Micro-mechanisms of microstructural damage due to low cycle fatigue in CoCuFeMnNi high entropy alloy

Micro-mechanisms of microstructural damage due to low cycle fatigue in CoCuFeMnNi high entropy alloy

International Journal of Fatigue 130 (2020) 105258 Contents lists available at ScienceDirect International Journal of Fatigue journal homepage: www...

16MB Sizes 0 Downloads 26 Views

International Journal of Fatigue 130 (2020) 105258

Contents lists available at ScienceDirect

International Journal of Fatigue journal homepage: www.elsevier.com/locate/ijfatigue

Micro-mechanisms of microstructural damage due to low cycle fatigue in CoCuFeMnNi high entropy alloy Fateh Bahadur, Krishanu Biswas, N.P. Gurao

T



Department of Materials Science and Engineering, IIT Kanpur, Kanpur 208016, India

ARTICLE INFO

ABSTRACT

Keywords: Advanced materials Low cycle fatigue Microstructures Dislocations Damage

Low cycle fatigue behavior of equiatomic CoCuFeMnNi high entropy alloy was investigated under fully reversible strain control mode followed with detailed microstructural characterization using electron back scatter diffraction. There is an increase in the intragranular misorientation and geometrically necessary dislocation density with increase in the strain amplitude that leads to accumulation of damage and intergranular cracking. The interaction of dislocations with copper rich nano-clusters, solute environment and grain boundaries affect slip reversibility, thereby deciding cyclic deformation behaviour. Annealing twin boundaries are resistant to damage and increasing their population density by grain boundary engineering can improve performance of CoCuFeMnNi alloy.

1. Introduction Design and development of new alloys for achieving excellent mechanical properties like high specific strength, optimum combination of strength and ductility, excellent fracture toughness at low temperature, superior fatigue and creep resistance has been pursued to cater the stringent demands of the engineering industries in the recent past. Recently, a new class of materials known as high entropy alloys (HEAs) or complex concentrated alloys (CCAs) that are multi-principal multicomponent alloys containing five or more elements in equal or near equal proportions have been developed [1,2]. Their high configurational entropy of mixing favors the formation of solid solution over brittle intermetallic phases at high temperature [1–4]. This innovative alloy design strategy overcomes the conventional approach in which properties are primarily governed by principal element, thus, providing enormous opportunities in terms of physical and functional properties by permitting extensive and flexible tuning of alloy composition. The past decade of dedicated research revealed that many HEAs possess unique properties in comparison with the conventional alloys for instance, high thermal and microstructural stability [5,6], high hardness [6], high strength at elevated temperature [7,8], superior fracture toughness [9], and excellent wear resistance [10]. Among HEAs, equiatomic CoCrFeMnNi Face Centered Cubic (FCC) HEA (Cantor alloy) exhibits high tensile strength (1130 MPa), sufficient ductility (65%), and higher fracture toughness (200 MPa.m1/2 ); especially at cryogenic temperatures due to the presence of deformation twins and has been studied in great details [4,9,11–13].



Most of these studies regarding mechanical behavior are usually investigated in uniaxial loading, [14–19] which are necessary but still insufficient to broaden the application range of HEAs for structural applications. Most engineering applications involve load reversal and cyclic loading in service, and therefore, it is essential to study them under cyclic loading. In particular, low cycle fatigue behavior of metals and alloys is important from technological perspective as well as from a scientific point of view to investigate the accumulation of damage wherein dislocations and their interaction with each other and other defects play an important role. Low cycle fatigue involves plastic strain and is generally investigated in strain control mode. Deformation behaviour and failure of different engineering materials like IF steel [20], dual phase steel, low alloy steel, TWIP/TRIP steel [21–24], stainless steel [25–27], commercially pure titanium, titanium alloy [28,29], and Ni-based superalloys [30,31] has been well-established under low cyclic fatigue regime. However, the research on fatigue behavior, particularly pertaining to low cycle fatigue of high entropy alloys, has been limited. Hemphill et al. [32] have examined the high cycle fatigue behavior of as-cast and wrought Al0.5CoCrCuFeNi alloy using four-point bending test at room temperature and found that the fatigue endurance limit lies in the range of 540–945 MPa. A similar composition of Al0.5CoCrCuFeNi alloy with high purity elements has also been studied by Tang et al. [33] who showed that the use of high purity elements results in better fatigue performance. This was attributed to the formation of nano-twins during cyclic loading in the alloy containing high purity elements leading to higher endurance limit and prolonged fatigue life compared to many

Corresponding author. E-mail address: [email protected] (N.P. Gurao).

https://doi.org/10.1016/j.ijfatigue.2019.105258 Received 26 June 2019; Received in revised form 12 August 2019; Accepted 28 August 2019 Available online 29 August 2019 0142-1123/ © 2019 Elsevier Ltd. All rights reserved.

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al.

Nomenclature

R APT IPF GROD KAM GND

List of symbols and acronyms Δε/2 c n

strain amplitude fatigue ductility exponent strain hardening co-efficient

conventional alloys. Niendorf et al. [34] reported the low cycle fatigue behavior of Fe50Mn30Co10Cr10 alloy showing remarkable changes in the deformation mechanisms under monotonic and cyclic loading conditions. The alloy showed poor strain hardening in tension but occurrence of strain-induced martensitic transformation in low cycle fatigue regime contributed to better fatigue performance. This peculiar behavior was attributed to partial deformation reversibility and the presence of the planar nature of slip. In other investigations, microstructural engineering approaches like grain refinement were employed to improve fatigue performance [35,36]. Multi-phase high entropy alloys investigated till date show remarkable fatigue strength compared to the conventional alloy systems. But the fundamental understanding of the fatigue deformation mechanism of single phase high entropy alloys, especially under low cycle fatigue regime has not been studied yet. Accordingly, the present study focuses on the low cycle fatigue behavior of single phase FCC CoCuFeMnNi HEA system developed by Tazuddin et al. [37], under strain control conditions. It was shown that operation of planar partial slip along with conventional octahedral slip led to the unique Goss-Brass type texture evolution during cold rolling of the alloy [38]. An atom probe tomography study by Sonkusare et al. [39] revealed the presence of copper rich nano-clusters (2.5 nm average

strain ratio atom probe tomography inverse pole figure grain reference orientation deviation kernel average misorentation geometrically necessary dislocation

size) in the FCC matrix of the alloy. Other investigations on the alloy have established the importance of the copper rich nano-clusters and their interactions with dislocations in tension and high pressure torsion in monotonic and cyclic mode [40,41]. Thus, the deformation behavior of the CoCuFeMnNi high entropy alloy is studied for different loading conditions. However, the cyclic deformation of the alloy is still unexplored. The present investigation deals with understanding the deformation mechanism and microstructural evolution of CoCuFeMnNi alloy subjected to strain controlled low cycle fatigue using various characterization tools like scanning electron microscopy (SEM), electron back scatter diffraction (EBSD) and back scattered electron imaging showing electron channeling contrast. 2. Experimental procedure 2.1. Material and methods Equiatomic CoCuFeMnNi alloy was prepared by arc melting buttons of ~12 g of equimolar mixture of elements with greater than 99.9% purity obtained from Sigma Aldrich. Alloy buttons were remelted at least four times to obtain compositional homogeneity. Four such alloy

Fig. 1. (a) X-ray diffraction pattern, (b) normal direction inverse pole figure map and (c) elemental distribution map from atom probe tomography of the homogenized CoCuFeMnNi HEA sample. 2

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al.

buttons were remelted again and suction cast into a billet of dimension 10 × 10 × 60 mm3. The billet was sealed in a quartz tube with argon and homogenized at 1273 K for 24 h followed by water quenching. Subsequently, the homogenized billet was cold rolled to 50% reduction in thickness at room temperature. The rolled specimen was subjected to annealing for 2 h at 1273 K to obtain a fully crystallized microstructure. Tensile and fatigue samples were machined from the annealed billet by electric discharge machining according to the ASTM E8 (10 mm × 2.5 mm × 2.5 mm) and ASTM E-606 (8 mm × 5 mm × 2.5 mm) standards respectively.

diffraction (EBSD, Oxford Instruments, U.K.) detector. EBSD scans were acquired with Aztec HKL software. The post-processing of the data was performed using TSL-OIM version 7.2 software. The EBSD samples were prepared by polishing to 2000 grit size using emery paper and then cloth polishing using the 0.5 μm alumina suspension followed with vibromet polishing (Buehler, USA) using colloidal silica suspension with a particle size of 50 nm. Backscattered electron imaging was carried out to obtain electron channeling contrast [44] by tilting the sample 10 degree at a working distance of 10 mm and rotating the sample to reveal the dislocation structure near the fracture tip of the fatigue tested samples.

2.2. Tensile and fatigue test

3. Results

The uniaxial tensile test was carried out at room temperature with a strain rate of 1 × 10 3s 1 using Instron-1195 Universal Testing Machine. Low cycle fatigue tests were also performed under straincontrol conditions at different strain amplitude (Δε/2) of 0.5%, 0.75%, 1% and 1.5% with fully reversible cyclic strain condition (R = −1) up to failure. All fatigue tests were conducted with the strain rate of 5 × 10−3 s−1 using the 100 kN Universal Testing Machine (BiSS, Bangalore, India).

3.1. Initial material X-ray diffraction (XRD) pattern of the homogenized CoCuFeMnNi alloy shown in Fig. 1(a) indicates the presence of FCC single phase with a lattice parameter of 0.361 nm. The normal direction (ND) inverse pole figure (IPF) map of the investigated HEA in Fig. 1(b) shows a fully recrystallized microstructure with annealing twins and an area average grain size of ~47 µm. Atom probe tomography carried out on the homogenized sample provided elemental distribution at the atomic length scale. Fig. 1(c) shows the elemental distribution from atom probe tomography indicating the homogeneous distribution of cobalt, iron, manganese, and nickel while there is heterogeneous distribution of copper, characterized by the presence of near spherical copper rich nano-clusters with an average diameter of 2.5 nm.

2.3. X-ray diffraction and microstructure analyses The crystal structure of the homogenized HEA was examined by Xray diffraction (PANalytical, Netherland), with a step size of 0.02 degree/ s, operated at 45 kV and 40 mA using CuKα (λ = 0.15402 nm) radiation. The grain morphology, fractography and compositional analyses of the synthesized sample were carried out utilizing field emission scanning electron microscope (FE-SEM) (JSM-7100F, JEOL, Japan) equipped with energy dispersive spectroscopy facility (EDS, Oxford Instruments, UK) operated at 20 kV. Atom probe tomography (APT) was carried out using the local electrode atom probe (LEAP 4000X HR™, Cameca Instruments) to obtain near-atomic scale elemental distribution. APT samples were prepared by using standard procedure [42,43] using a dual-beam focused ion beam (FEI Helios Nanolab 660, USA). A volume of 60 × 60 × 100 nm3 was acquired, through a pulse repetition rate of 250 kHz and pulse energy of 30 pJ. The tip temperature was maintained at around 60 K. The APT data was evaluated with the help of IVAS 3.6.10a software provided by Cameca Instruments.

3.2. Tensile and fatigue tests Engineering stress - engineering strain behaviour of the CoCuFeMnNi alloy as shown in Fig. 2(a) indicates that the investigated alloy exhibits yield strength of 324 ± 20 MPa and ultimate tensile strength of 710 ± 35 MPa. Cyclic tests performed with fully reversible strain amplitude (Δε/2) of 0.5%, 0.75%, 1%, and 1.5% showed a unique evolution of stress amplitude with number of cycles that is depicted in Fig. 2(b). The stress amplitude at the onset of cyclic loading increases with an increase in the strain amplitude. However, the evolution of the stress amplitude with number of cycles is different for 0.5% and 0.75% (low) strain amplitude compared to the 1% and 1.5% (high) strain amplitude with the former two showing a significantly higher number of cycles to failure than the latter two. There is initial hardening; that is an increase in stress amplitude with the number of cycles for all the four samples. However, 0.5% and 0.75% strain amplitude samples show softening followed with fluctuations in the stress amplitude. In

2.4. Electron back scatter diffraction (EBSD) and electron channeling contrast imaging Electron back scatter diffraction (EBSD) data acquisition was carried out in FE-SEM from JEOL equipped with electron back scatter

Fig. 2. (a) Engineering stress – engineering strain diagram and (b) variation of stress amplitude with number of cycles in low cycle fatigue test at different strain amplitude for CoCuFeMnNi HEA. 3

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al.

necessary dislocation density (GND) were determined from the EBSD scan and are listed in Table 1. It is observed that there is an increase in the average KAM and GAM value with increase in the strain amplitude. Similarly, the geometrically necessary dislocation density is also found to increase with the increase in strain amplitude. Fig. 3(e) shows that the kernel average misorientation (KAM) distribution shifts to the right and becomes broader as strain amplitude increases, indicating an increase in dislocation density and more heterogeneous distribution of the dislocations with increasing strain amplitude [45]. 3.4. Fractography Detailed analysis of the fractured surface of the fatigue tested samples was carried out using a scanning electron microscope. The illustrative SEM fractographs at lowest (0.5%) and highest (1.5%) strain amplitude are shown in Fig. 4(a–c) and Fig. 4(d–f) respectively. Fig. 4(a) shows formation of striation marks with different orientation observed within different grains. Fig. 4(b) and (c) display crack propagation path with higher number of cycles, which are found to be mostly intergranular in nature. Fig. 4(d) shows the presence of void or inhomogeneity in the material that can act as local stress raisers at grain boundaries, where crack initiation will take place. Once the cracks are initiated the stable propagation of the crack with each cycle is a clear representation of striation mark as shown in Fig. 4(e). However, as shown in Fig. 4(f), the spacing between the striation marks was found to increase with strain amplitude. The increase of strain amplitude causes faster crack growth rate. . Therefore, the mechanism of fatigue failure in the HEA is similar to that in conventional alloys based on crack initiation, stable crack propagation, and final fracture [46]. 4. Discussion

Fig. 3. Loading axis inverse pole figure map near the fracture tip of samples tested in fatigue at strain amplitude of (a) 0.5% (b) 0.75% (c) 1% (d) 1.5% and (e) Kernel average misorientation distribution from all the fatigue tested samples.

4.1. Cyclic stress response The variation of stress amplitude with strain amplitude shows two distinct behavior for the equiatomic CoCuFeMnNi alloy in the present investigation. The stress amplitude reaches a maximum and fluctuates over a prolonged cyclic stage and then decreases sharply for strain amplitude of 0.5% and 0.75%. However, at higher strain amplitudes of 1.0% and 1.5%, value of stress amplitude increases but the regime of high-stress amplitude is very small, and the sample fails after few cycles compared to low strain amplitude samples. Chen et al. [47] have shown that long range motion of dislocations at higher strain amplitude shows a distinct cyclic hardening/softening behaviour compared to short range motion of dislocations at low strain amplitude in Hadfield steel. Agrawal et al. [40] have reported four factors responsible for the strain hardening behavior of the CoCuFeMnNi HEA under monotonic condition, namely; (i) dislocation multiplication, (ii) grain boundary strengthening, (iii) solid solution strengthening and (iv) interaction between dislocations and copper rich nano-clusters. Cyclic loading leads to forward and backward movement of dislocations characterized by long range motion at higher strain amplitude and short-range motion at low strain amplitude. All the aforementioned four interactions play an important role in the reversible nature of dislocation motion in the CoCuFeMnNi high entropy alloy. Grain boundaries act as barrier for the motion of dislocations leading to formation of dislocation pile-up at grain boundaries. During reverse loading, the dislocations can move in the reverse direction at lower stress and annihilate each other, contributing to cyclic softening. This happens for materials with low strain hardening exponent (n ≤ 0.15), while for materials with higher strain hardening co-efficient (n > 0.15) dislocation accumulation occur, leading to cyclic hardening [48]. The alloy under investigation shows higher strain hardening in tension [38] and is a cyclic hardening material at all the strain amplitudes. Both solute atmosphere and copper rich nanoclusters are likely to pin dislocations and reduce the mobility of dislocation. This, in turn, can cause an increase in dislocation density leading to accumulation of intragranular misorientation that can cause accumulation

Table 1 Variation of average kernel average misorientation (KAM), grain average misorientation (GAM) and geometrically necessary dislocation (GND) density for the low cycle fatigue tested samples at different strain amplitude. Strain amplitude

Average KAM°

Average GAM°

Average GND density × 1013 (m−2)

0.5% 0.75% 1% 1.5%

0.48 0.56 0.69 0.91

1.07 1.51 1.53 1.97

2.1 3.2 5.3 5.7

addition, the 1% and 1.5% samples show a decrease in stress amplitude till failure. Despite the fluctuation in stress the amplitude, the 0.5% and 0.75% samples show overall increase in the stress amplitude at the onset of failure. 3.3. Microstructural evolution Fig. 3(a d) shows the loading axis inverse pole figure (IPF) maps of the deformed samples near the fracture tip. The IPF map is characterized by color gradient within the grains indicating slip activity and dislocation dominated deformation in all the samples. Low strain amplitude samples are characterized by voids near grain boundaries and triple junctions while 1% and 1.5% strain amplitude samples show profuse intergranular cracking accompanied with few transgranular cracks near the fracture tip. In order to understand the origin of damage in the fatigue tested samples and establish a link between the operation of slip and damage, intragranular misorientation parameters like kernel average misorientation (KAM), grain average misorientation (GAM) and geometrically 4

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al.

Fig. 4. Secondary electron images of the fractured surface of the low cycle fatigue tested CoCuFeMnNi HEA samples for (a–c) 0.5% and (d–f) 1.5% strain amplitude.

Fig. 5. (a) Coffin-Manson plot and (b) Hysteresis loop at half-life for different strain amplitude for low cycle fatigue tested CoCuFeMnNi HEA samples.

of damage and final failure. At low strain amplitudes, the path length of dislocations is smaller and there is a chance of an edge dislocation to get trapped in the copper rich nano-clusters. At higher strain amplitude, an edge dislocation is likely to cut through the cluster due to the higher slip length. In addition, the overall probability of interaction of an edge dislocation with a copper rich nano-cluster is higher at higher strain amplitude due to higher dislocation density at higher strain amplitude. A balance between dislocation accumulation by the interaction of dislocations with other dislocations, solute environment and nano-cluster and the annihilation of dislocations contribute to a steady state at low strain amplitude. However, for higher strain amplitude, dislocation accumulation is dominant, and the sample fails after few cycles. Fig. 5(a) displays the Coffin-Manson plot, which reveals a logarithmic behavior between strain amplitude and cycles to failure. The determined constant value of c (slope of the curve) is 0.34, which is significantly lower than the conventional range for steels (0.5–0.7) [49]. Recently, a new design strategy has been proposed to improve low cycle fatigue life of FeMnSi alloys, relying on increasing slip planarity by solute pinning that aids in slip reversibility and hence a higher fatigue life [50]. It has already been established that planar partial slip plays an important role in determining deformation behavior of CoCuFeMnNi high entropy alloy and it

is expected that the planar partial slip with smaller slip length will lead to better fatigue performance. However, at higher strain amplitude, it is expected that conventional octahedral slip dominates reducing slip reversibility and contributing to early failure. A similar observation of the transition from dominant partial slip to octahedral slip has been reported in the monotonic deformation by rolling using viscoelastic self-consistent simulations by Tazuddin et al. [38]. Fig. 5(b) illustrates the half-life hysteresis loops plotted in relative co-ordinates where the upper branches of hysteresis loops were almost congruent, revealing a perfect Masingbehavior [23]. This is a clear indication that the dislocation structure might remain unchanged during fatigue at various strain amplitudes. With respect to dislocation motion, the distribution of effective obstacles in Masing-type material may remain constant during fatigue deformation. 4.2. Intragranular misorientation and damage Figs. 6 and 7 show grain reference orientation deviation (GROD) and kernel average misorientation (KAM) maps near the fracture tip of the fatigue tested samples. The GROD map shows homogeneous evolution of GROD for 0.5% strain amplitude sample while all other samples show the heterogeneous distribution of GROD. These samples 5

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al.

Fig. 6. Grain reference orientation deviation (GROD) map of low cycle fatigue tested samples at strain amplitude of (a) 0.5% (b) 0.75% (c) 1.0% (d) 1.5%.

Fig. 7. Kernel average misorientation (KAM) map of low cycle fatigue tested samples at strain amplitude of (a) 0.5% (b) 0.75% (c) 1.0% and (d) 1.5%.

are characterized by grains with lower GROD values as well as grains with very high GROD values, clearly indicating partitioning of strain between the grains. The KAM map shows higher but homogeneous distribution of KAM values for 0.5% sample while other samples show

regions of low KAM and regions of very high KAM. The KAM value is highest near voids and cracks indicating initiation of cracks and voids due to strain incompatibility between neighboring grains and related to geometry necessary dislocations (GND) during the course of cyclic 6

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al.

Fig. 8. Geometrically necessary dislocation (GND) density map of low cycle fatigue tested samples at strain amplitude of (a) 0.5% (b) 0.75% (c) 1.0% and (d) 1.5%.

Fig. 9. Back scattered images with channeling contrast for low cycle fatigue tested samples at (a) 0.5% and (b) 1.5% stain amplitude.

deformation. The strain incompatibility between neighboring grains induces the local misorientation within the grains and can be correlated with the magnitude of the macroscopic plastic strain. Kamaya et al. [51] have proposed that the effect of plastic strain on cracking can be correlated with the evolution of geometrically necessary dislocation (GND) density in the microstructure. The GND density maps near the fracture tip of the fatigue tested samples obtained from EBSD are shown in Fig. 8(a–d), There is a clear indication of increase in GND density in the vicinity of grain boundaries due to significantly strong incompatibility between neighboring grains. The BSE image with electron channeling contrast in Fig. 9(a) near the fracture tip at 0.5% strain amplitude sample hardly shows any contrast within the grains and only orientation contrast is observed. Insignificant change was observed after changing the diffraction condition by rotating the sample, indicating very little dislocation density. In fact, this was uniform throughout the sample with only the orientation contrast visible for different grains. However, for the high strain amplitude deformed sample shown in Fig. 9(b), significant contrast was observed, and there was drastic change in the contrast for different diffraction conditions. This clearly indicates higher dislocation density and more heterogeneous distribution of dislocations at higher strain amplitude. The distribution of dislocation is measured in terms of strain heterogeneity due to locally different stress state during plastic deformation of the material. Due to the different crystallographic

orientation of grains, some grains are more preferentially oriented where the dislocation glide can take place easily than other grains during cyclic deformation. These preferential dislocation motion induces local plastic deformation and creates local stress concentrations and strain gradient. This results in locally different regions of cyclic softening and hardening behavior during cyclic deformation [52]. 4.3. Role of slip and grain boundary character in damage Fig. 10(a–d) shows Schmid factor maps determined using octahedral slip for tension/compression at various strain amplitude (0.5–1.5%) for a range of (0.3–0.5) Schmid factor value. Fig. 10(e and f) displays the variation of the relative fraction with Schmid factor and strain amplitudes to understand the activity of slip systems. The plot indicates distinct value of Schmid factor associated with cyclic plastic stain accommodation in terms of slip activity. It is expected that only single slip system is active for Schmid factor greater than 0.41 and multi-slip is operative for Schmid factor in the range of 0.32–0.41 while no slip is expected for a value less than 0.32 based on slip line trace analysis [53]. The distribution of Schmid factor in all the fatigue tested samples clearly show that with the increase in strain amplitude, there is a reduction in fraction of grains undergoing single slip and most grains undergo multi-slip thereby reducing slip reversibility [53]. Interaction between dislocations and grain boundaries play an 7

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al.

Fig. 10. (a–d) Schmid factor map, (e) Schmid factor distribution for cyclic deformed samples and (f) Schmid factor domain varying as function of strain amplitude.

fraction of special boundaries in general random grain boundary networks. Fig. 11(a and b) shows that the annealing twins generated during processing can act as barrier for crack propagation and void nucleation unlike random high angle grain boundaries during cyclic deformation. Thus, increasing the fraction of special twin boundaries at the expense of random high angle grain boundaries by grain boundary engineering can improve the low cycle fatigue performance of CoCuFeMnNi HEA. Previous researchers have also reported that the grain size, annealing twins and grain boundary network can improve low cycle fatigue performance of conventional alloys [54,55]. Fig. 12(a) and (b) show the interaction of dislocations with nanoclusters. At low stain amplitude the dislocations are pinned by the nanoclusters. However, at higher strain amplitude or high stress applied during cyclic loading, the copper rich nano-clusters get sheared. Fig. 12(c) shows the crack initiation taking place at grain boundary triple junction due to the presence of high geometrically necessary dislocation density around the triple point... There is a greater chance for the crack to initiate at the triple junction and follow an intergranular path. Fig. 11. Loading axis inverse pole figure (IPF) map near crack growth region of the 1.5% stain amplitude fatigue tested CoCuFeMnNi HEA.

5. Conclusions

important role during cyclic deformation and influenced the crack propagation. Therefore, engineering the grain boundary network can improve the mechanical properties of the materials by increasing the

Low cycle fatigue behavior of CoCuFeMnNi high entropy alloy was investigated at different strain amplitudes, and detailed microstructural characterization of the failed samples was carried out to decipher the 8

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al.

Fig. 12. Schematic showing interaction of dislocation with copper rich nano-clusters at (a) low strain amplitude (b) high strain amplitude and (c) crack initiation and propagation in CoCuFeMnNi high entropy alloy subjected to low cycle fatigue.

micro-mechanisms of damage. The important findings from the investigation are provided here.

[2]

1. There is a decrease in low cycle fatigue life with an increase in strain amplitude for CoCuFeMnNi high entropy alloy indicating poor low cycle fatigue behavior manifested by lower Coffin-Manson exponent compared to steels. 2. There is cyclic hardening followed with saturation for low strain amplitude while only softening regime is observed post initial hardening for high strain amplitude regime samples. 3. Higher strain amplitude leads to the evolution of higher intragranular misorientation contributing to the accumulation of damage near grain boundaries. 4. The planar nature of dislocation aids in the reversible motion of dislocations while the presence of solute atmosphere and copper rich nano-clusters retard the reversibility of dislocations. Higher dislocation density at higher strain amplitude leads to higher damage accumulation near grain boundaries which are weakened by the presence of clusters themselves and lead to intergranular failure. 5. Coherent annealing twin boundaries show little damage compared to random high angle grain boundaries indicating that increasing population of special boundaries may improve the low cycle fatigue performance of CoCuFeMnNi high entropy alloy. Further work is required to unequivocally establish the micro-mechanisms of low cycle fatigue failure in CoCuFeMnNi high entropy alloy.

[3]

[4] [5]

[6]

[7] [8] [9] [10] [11]

Acknowledgements

[12]

The authors would like to thank Science and Engineering Research Board under Department of Science and Technology (DST) for funding, IIT Kanpur for experimental facilities and Dr. K.G. Pradeep for Atom Probe Tomography measurements.

[13]

[14]

References

[15]

[1] Yeh JW, Chen SK, Lin SJ, Gan JY, Chin TS, Shun TT, et al. Nanostructured highentropy alloys with multiple principal elements: novel alloy design concepts and

9

outcomes. Adv Eng Mater 2004;6:299–303. https://doi.org/10.1002/adem. 200300567. Cantor B, Chang ITH, Knight P, Vincent AJB. Microstructural development in equiatomic multicomponent alloys. Mater Sci Eng, A 2004;375:213–8. https://doi. org/10.1016/j.msea.2003.10.257. Sharma AS, Yadav S, Biswas K, Basu B. High-entropy alloys and metallic nanocomposites: processing challenges, microstructure development and property enhancement. Materials Science and Engineering: R: Reports 2018;131:1–42. https:// doi.org/10.1016/j.mser.2018.04.003. Zhang Y, Zuo TT, Tang Z, Gao MC, Dahmen KA, Liaw PK, et al. Microstructures and properties of high-entropy alloys. Prog Mater Sci 2014;61:1–93. https://doi.org/10. 1016/j.pmatsci.2013.10.001. Wu Z, Troparevsky MC, Gao YF, Morris JR, Stocks GM, Bei H. Phase stability, physical properties and strengthening mechanisms of concentrated solid solution alloys. Curr Opin Solid State Mater Sci 2017;21:267–84. https://doi.org/10.1016/j. cossms.2017.07.001. Youssef KM, Zaddach AJ, Niu C, Irving DL, Koch CC. A novel low-density, highhardness, high-entropy alloy with close-packed single-phase nanocrystalline structures. Mater Res Lett 2015;3:95–9. https://doi.org/10.1080/21663831.2014. 985855. Senkov ON, Wilks GB, Miracle DB, Chuang CP, Liaw PK. Refractory high-entropy alloys. Intermetallics 2010;18:1758–65. https://doi.org/10.1080/21663831.2014. 985855. Senkov ON, Wilks GB, Scott JM, Miracle DB. Mechanical properties of Nb25Mo25Ta25W25 and V20Nb20Mo20Ta20W20 refractory high entropy alloys. Intermetallics 2011;19:698–706. https://doi.org/10.1016/j.intermet.2011.01.004. Gludovatz B, Hohenwarter A, Catoor D, Chang EH, George EP, Ritchie RO. A fracture-resistant high-entropy alloy for cryogenic applications. Science 2014;345:1153–8. https://doi.org/10.1126/science.1254581. Senkov ON, Scott JM, Senkova SV, Meisenkothen F, Miracle DB, Woodward CF. Microstructure and elevated temperature properties of a refractory TaNbHfZrTi alloy. J Mater Sci 2012;47:4062–74. https://doi.org/10.1007/s10853-012-6260-2. Laplanche G, Kostka A, Horst OM, Eggeler G, George EP. Microstructure evolution and critical stress for twinning in the CrMnFeCoNi high-entropy alloy. Acta Mater 2016;118:152–63. https://doi.org/10.1016/j.actamat.2016.07.038. Li Z, Zhao S, Ritchie RO, Meyers MA. Mechanical properties of high-entropy alloys with emphasis on face-centered cubic alloys. Prog Mater Sci 2018. https://doi.org/ 10.1016/j.pmatsci.2018.12.003. Thurston KV, Gludovatz B, Yu Q, Laplanche G, George EP, Ritchie RO. Temperature and load-ratio dependent fatigue-crack growth in the CrMnFeCoNi high-entropy alloy. J Alloy Compd 2019;794:525–33. https://doi.org/10.1016/j.jallcom.2019. 04.234. Zhang WR, Liaw PK, Zhang Y. Science and technology in high-entropy alloys. Sci Chin Mater 2018;61:2–22. https://doi.org/10.1007/s40843-017-9195-8. Cui P, Ma YM, Zhang LJ, Zhang MD, Fan JT, Dong WQ, et al. Microstructure and mechanical behaviors of CoFeNiMnTixAl1-x high entropy alloys. Mater Sci Eng A 2018;731:124–30. https://doi.org/10.1016/j.msea.2018.06.022.

International Journal of Fatigue 130 (2020) 105258

F. Bahadur, et al. [16] Wang J, Guo T, Li JS, Jia WJ, Kou HC. Microstructure and mechanical properties of non-equilibrium solidified CoCrFeNi high entropy alloy. Mater Chem Phys 2018;210:192–6. https://doi.org/10.1016/j.matchemphys.2017.06.037. [17] Shabani A, Toroghinejad MR, Shafyei A, Loge RE. Evaluation of the mechanical properties of the heat treated FeCrCuMnNi high entropy alloy. Mater Chem Phys 2019;221:68–77. https://doi.org/10.1016/j.matchemphys.2018.09.033. [18] Ghassemali E, Sonkusare R, Biswas K, Gurao NP. In-situ study of crack initiation and propagation in a dual phase AlCoCrFeNi high entropy alloy. J. Alloys Compd. 2017;710:539–46. https://doi.org/10.1016/j.jallcom.2017.03.307. [19] Ghassemali E, Sonkusare R, Biswas K, Gurao NP. Dynamic precipitation at elevated temperatures in a dual-phase AlCoCrFeNi highentropy alloy: an in situ study. Philosophical Magazine Letters 2018;98:400–9. https://doi.org/10.1080/ 09500839.2018.1563728. [20] Paul SK. Effect of anisotropy on ratcheting: an experimental investigation on IFHS steel sheet. Mater Sci Eng, A 2012;538:349–55. https://doi.org/10.1016/j.msea. 2012.01.058. [21] Paul SK, Stanford N, Taylor A. Hilditch, The effect of low cycle fatigue, ratcheting and mean stress relaxation on stress–strain response and microstructural development in a dual phase steel. Int J Fatigue 2015;80:341–8. https://doi.org/10.1016/j. ijfatigue.2015.06.003. [22] Rui S-S, Shang Y-B, Su Y, Qiu W, Niu L-S, Shi H-J, et al. EBSD analysis of cyclic load effect on final misorientation distribution of post-mortem low alloy steel: a new method for fatigue crack tip driving force prediction. Int J Fatigue 2018;113:264–76. https://doi.org/10.1016/j.ijfatigue.2018.04.016. [23] Nikulin I, Sawaguchi T, Kushibe A, Inoue Y, Otsuka H, Tsuzaki K. Effect of strain amplitude on the low-cycle fatigue behavior of a new Fe–15Mn–10Cr–8Ni–4Si seismic damping alloy. Int J Fatigue 2016;88:132–41. https://doi.org/10.1016/j. ijfatigue.2016.03.021. [24] Lambers HG, Rüsing CJ, Niendorf T, Geissler D, Freudenberger J, Maier HJ. On the low-cycle fatigue response of pre-strained austenitic Fe61Mn24Ni6. 5Cr8..5 alloy showing TWIP effect. Int J Fatigue 2012;40:51–60. https://doi.org/10.1016/j. ijfatigue.2012.01.002. [25] Ghosh A, Gurao NP. Effect of crystallographic texture on the planar anisotropy of ratcheting response in 316 stainless steel sheet. Mater Design 2016;109:186–96. https://doi.org/10.1016/j.matdes.2016.07.079. [26] Reddy GP, Kannan R, Mariappan K, Sandhya R, Sankaran S, Rao KBS. Effect of strain rate on low cycle fatigue of 316LN stainless steel with varying nitrogen content: part-I cyclic deformation behavior. Int J Fatigue 2015;81:299–308. https://doi.org/10.1016/j.ijfatigue.2015.07.033. [27] Tofique M, Bergström J, Svensson K, Johansson S, Peng R. ECCI/EBSD and TEM analysis of plastic fatigue damage accumulation responsible for fatigue crack initiation and propagation in VHCF of duplex stainless steels. Int J Fatigue 2017;100:251–62. https://doi.org/10.1016/j.ijfatigue.2017.03.035. [28] Ghosh A. In-plane anisotropy in deformation micro-mechanism of commercially pure titanium during monotonic tension and cyclic loading. Fratt Integrit Strutt 2019;48:585–98. https://doi.org/10.3221/IGF-ESIS.48.57. [29] Sasaoka S, Arakawa J, Akebono H, Sugeta A, Shirai Y, Nakayama E, et al. The effects of crystallographic orientation on fatigue crack initiation behavior in Ti-6Al4V. Int J Fatigue 2018;117:371–83. https://doi.org/10.1016/j.ijfatigue.2018.07. 039. [30] Zhang P, Zhu Q, Hu C, Wang CJ, Chen G, Qin HY. Cyclic deformation behavior of a nickel-base superalloy under fatigue loading. Mater Design 2015;69:12–21. https:// doi.org/10.1016/j.matdes.2014.12.047. [31] Brien V, Décamps B. Low cycle fatigue of a nickel based superalloy at high temperature: deformation microstructures. Mater Sci Eng A 2001;316:18–31. https:// doi.org/10.1016/S0921-5093(01)01235-7. [32] Hemphill MA, Yuan T, Wang GY, Yeh JW, Tsai CW, Chuang A, et al. Fatigue behavior of Al0.5CoCrCuFeNi high entropy alloys. Acta Mater 2012;60:5723–34. https://doi.org/10.1016/j.actamat.2012.06.046. [33] Tang Z, Yuan T, Tsai CW, Yeh JW, Lundin CD, Liaw PK. Fatigue behavior of a wrought Al0.5CoCrCuFeNi two-phase high-entropy alloy. Acta Mater 2015;99:247–58. https://doi.org/10.1016/j.actamat.2015.07.004. [34] Niendorf T, Wegener T, Li ZM, Raabe D. Unexpected cyclic stress-strain response of dual-phase high-entropy alloys induced by partial reversibility of deformation. Scr Mater 2018;143:63–7. https://doi.org/10.1016/j.scriptamat.2017.09.013.

[35] Liu K, Nene S, Frank M, Sinha S, Mishra R. Extremely high fatigue resistance in an ultrafine grained high entropy alloy. Appl Mater Today 2019;15:525–30. https:// doi.org/10.1016/j.apmt.2019.04.001. [36] Liu K, Nene S, Frank M, Sinha S, Mishra R. Metastability-assisted fatigue behavior in a friction stir processed dual-phase high entropy alloy. Mater Res Lett 2018;6:613–9. https://doi.org/10.1080/21663831.2018.1523240. [37] Tazuddin, Biswas K, Gurao NP. In the quest of single phase multi-component multiprincipal high entropy alloys. J Alloy Compd 2017;697:434–42. https://doi. org/10.1016/j.jallcom.2016.11.383. [38] Tazuddin, Biswas K, Gurao NP. Deciphering micro-mechanisms of plastic deformation in a novel single phase fcc-based MnFeCoNiCu high entropy alloy using crystallographic texture. Mater Sci Eng A 2016;657:224–33. https://doi.org/10. 1016/j.msea.2016.01.065. [39] Sonkusare R, Janani PD, Gurao NP, Sarkar S, Sen S, Pradeep KG, et al. Phase equilibria in equiatomic CoCuFeMnNi high entropy alloy. Mater Chem Phys 2018;210:269–78. https://doi.org/10.1016/j.matchemphys.2017.08.051. [40] Agarwal R, Sonkusare R, Jha SR, Gurao NP, Biswas K, Nayan N. Understanding the deformation behavior of CoCuFeMnNi high entropy alloy by investigating mechanical properties of binary ternary and quaternary alloy subsets. Mater Design 2018;157:539–50. https://doi.org/10.1016/j.matdes.2018.07.046. [41] Sonkusare R, Khandelwal N, Ghosh P, Biswas K, Gurao NP. A comparative study on the evolution of microstructure and hardness during monotonic and cyclic high pressure torsion of CoCuFeMnNi high entropy alloy. J Mater Res 2019;34:732–43. https://doi.org/10.1557/jmr.2018.479. [42] Zhang H, Pradeep KG, Mandal S, Ponge D, Springer H, Raabe D. Dynamic straininduced transformation: an atomic scale investigation. Scr Mater 2015;109:23–7. https://doi.org/10.1016/j.scriptamat.2015.07.010. [43] Basu I, Pradeep K, Mießen C, Barrales-Mora L, Al-Samman T. The role of atomic scale segregation in designing highly ductile magnesium alloys. Acta Mater 2016;116:77–94. https://doi.org/10.1016/j.actamat.2016.06.024. [44] Zaefferer S, Elhami NN. Theory and application of electron channelling contrast imaging under controlled diffraction conditions. Acta Mater 2014;75:20–50. https://doi.org/10.1016/j.actamat.2014.04.018. [45] Zhong Y, Yin F, Sakaguchi T, Nagai K, Yang K. Dislocation structure evolution and characterization in the compression deformed Mn–Cu alloy. Acta Mater 2007;55:2747–56. https://doi.org/10.1016/j.actamat.2006.12.012. [46] Ritchie RO. Mechanisms of fatigue-crack propagation in ductile and brittle solids. Int J Fract 1999;100:55–83. https://doi.org/10.1023/A:1018655917051. [47] Chen C, Lv B, Wang F, Zhang FC. Low-cycle fatigue behaviors of pre-hardening Hadfield steel. Mater Sci Eng 2017;695:144–53. https://doi.org/10.1016/j.msea. 2017.03.107. [48] Li P, Li S, Wang Z, Zhang Z. Fundamental factors on formation mechanism of dislocation arrangements in cyclically deformed fcc single crystals. Prog Mater Sci 2011;56:328–77. https://doi.org/10.1016/j.pmatsci.2010.12.001. [49] Roessle M, Fatemi A. Strain-controlled fatigue properties of steels and some simple approximations. Int J Fatigue 2000;22:495–511. https://doi.org/10.1016/S01421123(00)00026-8. [50] Sawaguchi T, Nikulin I, Ogawa K, Sekido K, Takamori S, Maruyama T, et al. Designing Fe–Mn–Si alloys with improved low-cycle fatigue lives. Scr Mater 2015;99:49–52. https://doi.org/10.1016/j.scriptamat.2014.11.024. [51] Kamaya M. Measurement of local plastic strain distribution of stainless steel by electron backscatter diffraction. Mater Charact 2009;60:125–32. https://doi.org/ 10.1016/j.matchar.2008.07.010. [52] Nellessen J, Sandlöbes S, Raabe D. Effects of strain amplitude, cycle number and orientation on low cycle fatigue microstructures in austenitic stainless steel studied by electron channelling contrast imaging. Acta Mater 2015;87:86–99. https://doi. org/10.1016/j.actamat.2014.12.024. [53] Das A. Slip system activity during cyclic plasticity. Metallurg Mater Trans A 2014;45:2927–30. https://doi.org/10.1007/s11661-014-2295-0. [54] Kobayashi S, Hirata M, Tsurekawa S, Watanabe TV. Grain boundary engineering for control of fatigue crack propagation in austenitic stainless steel. Proc Eng 2011;10:112–7. https://doi.org/10.1016/j.proeng.2011.04.021. [55] Das A. Grain boundary engineering: fatigue fracture. Phil Mag 2017;97:867–916. https://doi.org/10.1080/14786435.2017.1285072.

10