel bulk eutectic high-entrogy alloy with outstanding as-cast specific yield strengths at elevated temperatures. Scr. Mater. 2021, 204, 114132. [Google Scholar] [CrossRef]
  15. Ma, S.G.; Jiao, Z.M.; Qiao, J.W.; Yang, H.J.; Zhang, Y.; Wang, Z.H. Strain rate effects on the dynamic mechanical properties of the AlCrCuFeNi2 high-entropy alloy. Mater. Sci. Eng. A 2016, 649, 35–38. [Google Scholar] [CrossRef]
  16. Dong, Y.; Gao, X.X.; Lu, Y.P.; Wang, T.M.; Li, T.J. A multi-component AlCrFe2Ni2 alloy with excellent mechanical properties. Mater. Lett. 2016, 169, 62. [Google Scholar] [CrossRef]
  17. Li, Z.; Zhao, S.; Diao, H.; Liaw, P.K.; Meyers, M.A. High-velocity deformation of Al0.3CoCrFeNi high-entropy alloy: Remarkable resistance to shear failure. Sci. Rep. 2017, 7, 42742. [Google Scholar] [CrossRef]
  18. Meyers, M.A. (Ed.) Dynamic Behavior of Materials; John Wiley and Sons: Hoboken, NJ, USA, 1994. [Google Scholar]
  19. Cho, K.M.; Lee, S.H.; Nutt, S.R.; Duffy, J. Adiabatic shear band formation during dynamic torsional deformation of an HY-100 steel. Acta Mater. 1993, 41, 923–932. [Google Scholar] [CrossRef]
  20. Lee, H.S.; Sohn, S.S.; Jeon, C.W.; Jo, I.G.; Lee, S.K.; Lee, S.H. Dynamic compressive deformation behavior of SiC-particulate-reinforced A356 Al alloy matrix composites fabricated by liquid pressing process. Mater. Sci. Eng. A 2017, 680, 368–377. [Google Scholar] [CrossRef]
  21. Dirras, G.; Couque, H.; Lilensten, L.; Heczel, A.; Tingaud, D.; Couzinié, J.-P.; Perriére, L.; Gubicza, J.; Guillot, I. Mechanical behavior and microstructure of Ti20Hf20Zr20Ta20Nb20 high-entropy alloy loaded under quasi-static and dynamic compression conditions. Mater. Charact. 2016, 111, 106–113. [Google Scholar] [CrossRef]
  22. Wang, L.; Qiao, J.W.; Ma, S.G.; Jiao, Z.M.; Zhang, T.W.; Chen, G.; Zhao, D.; Zhang, Y.; Wang, Z.H. Mechanical response and deformation behavior of Al0.6CoCrFeNi high-entropy alloys upon dynamic loading. Mater. Sci. Eng. 2018, 727, 208–213. [Google Scholar] [CrossRef]
  23. Tian, L.; Jiao, Z.M.; Yuan, G.Z.; Ma, S.G.; Wang, Z.H.; Yang, H.J.; Zhang, Y.; Qiao, J.W. Effect of Strain Rate on Deformation Behavior of AlCoCrFeNi High-Entropy Alloy by Nanoindentation. Mater. Eng. Perform. 2016, 25, 2255–2260. [Google Scholar] [CrossRef]
  24. Cao, T.Q.; Zhang, Q.; Wang, L.; Wang, L.; Xiao, Y.; Yao, J.H.; Liu, H.Y.; Ren, Y.; Liang, J.; Xue, Y.F. Dynamic deformation behaviors and mechanisms of CoCrFeNi high-entropy alloys. Acta Mater. 2023, 260, 119343. [Google Scholar] [CrossRef]
  25. Yuan, K.B.; Yao, X.H.; Yu, Y.Q.; Wang, R.F.; Chai, Z.S.; Zhou, K.X.; Wang, Z.J. Dynamic thermomechanical response and constitutive modeling of eutectic high-entropy alloy. Int. J. Mech. Sci. 2023, 246, 108148. [Google Scholar] [CrossRef]
  26. Park, J.M.; Moon, J.; Bea, J.W.; Jang, M.J.; Park, J.; Lee, S.; Kim, H.S. Strain rate effects of dynamic compressive deformation on mechanical properties and microstructure of CoCrFeMnNi high-entropy alloy. Mater. Sci. Eng. 2018, 719, 155–163. [Google Scholar] [CrossRef]
  27. Kumar, N.; Ying, Q.; Nie, X.; Mishra, R.S.; Tang, Z.; Liaw, P.K.; Brenan, R.E.; Doherty, K.J.; Cho, K.C. High strain-rate compressive deformation behavior of the Al0.1CrFeCoNi high entropy alloy. Mater. Des. 2015, 86, 598–602. [Google Scholar] [CrossRef]
  28. 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]
  29. Wang, B.F.; Fu, A.; Huang, X.X.; Liu, B.; Liu, Y.; Li, Z.Z.; Zan, X. Mechanical properties and microstructure of the CoCrFeMnNi high entropy alloy under high strain rate compression. Mater. Eng. Perform. 2016, 25, 2985–2992. [Google Scholar] [CrossRef]
  30. Dong, Y.; Zhou, K.Y.; Lu, Y.P.; Gao, X.X.; Wang, T.M.; Li, T.J. Effect of vanadium addition on the microstructure and properties of AlCoCrFeNi high entropy alloy. Mater. Des. 2014, 57, 67–72. [Google Scholar] [CrossRef]
  31. Singh, S.; Wanderka, N.; Murty, B.S.; Glatzel, U.; Banhart, J. Decomposition in multi-component AlCoCrCuFeNi high-entropy alloy. Acta Mater. 2011, 59, 182–190. [Google Scholar] [CrossRef]
  32. Santodonato, L.J.; Zhang, Y.; Feygenson, M.; Parish, C.M.; Gao, M.C.; Weber, R.J.K.; Neuefeind, J.C.; Tang, Z.; Liaw, P.K. Deviation from high-entropy configurations in the atomic distributions of a multi-principal-element alloy. Nat. Commun. 2015, 6, 5964. [Google Scholar] [CrossRef]
  33. Linden, Y.; Pinkas, M.; Munitz, A.; Meshi, L. Long-period antiphase domains and short-range order in a B2 matrix of the AlCoCrFeNi high-entropy alloy. Scr. Mater. 2017, 139, 49–52. [Google Scholar] [CrossRef]
  34. Moon, J.; Hong, S.I.; Bae, J.W.; Jang, M.J.; Yim, D.; Kim, H.S. On the strain rate-dependent deformation mechanism of CoCrFeMnNi high-entropy alloy at liquid nitrogen temperature. Mater. Res. Lett. 2017, 5, 472–477. [Google Scholar] [CrossRef]
  35. Tang, Y.; Wang, R.X.; Xiao, B.; Zhang, Z.R.; Li, S.; Qiao, J.W.; Bai, S.X.; Zhang, Y.; Liaw, P.K. A review on dynamic-mechanical behaviors of high entropy alloys. Prog. Mater. Sci. 2023, 135, 101090. [Google Scholar] [CrossRef]
  36. Hong, S.I.; Moon, J.; Hong, S.K.; Kim, H.S. Thermally activated deformation and the rate controlling mechanism in CoCrFeMnNi high entropy alloy. Mater. Sci. Eng. A 2017, 682, 569–576. [Google Scholar] [CrossRef]
  37. Hines, J.A.; Vecchio, K.S. Recrystallization kinetics within adiabatic shear bands. Acta Mater. 1997, 45, 635–649. [Google Scholar] [CrossRef]
  38. Chen, W.; Song, B. Split Hopkinson (Kolsky) Bar; Springer: New York, NY, USA, 2011. [Google Scholar]
  39. Conque, H. The use of the direct impact Hopkinson pressure bar technique to describe thermally activated and viscous regimes of metallic materials. Philos. Trans. A Math. Phys. Eng. Sci. 2014, 372, 20130218. [Google Scholar]
  40. Meyers, M.A.; Nesterenko, V.F.; Lasalvis, J.C.; Xue, Q. Shear localization in dynamic deformation of materials: Microstructural evolution and self-organization. Mater. Sci. Eng. A 2001, 317, 204–225. [Google Scholar] [CrossRef]
  41. Ahn, D.H.; Lee, D.J.; Kang, M.J.; Park, L.J.; Lee, S.H.; Kim, H.S. Bi-modal structure of copper via room-temperature partial recrystallization after cryogenic dynamic compression. Metall. Mater. Trans. A 2016, 47, 1600–1606. [Google Scholar] [CrossRef]
  42. Bai, J.; Kou, H.C.; Wang, J.; Li, J.S.; Hu, R. Strain rate response of a Ti-based metallic glass composite at cryogenic temperature. Mater. Lett. 2014, 117, 228–230. [Google Scholar] [CrossRef]
  43. Jiang, K.; Ren, T.F.; Shan, G.B.; Ye, T.; Chen, L.Y.; Wang, C.X.; Zhao, F.; Li, J.G.; Suo, T. Dynamic mechanical responses of the Al0.1CoCrFeNi high entropy alloy at cryogenic temperature. Mater. Sci. Eng. A 2020, 797, 140125. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction of the as-cast AlCrFe2Ni2 alloy.
Figure 1. X-ray diffraction of the as-cast AlCrFe2Ni2 alloy.
Materials 18 01191 g001
Figure 2. Microstructure of the as-cast AlCrFe2Ni2 alloy: (a) SEM image; (b) the high-magnified SEM secondary electron image; (c) and (d) are magnified images of regions A and B in (b), respectively.
Figure 2. Microstructure of the as-cast AlCrFe2Ni2 alloy: (a) SEM image; (b) the high-magnified SEM secondary electron image; (c) and (d) are magnified images of regions A and B in (b), respectively.
Materials 18 01191 g002
Figure 3. (a) EBSD phase diagram of the as-cast AlCrFe2Ni2 alloy; (b) EBSD inverse pole figure (IPF).
Figure 3. (a) EBSD phase diagram of the as-cast AlCrFe2Ni2 alloy; (b) EBSD inverse pole figure (IPF).
Materials 18 01191 g003
Figure 4. (a,d) Bright-field images of the AlCrFe2Ni2; (b,e) selected area diffractions from zone axes [011] and [001], respectively; (c,f) dark-field images.
Figure 4. (a,d) Bright-field images of the AlCrFe2Ni2; (b,e) selected area diffractions from zone axes [011] and [001], respectively; (c,f) dark-field images.
Materials 18 01191 g004
Figure 5. TEM images of the as-cast AlCrFe2Ni2 alloy: (a) bright-field TEM image; (bf) the element distribution by TEM-EDS.
Figure 5. TEM images of the as-cast AlCrFe2Ni2 alloy: (a) bright-field TEM image; (bf) the element distribution by TEM-EDS.
Materials 18 01191 g005
Figure 6. (a) Static compression stress–strain curves of the AlCrFe2Ni2 high-entropy alloy under quasi-statics; (b) dynamic compression stress–strain curves at various strain rates.
Figure 6. (a) Static compression stress–strain curves of the AlCrFe2Ni2 high-entropy alloy under quasi-statics; (b) dynamic compression stress–strain curves at various strain rates.
Materials 18 01191 g006
Figure 7. The yield strength as a function of the logarithmic strain rate for the AlCrFe2Ni2 high-entropy alloy.
Figure 7. The yield strength as a function of the logarithmic strain rate for the AlCrFe2Ni2 high-entropy alloy.
Materials 18 01191 g007
Figure 8. Orientation image maps with the IPF and IQ + misorientation boundary maps of the samples after high-strain compression: (a,b) at 1300 s−1; (c,d) at 2100 s−1.
Figure 8. Orientation image maps with the IPF and IQ + misorientation boundary maps of the samples after high-strain compression: (a,b) at 1300 s−1; (c,d) at 2100 s−1.
Materials 18 01191 g008
Figure 9. The work-hardening rate (WHR): d σ /d ε ( σ is the flow stress and ε is the plastic strain).
Figure 9. The work-hardening rate (WHR): d σ /d ε ( σ is the flow stress and ε is the plastic strain).
Materials 18 01191 g009
Table 1. A summary of the compressive properties of the AlCrFe2Ni2 alloy under quasi-static deformation at room temperature.
Table 1. A summary of the compressive properties of the AlCrFe2Ni2 alloy under quasi-static deformation at room temperature.
Strain Rate [s−1]Yield Strength [MPa]Plastic Strain [%]
1 × 10−4606.1 ± 3.0>30.0
1 × 10−3616.3 ± 3.1>30.0
1 × 10−2691.6 ± 3.5>30.0
1 × 10−1655.8 ± 3.3>30.0
1 × 100658.3 ± 3.3>30.0
Table 2. A summary of the compressive properties of the AlCrFe2Ni2 alloy under dynamic deformation at room temperature.
Table 2. A summary of the compressive properties of the AlCrFe2Ni2 alloy under dynamic deformation at room temperature.
Strain Rate
[s−1]
Yield Strength [MPa]Compressive Stress [MPa]Plastic Strain
[%]
1300891.9 ± 4.51521.8 ± 7.612.8 ± 2.3
17001132.4 ± 5.71815.5 ± 9.114.7 ± 2.0
21001091.7 ± 5.51925.3 ± 9.618.2 ± 1.6
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

Yan, H.; Wang, Y.; Gan, X.; Dong, Y.; Liu, S.; Duan, S.; Mao, L. Microstructure Evolution and Mechanical Properties of Dual-Phase AlCrFe2Ni2 High-Entropy Alloy Under High-Strain-Rate Compression. Materials 2025, 18, 1191. https://doi.org/10.3390/ma18061191

AMA Style

Yan H, Wang Y, Gan X, Dong Y, Liu S, Duan S, Mao L. Microstructure Evolution and Mechanical Properties of Dual-Phase AlCrFe2Ni2 High-Entropy Alloy Under High-Strain-Rate Compression. Materials. 2025; 18(6):1191. https://doi.org/10.3390/ma18061191

Chicago/Turabian Style

Yan, Hang, Yu Wang, Xilin Gan, Yong Dong, Shichao Liu, Shougang Duan, and Lingbo Mao. 2025. "Microstructure Evolution and Mechanical Properties of Dual-Phase AlCrFe2Ni2 High-Entropy Alloy Under High-Strain-Rate Compression" Materials 18, no. 6: 1191. https://doi.org/10.3390/ma18061191

APA Style

Yan, H., Wang, Y., Gan, X., Dong, Y., Liu, S., Duan, S., & Mao, L. (2025). Microstructure Evolution and Mechanical Properties of Dual-Phase AlCrFe2Ni2 High-Entropy Alloy Under High-Strain-Rate Compression. Materials, 18(6), 1191. https://doi.org/10.3390/ma18061191

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