Next Article in Journal
Mechanism Design with Singularity Avoidance of Crystal-Inspired Deployable Structures
Next Article in Special Issue
An Investigation of Friction Coefficient on Microstructure and Texture Evolution of Interstitial-Free Steel during Warm Rolling and Subsequent Annealing
Previous Article in Journal
Field-Induced Magnetic Phase Transitions and Rich Phase Diagram of HoMnO3 Single Crystal
Previous Article in Special Issue
An Accurate Method for Crystallographic Reconstruction of Parent Austenite from Inherited Martensite in a Low-Alloy Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparing Cyclic Tension-Compression Effects on CoCrFeMnNi High-Entropy Alloy and Ni-Based Superalloy

1
Department of Materials Science and Engineering, National Chiao Tung University, Hsinchu 30010, Taiwan
2
Department of Materials Science and Engineering, National Tsing Hua University, Hsinchu 30013, Taiwan
3
High Entropy Materials Center, National Tsing Hua University, Hsinchu 30013, Taiwan
4
J-PARC Center, Japan Atomic Energy Agency, 2–4 Shirane Shirakata, Tokai-mura, Naka-gun, Ibaraki 319-1195, Japan
5
Department of Materials Science and Engineering, Chungnam National University, Daejeon 34134, Korea
6
Department of Materials Science and Engineering, Indian Institute of Technology, New Delhi 110016, India
*
Authors to whom correspondence should be addressed.
Submission received: 12 July 2019 / Revised: 1 August 2019 / Accepted: 8 August 2019 / Published: 13 August 2019

Abstract

:
An equal-molar CoCrFeMnNi, face-centered-cubic (fcc) high-entropy alloy (HEA) and a nickel-based superalloy are studied using in situ neutron diffraction experiments. With continuous measurements, the evolution of diffraction peaks is collected for microscopic lattice strain analyses. Cyclic hardening and softening are found in both metallic systems. However, as obtained from the diffraction-peak-width evolution, the underneath deformation mechanisms are quite different. The CoCrFeMnNi HEA exhibits distinct lattice strain and microstructure responses under tension-compression cyclic loadings.

1. Introduction

Fracture toughness and fatigue behavior are the key to the practical applications of high entropy alloys (HEAs) [1]. Hemphill et al.’s [2] four-point-bending fatigue experiments suggest that the fatigue behavior of Al0.5CoCrCuFeNi HEA is affected by microstructural defects of oxides and microcracks while Tang et al.’s [3] high-cycle four-point-bending fatigue experiments reveal that nanotwinning is the major deformation behavior before fracture in a Al0.5CoCrCuFeNi two-phase HEA. Despite the convenience of four-point-bending high-cycle fatigue (HCF), more detailed fatigue mechanisms, especially in the low-cycle fatigue (LCF) in HEAs, need to be further explored, since the deformation behaviors under HCF and LCF regimes are very different, as suggested by Chen et al.’s review [4].
Thurston et al. [5] report that fatigue behavior of the Cantor alloy is comparable to the austenitic stainless steels and twinning induced plasticity (TWIP) steels. Recently, Niendorf et al.’s [6] tension-compression LCF experiments indicate the influence of planarity of slip and partial reversibility of deformation on the unobvious cyclic hardening at given plastic strain in Fe50Mn30Co10Cr10 HEA. There are few reports revealing the polymorphism of CoCrFeMnNi, which is induced by tension [7] and compression [8,9,10], respectively. Moreover, the cyclic tension-compression effects as well as predominant mechanisms during the LCF process in CoCrFeMnNi still remain elusive. Meanwhile, the cyclic-loading deformation in face-centered-cubic (fcc) nickel-based superalloy has been extensively examined in which lattice strain asymmetry associated with dislocation density evolution governs the irreversible plastic deformation [11,12,13,14]. We herein investigate the in-situ tension-compression cyclic loading effects and the dominant deformation behaviors of the CoCrFeMnNi HEA in comparison with the commercially popular nickel-based superalloy by exploiting in-situ neutron diffraction measurements.

2. Materials and Methods

The nickel-based superalloy HASTELLOY C22™ (Ni-21Cr-17Mo in weight percentage) was annealed to dissolve most precipitates into the fcc matrix structure. C22™ has been chosen as a reference because this Ni-based superalloy has been extensively studied using neutron diffraction and other complimentary experiments [11,12,13,14,15,16,17]. Several correlations from diffraction-data characteristics, microscopic images and underneath deformation mechanisms have been found as benchmarks and references. Hence, we follow the same protocol and experimental design performed on the C22™ to apply to the current CoCrFeMnNi HEA examinations. In addition, the LCF experiments and cyclic-loading deformation mechanisms of the C22™ are clarified and cited in many articles [11,12,13,14,15,16].
For the CoCrFeMnNi HEA, the sample was fabricated by vacuum arc-melting (arc) with equal molar compositions of constituent powders whose purity is more than 99.9% (in weight percent) and the diffraction characteristics of the CoCrFeMnNi HEA are also well studied for its microstructure [10,18,19,20].
The selected gauge volume for both samples was approximately 120 mm3 to associate the microstructure with the bulk properties. During cyclic deformation of the C22™ alloy, in situ neutron diffraction measurements with the Spallation Neutron Source were carried out at the Spectrometer for Materials Research at Temperature and Stress (SMARTS) beamline of Los Alamos Neutron Science Center (LANSCE). The in situ neutron diffraction measurements for LCF of the HEA were conducted at TAKUMI beamline in Materials and Life Science Experimental Facilities at J-PARC, Japan. The loading direction is 45° relative to the incident neutron beam direction, with two detector banks located at ±90° concomitantly recording diffraction vector patterns in the loading and transverse directions. Both the in situ neutron diffraction LCF tests were conducted at room temperature with the maximum tensile strain of 1% and the maximum compressive strain of −1%. The cyclic frequency was 1 Hz. The neutron diffraction profiles were analyzed by applying pseudo-Voigt function for single peak fitting with the General Structure Analysis Software (GSAS, Los Alamos National Laboratory, version 1, Los Alamos County, New Mexico, United States) [21] and Convolutional Multiple Whole Profile (CMWP) [22] software, following the previously used protocol [13,15].

3. Results

3.1. Chemical Composition of Alloying Elements in CoCrFeMnNi

We employed JEOL 6500 scanning electron microscopy (SEM, Japan Electron Optics Laboratory, Tokyo, Japan) with energy-dispersive x-ray (EDX, Japan Electron Optics Laboratory, Tokyo, Japan) spectroscopy for determining the elemental composition. The EDX analysis in Figure 1 discloses the alloy composition of CoCrFeMnNi with an equal weight percentage distribution of each constituent element, determining the Co20Cr20Fe20Mn20Ni20 HEA in the present study.

3.2. Low-Cycle Fatigue Experiments

Figure 2 presents the macroscopic stresses evolution versus certain selected fatigue cycles measured at the maximum strain of 1% and the minimum strain of −1% for C22™ and CoCrFeMnNi alloys. Both the C22™ and HEA performed similar cyclic hardening followed by softening tendency, however, the microscopic mechanisms underlying their cyclic hardening/softening-structural transformation were in different behaviors. Three distinct regions of the fatigue cycles based on the applied stress development can be clearly observed. During continuous tension-compression LCF deformation, the C22™ alloy revealed the cyclic hardening within the 100th cycle region with increasing applied stress (stage I), followed by the cyclic softening upon the 1000th cycle with a gradual decrease in applied stress (stage II) and remaining saturated (stage III) until the 2500th cycle. Compared to the C22™, the HEA experienced the hardening/softening transition at the 10th fatigue cycle, and then cyclic softening. At the same strain of tensile (1%) or compressive (− 1%) deformation, an increase in stress up to the hardening/softening transition on the C22™ was more pronounced than that on the HEA, which demonstrated the greater hardening behavior of the C22™ alloy.

3.3. Lattice Strain Evolution during Tension-Compression Cyclic Loadings

The lattice strain evolution was calculated based on the variations in the d-spacing of the refined neutron diffraction peak positions of lattice planes as a function of fatigue cycle with respect to the initial d-spacing before deformation. Figure 3 depicts the lattice strain-stress dependence of the {311} plane at the maximum tensile (1%) and compressive (−1%) strains in both transverse and axial loading directions since {311} grain families are the least affected crystallographic planes by intrinsic microstructure deformation rather than external loadings in fcc polycrystalline materials under loading [12,23]. The yellow arrows show the trends of lattice strain evolution under cyclic loadings. Within the cyclic hardening stage in the early fatigue cycles, we can observe a positive correlation between the elastic strain and applied stress in both C22™ and HEA, which seems to follow Hooke’s law as Suzuki and Egami reported for cyclic deformation [24]. Specifically, C22™ has a nearly linear relationship up to the hardening/softening transition. However, although the lattice strain-stress relationship of HEA is also positive up to the 10th fatigue cycle, the trend is not linear. Several reports suggest that such a nonlinear trend may be governed by the texture, patterned dislocation structure, or other anisotropy of the materials [11].
In the hardening cycles of stage I, both the C22™ and HEA displayed the same propensity in which the lattice strain increases with the number of fatigue cycles under both tensile and compressive deformation in both axial loading and transverse directions, ascribed to the cyclic hardening behavior in this region. However, the increase of elastic strain in the C22™ was much greater than that of the HEA under maximum tension shown in Figure 3a,b. Meanwhile, the lattice strain evolution of the C22™ and HEA was almost the same under maximum compressive deformation within the cycling hardening region. In stage II, there was an opposite trend of lattice strain evolution between the two alloys. The lattice strain of HEA continued to increase while that of C22™ experienced an initially slight decrease within the cyclic softening region. Furthermore, the elastic strain of C22™ increases under tension in Figure 3a but decreases under compression in Figure 3c from the 250th to the 1000th fatigue cycle for the grains oriented along the transverse direction. A similar tendency was also obtained for the grains of C22™ oriented along the axial loading direction in tension in Figure 3b and compression in Figure 3d. The deviation of elastic strain within the cyclic softening region was ascribed to the development of the intergranular strains [16]. In stage III, there was a gradually decrease of lattice strain of C22™ under both tensile and compressive tests for grains oriented along both axial loading and transverse directions after the 1000th cycle. Similar to the C22™ alloy, the lattice strain of HEA also depicted a slight decrease after the 200th cycle. The major discrepancy in the lattice strain evolution of the C22™ and HEA was further clarified in terms of the dominant development of dislocation slip activities and other fatigue-induced defects such as stacking faults and twins.

4. Discussion

In this work, we determine the evolution of average dislocation density (n) and dislocation-wall spacings (D) by using CMWP software [22] in fitting the neutron diffraction profiles. Figure 4a depicts the dislocation density and dislocation wall spacings as a function of fatigue cycle at the maximum strain of 1% in both C22™ [13] and CoCrFeMnNi alloys along the axial loading direction. For CoCrFeMnNi, the dislocation density sharply increased together with an increasing number of fatigue cycles in the very early stages in the axial loading direction and reached the maximum value at the 40th cycle, then decreased with further fatigue cycles undergoing the softening cycles. The decrease in the density of the randomly distributed dislocations after the hardening cycles seen in the CoCrFeMnNi was analogous to that obtained in the C22™ alloy [13]. The average distances between the dislocation-wall structures are largest at a very early cycle because the annealed C22™ alloy has fewer cell structures than Mughrabi’s dislocation wall [25] in the beginning [17]. The C22™ alloy mainly has randomly distributed dislocations before cyclic loading [17]. Upon cyclic loading, the average distances between the dislocation-wall structures dropped within the hardening region, and remained constant with higher fatigue cycles undergoing the softening and saturation regimes. Due to dislocation movement and interaction, a more orderly arrangement of dislocations associated with the formation of the Mughrabi’s dislocation-wall for cell structures resulted in a decrease and final saturation of the dislocation-wall distance when the dislocation densities approached a more stable state in CoCrFeMnNi. The dislocation hardening was demonstrated to be the main deformation mechanism of the cyclic hardening effect in the C22™ alloy. The calculated dislocation-wall-spacing evolution within the hardening region of HEA is similar to that of C22™ but the dislocation density evolution of HEA is quite different from that of C22™, implying that there are other microstructure formations coupled with cyclic deformation activities in governing the LCF regime. It has been reported that mechanical nanotwinning is the major deformation mechanism before fatigue crack initiation and propagation in HEA [3,4,26,27].
The generation of different kinds of defects associated with the dislocation density evolution under LCF was further elucidated. As reported by Ungar, each kind of defect is simultaneously governed by distinct diffraction peak aberrations such as peak shift, peak broadening, peak asymmetry, anisotropic peak broadening, and peak shape [28]. We first examined the deformation fault probability of CoCrFeMnNi obtained directly from the changes in only diffraction peak separations, which is applied only for fcc structure, as summarized by Warren [29]. Using this method, we can predict the presence of deformation fault probability based on only peak-shift correlations between (1 1 1) and (2 0 0) of polycrystalline metals in fcc structure, specifically, the displacement of d111-d200 separation toward smaller or larger deviation between these two groups of lattice planes, as shown in Figure 4b. However, we cannot distinguish whether the deformation fault is intrinsic or extrinsic stacking faults. To attain the quantitative probability of intrinsic, extrinsic stacking faults, and twinning, a more complete and reliable analysis of CMWP program [22] was further employed to resolve the convoluted diffraction data and the corresponding results are presented in Figure 4c.
It can be seen from CMWP fitting results in Figure 4c that the formation probability of twinning, extrinsic and intrinsic stacking faults defects started to slightly increase with the dislocation density evolution and the extrinsic stacking fault plays a major role in deformation in the hardening cycles. More twinning and extrinsic stacking faults were generated easily, and they governed the cyclic hardening behavior. Further increase of dislocation densities led to the decrease of twinning, and the probability of extrinsic and intrinsic stacking faults in the first half of softening cycles. When the dislocation densities were in the more stable condition, there was a slight variation of extrinsic and intrinsic stacking faults but an obvious increasing trend of twinning probability, suggesting the dominant role of twinning in the latter half of softening cycles. The formation of nanotwins and stacking faults is attributed to the low stacking fault energy (SFE) of HEA.

5. Conclusions

During low-cycle fatigue experiments, both the CoCrFeMnNi and C22™ alloys display obvious cyclic hardening and softening regimes, however, the CoCrFeMnNi indicated fewer hardening cycles comparing to the C22™ alloy. The lattice strain-stress relationship of these two alloys are quite different, indicating that dominant deformation-induced microstructures were distinct under tension-compression cyclic loadings. The twinning, extrinsic and intrinsic stacking faults are found to be formed during cyclic hardening regimes, and twinning is the most predominant defect in governing the main deformation behavior of CoCrFeMnNi in the latter half of softening regimes. The role of twin formation subjected to tension-compression cyclic loadings on CoCrFeMnNi is revealed, which may be useful for HEA design for fatigue endurance.

Author Contributions

Conceptualization, E.-W.H.; data analysis, T.-N.L., Y.-S.C., B.-H.L. and E.-W.H.; investigation, Y.-J.C. and T.-R.S.; project administration, E.-W.H.; resources, A.-C.Y. and S.H.; supervision, E.-W.H.; writing—original draft, T.-N.L. and E.-W.H.; writing—review and editing, A.C.Y., S.Y.L. and J.J.

Funding

The neutron diffraction experiments were performed at BL19 in Materials and Life Science Experimental Facility of J-PARC with the proposal 2016A0322. The authors are grateful to the support of the Ministry of Science and Technology (MOST) Programs, 107-2218-E-007-012, 107-3017-F-009-002, 107-3017-F-007-003, 107-3017-F-009-002, 107-3011-F-002-002, and 108-2221-E-009-131-MY4. This work was financially supported by the “Center for the Semiconductor Technology Research” from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE) in Taiwan. Additionally supported in part by the MOST, Taiwan, under Grant MOST-108-3017-F-009-003. EWH and TNL are supported by MOST 107-2628-E-009-001-MY3. SYL was supported by a National Research Foundation (NRF) grant funded by the Korean government (No.2017K1A3A7A09016308).

Acknowledgments

We sincerely appreciate Tamás Ungár for his Convolutional Multiple Whole Profile (CMWP) software. EWH very much appreciate the financial support from the National Synchrotron Radiation Research Center (NSRRC) Neutron Program. This work has benefited from the use of Spectrometer for Materials Research at Temperature and Stress (SMARTS) of the Lujan Neutron Scattering Center at LANSCE, which is funded by the Department of Energy’s Office of Basic Energy Sciences. Los Alamos National Laboratory is operated by Los Alamos National Security LLC under DOE.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Seifi, M.; Li, D.; Yong, Z.; Liaw, P.K.; Lewandowski, J.J. Fracture Toughness and Fatigue Crack Growth Behavior of As-Cast High-Entropy Alloys. Jom 2015, 67, 2288–2295. [Google Scholar] [CrossRef]
  2. 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 Materialia 2012, 60, 5723–5734. [Google Scholar] [CrossRef]
  3. Tang, Z.; Yuan, T.; Tsai, C.W.; Yeh, J.W.; Lundin, C.D.; Liaw, P.K. Fatigue behavior of a wrought Al0.5CoCrCuFeNi two-phase high-entropy alloy. Acta Materialia 2015, 99, 247–258. [Google Scholar] [CrossRef]
  4. Chen, P.; Lee, C.; Wang, S.Y.; Seifi, M.; Lewandowski, J.J.; Dahmen, K.A.; Jia, H.; Xie, X.; Chen, B.; Yeh, J.W.; et al. Fatigue behavior of high-entropy alloys: A review. Sci. China Technol. Sci. 2017, 61, 168–178. [Google Scholar] [CrossRef]
  5. Thurston, K.V.S.; Gludovatz, B.; Hohenwarter, A.; Laplanche, G.; George, E.P.; Ritchie, R.O. Effect of temperature on the fatigue-crack growth behavior of the high-entropy alloy CrMnFeCoNi. Intermetallics 2017, 88, 65–72. [Google Scholar] [CrossRef]
  6. Niendorf, T.; Wegener, T.; Li, Z.; Raabe, D. Unexpected cyclic stress-strain response of dual-phase high-entropy alloys induced by partial reversibility of deformation. Scr. Mater. 2018, 143, 63–67. [Google Scholar] [CrossRef]
  7. Niu, C.; LaRosa, C.R.; Miao, J.; Mills, M.J.; Ghazisaeidi, M. Magnetically-driven phase transformation strengthening in high entropy alloys. Nat. Commun. 2018, 9, 1363. [Google Scholar] [CrossRef]
  8. Tracy, C.L.; Park, S.; Rittman, D.R.; Zinkle, S.J.; Bei, H.; Lang, M.; Ewing, R.C.; Mao, W.L. High pressure synthesis of a hexagonal close-packed phase of the high-entropy alloy CrMnFeCoNi. Nat. Commun. 2017, 8, 15634. [Google Scholar] [CrossRef] [Green Version]
  9. Zhang, F.; Wu, Y.; Lou, H.; Zeng, Z.; Prakapenka, V.B.; Greenberg, E.; Ren, Y.; Yan, J.; Okasinski, J.S.; Liu, X.; et al. Polymorphism in a high-entropy alloy. Nat. Commun. 2017, 8, 15687. [Google Scholar] [CrossRef] [Green Version]
  10. Huang, E.W.; Lin, C.M.; Jain, J.; Shieh, S.R.; Wang, C.P.; Chuang, Y.C.; Liao, Y.F.; Zhang, D.Z.; Huang, T.; Lam, T.-N.; et al. Irreversible phase transformation in a CoCrFeMnNi high entropy alloy under hydrostatic compression. Mater. Today Commun. 2018, 14, 10–14. [Google Scholar] [CrossRef]
  11. Huang, E.W.; Barabash, R.I.; Clausen, B.; Liaw, P.K. Cyclic-Loading Induced Lattice-Strain Asymmetry in Loading and Transverse Directions. Metall. Mater. Trans. A 2012, 43, 1454–1461. [Google Scholar] [CrossRef]
  12. Huang, E.W.; Barabash, R.I.; Wang, Y.; Clausen, B.; Li, L.; Liaw, P.K.; Ice, G.E.; Ren, Y.; Choo, H.; Pike, L.M.; et al. Plastic behavior of a nickel-based alloy under monotonic-tension and low-cycle-fatigue loading. Int. J. Plast. 2008, 24, 1440–1456. [Google Scholar] [CrossRef]
  13. Huang, E.W.; Barabash, R.I.; Clausen, B.; Liu, Y.L.; Kai, J.J.; Ice, G.E.; Woods, K.P.; Liaw, P.K. Fatigue-induced reversible/irreversible structural-transformations in a Ni-based superalloy. Int. J. Plast. 2010, 26, 1124–1137. [Google Scholar] [CrossRef]
  14. Huang, E.; Clausen, B.; Wang, Y.; Choo, H.; Liaw, P.; Benson, M.; Pike, L.; Klarstrom, D. A neutron-diffraction study of the low-cycle fatigue behavior of HASTELLOY® C-22HSTM alloy. Int. J. Fatigue 2007, 29, 1812–1819. [Google Scholar] [CrossRef]
  15. Huang, E.W.; Chang, C.K.; Liaw, P.K.; Suei, T.R. Fatigue induced deformation and thermodynamics evolution in a nano particle strengthened nickel base superalloy. Fatigue Fract Eng Mater Struct. 2016, 39, 675–685. [Google Scholar] [CrossRef]
  16. Huang, E.W.; Barabash, R.I.; Ice, G.E.; Liu, W.; Liu, Y.L.; Kai, J.J.; Liaw, P.K. Cyclic-loading-induced accumulation of geometrically necessary dislocations near grain boundaries in an Ni-based superalloy. JOM: J. Miner., Met. Mater. Sci. 2009, 61, 53–58. [Google Scholar] [CrossRef]
  17. Huang, E.W.; Barabash, R.; Jia, N.; Wang, Y.D.; Ice, G.E.; Clausen, B.; Horton, J.; Liaw, P.K. Slip-System-Related Dislocation Study from In-Situ Neutron Measurements. Metall. Mater. Trans. A 2008, 39, 3079. [Google Scholar] [CrossRef]
  18. Huang, E.W.; Lin, C.M.; Juang, J.Y.; Chang, Y.J.; Chang, Y.W.; Wu, C.S.; Tsai, C.W.; Yeh, A.C.; Shieh, S.R.; Wang, C.P.; et al. Deviatoric deformation kinetics in high entropy alloy under hydrostatic compression. J. Alloys. Compd. 2019, 792, 116–121. [Google Scholar] [CrossRef]
  19. Huang, E.W.; Yu, D.; Yeh, J.W.; Lee, C.; An, K.; Tu, S.Y. A study of lattice elasticity from low entropy metals to medium and high entropy alloys. Scr. Mater. 2015, 101, 32–35. [Google Scholar] [CrossRef] [Green Version]
  20. Woo, W.; Huang, E.W.; Yeh, J.W.; Choo, H.; Lee, C.; Tu, S.Y. In-situ neutron diffraction studies on high-temperature deformation behavior in a CoCrFeMnNi high entropy alloy. Intermetallics 2015, 62, 1–6. [Google Scholar] [CrossRef]
  21. Von Dreele, R.B.; Larson, A.C. General Structure Analysis System (GSAS). Los Alamos Natl. Lab. Rep. LAUR 2004, 86–748. [Google Scholar]
  22. Ribárik, G.; Ungár, T.; Ungár, J. MWP-fit a program for multiple whole-profile fitting of diffraction peak profiles by ab initio theoretical functions. J. Appl. Crystallogr. 2001, 34, 669–676. [Google Scholar] [CrossRef]
  23. Velterop, L.; Delhez, R.; de Keijser, Th.H.; Mittemeijera, E.J.; Reefman, D. X-ray diffraction analysis of stacking and twin faults in f.c.c. metals. J. Appl. Cryst. 2000, 33, 296–306. [Google Scholar] [CrossRef]
  24. Suzuki, Y.; Egami, T. Shear deformation of glassy metals: Breakdown of cauchy relationship and anelasticity. J. Non. Cryst. Solids 1985, 75, 361–366. [Google Scholar] [CrossRef]
  25. Mughrabi, H. Dislocation clustering and long-range internal stresses in monotonically and cyclically deformed metal crystals. Rev. Phys. Appl. (Paris) 1988, 23, 367–379. [Google Scholar] [CrossRef]
  26. Laplanche, G.; Kostka, A.; Horst, O.M.; Eggeler, G.; George, E.P. Microstructure evolution and critical stress for twinning in the CrMnFeCoNi high-entropy alloy. Acta Materialia 2016, 118, 152–163. [Google Scholar] [CrossRef] [Green Version]
  27. Yu, P.F.; Cheng, H.; Zhang, L.J.; Zhang, H.; Ma, M.Z.; Li, G.; Liaw, P.K.; Liu, R.P. Nanotwin’s formation and growth in an AlCoCuFeNi high-entropy alloy. Scr. Mater. 2016, 114, 31–34. [Google Scholar] [CrossRef]
  28. Ungár, T. Microstructural parameters from X-ray diffraction peak broadening. Scr. Mater. 2004, 51, 777–781. [Google Scholar] [CrossRef]
  29. Warren, B.E. X-ray diffraction. Phys. Today 1970, 23, 53. [Google Scholar] [CrossRef]
Figure 1. EDX spectrum of the CoCrFeMnNi HEA.
Figure 1. EDX spectrum of the CoCrFeMnNi HEA.
Crystals 09 00420 g001
Figure 2. The stress evolution at the maximum tension (1%) and compression (−1%) as a function of fatigue cycle (presented in log scale) for both the C22™ and CoCrFeMnNi alloys.
Figure 2. The stress evolution at the maximum tension (1%) and compression (−1%) as a function of fatigue cycle (presented in log scale) for both the C22™ and CoCrFeMnNi alloys.
Crystals 09 00420 g002
Figure 3. The {311} lattice strain-stress relationship at the maximum tensile strain of 1% in (a) transverse and (b) axial loading; and at the maximum compressive strain of −1% in (c) transverse and (d) axial loading directions for both the C22™ and CoCrFeMnNi alloys.
Figure 3. The {311} lattice strain-stress relationship at the maximum tensile strain of 1% in (a) transverse and (b) axial loading; and at the maximum compressive strain of −1% in (c) transverse and (d) axial loading directions for both the C22™ and CoCrFeMnNi alloys.
Crystals 09 00420 g003
Figure 4. (a) The dislocation density and dislocation-wall-spacing evolution in C22™ and CoCrFeMnNi alloys. The magnification of n and D of C22™ alloy is indicated in the inset. (b) The difference in d-spacing between diffraction peaks 111 and 200 of CoCrFeMnNi. (c) Twining, extrinsic and intrinsic stacking faults defects as a function of fatigue cycle (presented in log scale) in CoCrFeMnNi.
Figure 4. (a) The dislocation density and dislocation-wall-spacing evolution in C22™ and CoCrFeMnNi alloys. The magnification of n and D of C22™ alloy is indicated in the inset. (b) The difference in d-spacing between diffraction peaks 111 and 200 of CoCrFeMnNi. (c) Twining, extrinsic and intrinsic stacking faults defects as a function of fatigue cycle (presented in log scale) in CoCrFeMnNi.
Crystals 09 00420 g004

Share and Cite

MDPI and ACS Style

Lam, T.-N.; Chou, Y.-S.; Chang, Y.-J.; Sui, T.-R.; Yeh, A.-C.; Harjo, S.; Lee, S.Y.; Jain, J.; Lai, B.-H.; Huang, E.-W. Comparing Cyclic Tension-Compression Effects on CoCrFeMnNi High-Entropy Alloy and Ni-Based Superalloy. Crystals 2019, 9, 420. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9080420

AMA Style

Lam T-N, Chou Y-S, Chang Y-J, Sui T-R, Yeh A-C, Harjo S, Lee SY, Jain J, Lai B-H, Huang E-W. Comparing Cyclic Tension-Compression Effects on CoCrFeMnNi High-Entropy Alloy and Ni-Based Superalloy. Crystals. 2019; 9(8):420. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9080420

Chicago/Turabian Style

Lam, Tu-Ngoc, You-Shiun Chou, Yao-Jen Chang, Tsung-Ruei Sui, An-Chou Yeh, Stefanus Harjo, Soo Yeol Lee, Jayant Jain, Bo-Hong Lai, and E-Wen Huang. 2019. "Comparing Cyclic Tension-Compression Effects on CoCrFeMnNi High-Entropy Alloy and Ni-Based Superalloy" Crystals 9, no. 8: 420. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst9080420

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