ArticlePDF Available

Revisiting the role of prestrain history in the mechanical properties of ultrafine-grained CoCrFeMnNi high-entropy alloy

Authors:

Abstract and Figures

The yield strength of face-centered cubic (FCC) alloy is always insufficient for applications. In this work, different prestrain histories were imposed to improve the yield strength and strain-hardening capability of an ultrafine-grained (UFG) CoCrFeMnNi high-entropy alloy (HEA). In contrast to the specimens prestrained at 293 K, the specimens prestrained at 77 K possess higher yield strength, elongation, and strain-hardening capability, which were intensely related to the formation of deformation twins. The cryogenic strengthening magnitude is found to be strongly associated with the grain size, but slightly affected by dislocations. By modulating the prestrain history, an ultrahigh yield strength of 1.84 GPa and a considerable uniform elongation of 13% were achieved at 77 K in the CoCrFeMnNi HEA. Hence, imposing prestrain on HEAs at 77 K could be an efficient strategy to harmonize the mechanical properties of the FCC HEAs, which would enrich the application of HEAs in cryogenic fields.
Content may be subject to copyright.
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
1
Revisiting the role of prestrain history in the mechanical properties of
ultrafine-grained CoCrFeMnNi high-entropy alloy
S.J. Sun a, b, Y.Z. Tian c, d,*, H.R. Lin a, b, Z.J. Wang e, Z.F. Zhang a, b **
a Shi-changxu Innovation Center for Advanced Materials, Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China
b School of Materials Science and Engineering, University of Science and Technology of China, Hefei 230026, China
c Key Laboratory for Anisotropy and Texture of Materials (Ministry of Education), School of Materials Science and Engineering, Northeastern
University, Shenyang 110819, China
d Research Center for Metal Wires, Northeastern University, Shenyang 110819, China
e State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi'an 710072, China
Abstract
The yield strength of face-centered cubic (FCC) alloy is always insufficient for applications. In this work, different
prestrain histories were imposed to improve the yield strength and strain-hardening capability of an ultrafine-grained
(UFG) CoCrFeMnNi high-entropy alloy (HEA). In contrast to the specimens prestrained at 293 K, the specimens
prestrained at 77 K possess higher yield strength, elongation, and strain-hardening capability, which were intensely
related to the formation of deformation twins. The cryogenic strengthening magnitude is found to be strongly associated
with the grain size, but slightly affected by dislocations. By modulating the prestrain history, an ultrahigh yield strength
of 1.84 GPa and a considerable uniform elongation of 13% were achieved at 77 K in the CoCrFeMnNi HEA. Hence,
imposing prestrain on HEAs at 77 K could be an efficient strategy to harmonize the mechanical properties of the FCC
HEAs, which would enrich the application of HEAs in cryogenic fields.
Keywords: High-entropy alloy; Prestrain; Deformation twins; Cryogenic strengthening; Yield strength
* Corresponding authors
E-mail addresses: tianyanzhong@mail.neu.edu.cn (Y.Z. Tian), zhfzhang@imr.ac.cn (Z.F. Zhang).
1. Introduction
In contrast to the conventional alloys with a principal element, high-entropy alloys (HEAs), also known as multi-
principal-element alloys (MPEAs) or complex concentrated alloys (CCAs), usually contain multiple principal elements
[1-6]. The equiatomic or near-equiatomic compositions in HEAs maximize the configuration entropy, which generally
results in the formation of single-phase microstructure [7-10]. Among these HEAs, single-phase HEAs with face-
centered cubic (FCC) structure were widely investigated because of their outstanding properties, such as the irradiation
resistance, corrosion resistance and mechanical performance especially at cryogenic temperature [11-15]. In particular,
there is no ductile-to-brittle (DTB) transition during the deformation process of HEAs and medium-entropy alloys
(MEAs) at cryogenic temperature, which usually occurs in ferritic, martensitic and duplex stainless steels [14, 16-23].
The superior cryogenic mechanical properties of the HEAs or MEAs are mainly due to the deformation twins, which
introduce new interfaces and decrease the mean free path of dislocations, as referred as the dynamic Hall-Petch effect
[16]. However, deformation twins are difficult to occur at room temperature due to the insufficient flow stress to
approach the critical twinning stress [16, 24, 25]. In contrast, when the deformation temperature decreases to the liquid
nitrogen temperature, the flow stress increases dramatically to activate deformation twins, and multiple deformation
mechanisms including dislocations, stacking faults and deformation twins were activated during the deformation process,
resulting in the simultaneous improvement of strength and ductility [24]. Jo et al. designed a HEA with partially
recrystallized microstructures containing retained deformation twins, which exhibited ultra-high strength at liquid
nitrogen temperature [23]. Furthermore, in CoCrNi MEA with lower stacking fault energy (SFE) relative to
CoCrFeMnNi HEA, the twinning stress could be reached at lower strains, which results in an extended strain-hardening
range, leading to much more excellent mechanical properties compared to CoCrFeMnNi HEA [26]. Albeit the dramatic
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
2
temperature dependence of mechanical properties of HEAs, the intrinsic mechanisms for enhanced yield strength (σYS)
and strain-hardening behavior at cryogenic temperature is unclear.
Besides the deformation temperature, grain size can also influence the deformation mechanisms especially the
twinning capability. Our previous study on the ultrafine-grained (UFG) CoCrFeMnNi HEA indicates that deformation
twins and dislocations could be activated at liquid nitrogen temperature, while only dislocations could be captured at
room temperature during tensile tests [24]. In this case, an UFG CoCrFeMnNi HEA was chosen as model material and
different prestrains at 293 K and 77 K were applied to introduce distinct defects. Then the roles of dislocations and
deformation twins in the cryogenic strengthening effect were clarified by comparing tensile properties at 293 K and 77
K. We aimed to unveil the roles of prestored dislocations and twins in the yield strength and subsequent strain-hardening
capability, and achieve a superior balance of strength and ductility to extend the cryogenic applications of HEAs.
2. Experimental method
2.1. Starting material
A bulk CoCrFeMnNi HEA ingot with 110 mm in diameter and 90 mm in height was fabricated using magnetic
levitation melting technique with the starting elemental metals (>99.7% wt%). The bulk ingot was homogenized at 1373
K for 2 h. Then the ingot was hot forged at 1273 K and cold rolled to sheets with 1 mm in thickness at room temperature.
Details about the casting and manufacture procedures can be found in the earlier work [27]. The original specimens
(defined as ORI) were obtained by annealing the cold-rolling sheets for 5 minutes at 998 K.
2.2. Mechanical testing
Tensile specimens with rectangular dog-bone shape (gauge dimensions, 10×2×1 mm) were cut from the recrystallized
sheets, and the tensile loading direction was parallel to the rolling direction. The specimens were mechanically ground
using 400-2000 grid sized fine sandpapers before the tensile tests. One group of specimens were firstly stretched to
specific plastic strains (10% and 20%) at room temperature, then unloaded. The other group of specimens were stretched
to specific plastic strains (10%, 20% and 30%) at liquid nitrogen temperature, then unloaded. All prestrain tests were
conducted on an Instron 5982 testing machine with a strain rate of 10-3 s-1. The sample code and the corresponding
processing states were also listed in Table 1. After the prestrained tests, all the specimens were ground afresh using the
SiC paper. At room temperature, the strains could be measured with an extensometer attached to the tensile specimens.
At liquid nitrogen temperature, the extensometer could not be used, so strains had to be determined from the crosshead
displacement. However, the displacement of the crosshead is typically larger than the real strains. In this work, we
calculated the difference between the extensometer and crosshead displacements at room temperature and assumed the
same difference at liquid nitrogen temperature. The strains at both temperatures were corrected by the difference.
Furthermore, the elongations of specimens after fracture were calculated using the change in spacing of the indents with
10 mm initial interval, and they showed good agreement with the above corrected strains. The tensile tests were
conducted on the testing machine with a strain rate of 10-3 s-1 at 293 K and 77 K until fracture. During the cryogenic
tests, the specimens were firstly immersed in the liquid nitrogen for about 5 minutes and the tests were performed on
the specimens immersed in the liquid nitrogen. Several tests for each condition were conducted to ensure the
reproducibility of the experiments.
Table 1. Sample code and processing states.
Sample code
Processing states
ORI
Fully recrystallized UFG specimens
PL1
ORI prestrained to strain of 10% and unloaded at 77 K
PL2
ORI prestrained to strain of 20% and unloaded at 77 K
PL3
ORI prestrained to strain of 30% and unloaded at 77 K
PR1
ORI prestrained to strain of 10% and unloaded at 293 K
PR2
ORI prestrained to strain of 20% and unloaded at 293 K
2.3. Microstructure characterization
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
3
The prestrained specimens were characterized by electron backscattering diffraction (EBSD), electron channeling
contrast imaging (ECCI) and transmission electron microscopy (TEM). The microstructural and crystallographic
information were obtained by an EBSD system equipped on a field emission scanning electron microscope (FE-SEM,
JSM 7100F). An accelerated voltage of 20 kV, a working distance of ~15 mm and a step size of 40 nm were used to
acquire the EBSD data. The TSL OIM Analysis 7 software was used to analyze the ESBD data. The microstructures of
the specimens were characterized by the ECCI technique conducted on a LEO Supra 35 FE-SEM, and a working distance
of ~6 mm was used to acquire the ECCI images. TEM analyses were conducted on an FEI Tecnai F20 with an accelerated
voltage of 200 kV. The specimens for EBSD and ECCI characterization were prepared by electropolishing at 30 V using
an electrolyte consisting of 90 vol% ethanol and 10 vol% perchloric acid at room temperature. Thin foils for TEM
characterization were prepared by twin-jet electropolishing method in the above-mentioned solution at 253 K under a
working voltage of 20 V.
3. Results and discussions
3.1. Original microstructure of the UFG specimen
Fig. 1 shows the typical microstructure of the ORI specimen. The observation plane is the rolling direction-normal
direction plane (RD-ND). The ECCI image and grain orientation map show that the equiaxed grains containing
annealing twins distribute homogeneously, and the inverse pole figure (IPF) inset shows the maximum pole density
value is 2.1, which indicates that the crystallographic orientation is nearly random and the texture is weak. The average
grain size determined by the linear intercept method is about 0.65 μm by counting high-angle grain boundaries including
twin boundaries.
Fig.1. Typical microstructure of the ORI CoCrFeMnNi HEA. (a) ECCI image, (b) grain orientation map. The color code used for grain orientation
map is IPF coloring-Z inserted in the upper right corner in Fig.1 (b).
3.2. Mechanical behavior of prestrained HEAs
Various prestrain levels (strain of 10%, 20% and 30%) were applied to ORI HEA specimens at 293 K (simplified as
PR1 and PR2) and 77 K (PL1, PL2 and PL3), respectively, as given in Fig. 2(a), which also exhibited promising
experimental reproducibility. Figs. 2(b) and (c) show the uniaxial tensile engineering stress-strain curves of the ORI and
prestrained HEAs at 293 K and 77 K, respectively. It is found that the yield point phenomenon occurred at both
temperatures, as shown in Fig. 1(a). The recrystallized UFG grains are clean inside and contain nearly no dislocations.
In that case, high stress is required to activate dislocations, then a lower stress is enough to glide these activated
dislocations, exhibiting the yield point phenomenon. [28, 29]. The yield strength (σYS) and uniform elongation (εUE) are
summarized in Fig.3. Compared to the ORI specimen, the prestrain has led to significant increases in σYS of the HEAs,
but slight decreases in εUE, as shown in Fig. 2 and Fig. 3. With the increase of applied prestrain level, the σYS increases
and the εUE decreases at both temperatures. Furthermore, for the PR specimens, the σYS increases slowly with the increase
of prestrain levels, which shows weak sensitivity to the prestrain levels. For the PL specimens, the σYS increases sharply
with the increase of prestrain levels, which shows strong sensitivity to the prestrain levels. When the deformation
temperature decreases from 293 K to 77 K, the strength, ductility, as well as the strain-hardening ability increase
remarkably. It is found that the PL specimens exhibit higher strength and ductility relative to the PR specimens with the
same prestrain level, as discussed in the previous work [28]. The PL3 specimen exhibits striking properties with an
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
4
ultrahigh σYS of 1.84 GPa and a εUE of 13% at 77 K, which are even better than the partially recrystallized CoCrFeMnNi
HEA [30]. For the engineering structural materials, the σYS is a critical value that limits the utmost stress that can be
loaded on a material without bringing about the plastic deformation. It can be seen easily that σYS could be enhanced
efficiently through managing prestrain on the HEAs. Hence, modifying the prestrain history and building advanced
defect architecture could be an efficient way to design high strength HEAs used as cryogenic components.
Fig. 2. (a) Tensile stress-strain curves of the ORI HEA prestrained to various amounts of strain at 293 K (named as PR1 and PR2) and 77 K (named
as PL1, PL2 and PL3), (b) tensile stress-strain curves of the ORI, PR and PL specimens at 293 K, (c) tensile stress-strain curves of the ORI, PR
and PL specimens at 77 K.
Fig. 4 shows the cryogenic strengthening magnitude of the ORI and prestrained specimens, which was calculated by
the following equation:
     (1)
where σYS,77 K and σYS,293 K are the yield strengths of specimens at 77 K and 293 K, respectively. For the PR specimens,
ΔσYS increases slightly after imposing higher prestrain, indicating that the strength enhancement due to temperature
effect is insensitive to the imposed prestrain at 293 K. In contrast, ΔσYS increases dramatically after imposing prestrain
at 77 K. These results indicate that the prestrain history can be crucial for the cryogenic strengthening effect. In fact,
this cryogenic strengthening phenomenon has been frequently detected in HEAs, twinning-induced plasticity (TWIP)
steels and Cu-Al alloys [20, 31-33], but the intrinsic mechanism for the cryogenic strengthening magnitude is still
unclear. It is thus desired to characterize the microstructures of the prestrained specimens to explore evidences that
induce different cryogenic strengthening effects.
Fig. 3. Plots of yield strength (σYS) and uniform elongation (εUE) of ORI and prestrained HEAs: (a) σYS for ORI and PR specimens, (b) σYS for ORI
and PL specimens, (c) εUE for ORI and PR specimens and (d) εUE for ORI and PL specimens.
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
5
Fig. 4. The cryogenic strengthening magnitude ΔσYS for PR and PL specimens with different prestrain histories.
3.3. Microstructures of the HEAs with different prestrain histories
It is particularly found that the variations in dislocations and deformation twins in the prestrained specimens interfere
the ΔσYS, and examining the density of dislocations and deformation twins are insightful to understand the mechanism
of the cryogenic strengthening in the prestrained CoCrFeMnNi HEA. Plastic deformation involves many dislocations
in the microstructures of polycrystalline materials, and the dislocations usually consist of geometrically-necessary
dislocations (GNDs) and statistically-stored dislocations (SSDs) [34-36]. Generally, the density of SSDs with a net-zero
Burgers vector is irrelevant to the characteristics of the microstructures, while the density of GNDs with a non-zero
accumulated Burgers vector is dependent on the microstructural characteristics, such as grain size, texture and phase
[35, 37]. Therefore, the variations of the dislocations in the ORI and prestrained specimens could be regarded as the
variations of the GNDs. The density of the GNDs could be estimated through the EBSD misorientation calculated by
the following equation [38, 39]:
 
 (2)
where β is a constant determined by the type of dislocation boundaries, θ is the average misorientation within a certain
distance Δx, and b is the Burgers vector.
Fig. 5. EBSD results of kernel average misorientation (KAM) maps of the prestrained HEAs: (a) PL1, (b) PL2, (c) PL3, (d) PR1 and (e) PR2; (f)
KAM value distributions of the prestrained HEAs, and the enlarged misorientation curve of the prestrained HEAs are displayed in the inset.
The average misorientation θ in the above Eq. (2) is usually referred as the kernel average misorientation (KAM),
which quantifies the average misorientation around a kernel point with respect to the next-nearest neighbors. The KAM
maps and corresponding peak values of the prestrained HEAs are shown in Fig. 5. With increasing the prestrain level,
the KAM values increase obviously, and the density of the GNDs would be enhanced, as calculated by Eq. (2). It is also
found that the KAM values in PL specimens are slightly higher relative to the PR specimens at the same prestrain level.
High KAM values were observed near the grain boundaries, which indicates that more GNDs were accumulated near
the grain boundaries during the plastic deformation. The average KAM values and the GNDs densities of the prestrained
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
6
specimens were obtained through analyzing the EBSD data, as shown in Fig. 6. For the prestrained specimens, with the
increase of the prestrain level, the density of the GNDs raises evidently. It is worth noting that the density of GNDs in
PR specimens is only slightly lower than that in the PL specimens with the same prestrain level. This could be resulted
from the low SFE of CoCrFeMnNi HEA, which inhibits dislocation recovery even at 293 K. Hence, coupling with the
fact that the significant cryogenic strengthening in PL HEAs relative to the PR HEAs, while they had roughly the same
density of GNDs, we deduced that the density of GNDs did not contribute notably to the ΔσYS. Therefore, other factors
including deformation twins may contribute significantly to ΔσYS in the PL specimens.
Fig. 6. EBSD results of (a) average KAM value and (b) geometrically necessary dislocations (GND) density of the prestrained HEAs.
Fig. 7. Typical microstructures of the PR HEAs: (a), (b) ECCI and TEM images of the PR1 HEA. (c), (d) ECCI and TEM images of the PR2 HEA.
Fig. 8. Typical microstructures of the PL HEAs: (a), (b) ECCI and TEM images of the PL1 HEA. (c), (d) ECCI and TEM images of the PL2 HEA.
(e), (f) ECCI and TEM images of the PL3 HEA.
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
7
Fig. 7 and Fig. 8 present microstructures in the PR and PL specimens, respectively. The interactions of the dislocations
were captured in the PR HEAs, but no deformation twins were detected, as shown in Fig. 7. It is also found that the
quantity of the dislocations accumulated in grains and annealing twins raises evidently with the increase of the prestrain
level, which is consistent with the above-mentioned density of GNDs. The absence of deformation twins can be related
to the constrained twinning capability with grain refinement [25]. These results substantiate that the dislocations are the
main deformation modes in the UFG HEA, which is consistent with the reported work [24]. In contrast, the ECCI and
TEM images in Fig. 8 reveal that deformation twins were activated in the PL specimens, as indicated by the arrows, and
the fraction of deformation twins raises with the increase of the prestrain level. The interaction between dislocations and
deformation twins were captured, and more twinning systems were activated with increasing prestrain level, as shown
in Fig. 8b, c and d. The volume fraction (f) of the deformation twins was calculated through measuring the average twin
width (t) and mean twin spacing (λ) in the TEM images using the method [17, 40]:
 

 (3)
The calculated volume fraction is shown in Fig. 9. Furthermore, the volume fraction was also evaluated by counting
the area fraction of deformation twins in the ECCI images as reported by Luo and Huang [41], which is consistent with
the calculated values. With increasing the prestrain level, both the average twin width and the fraction increase, and the
mean twin spacing decreases, as shown in Fig. 9. In addition, with the increase of prestrain level, more twinning systems
from one system in PL1 HEA to three systems in PL3 HEA were activated, as shown in the high-resolution TEM in Fig.
10. The fine twins could become obstacles for hindering dislocation slip, which is similar to the effect of grain
boundaries that lead to the Hall-Petch strengthening effect [42]. Hence, the σYS increases sharply with the increase of
prestrain levels in PL specimens, which shows sensitive to the prestrain levels. In contrast, the σYS increases slowly with
the increase of prestrain levels in PR specimens because only dislocations were introduced in the microstructures of
specimens prestrained at room temperature.
Fig. 9. (a) Average twin width and mean twin spacing of the deformation twins, (b) the volume fraction of deformation twins in the PL
specimens.
Fig. 10. High-revolution TEM images for the PL specimens: (a) PL1 HEA, twinning occurring in one system, (b) PL2 HEA, twinning occurring
in two systems, (c) PL3 HEA, twinning occurring in three systems.
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
8
3.4. Cryogenic strengthening magnitude
Fig. 11. Plots of the cryogenic strengthening magnitude against grain size for various FCC alloys [16, 19, 21-23, 26, 43, 49-52].
Considering the significant difference of ΔσYS of the PR and PL specimens, as shown in Fig. 4, the roles of dislocations
and deformation twins were considered and discussed as follows. For the ORI specimen without dislocations and
deformation twins, the yield strength could be calculated by the following empirical equation [43]:
     (4)
where d is grain size in micrometer, T is the temperature in Kelvin. Therefore, the ΔσYS for ORI specimens resulted from
the grain boundaries (σg) and friction stress (σ0) is about 370 MPa, which is determined directly by experimental results.
In the prestrained HEAs with dislocations and deformation twins, the σYS could be written as:
      (5)
where σ0, σg, σρ and σT are the contribution of friction stress, grain boundaries, prestored dislocations and deformation
twins, respectively. Here the contribution of dislocations can be expressed by the classical Taylor relation [44]:
  (6)
where M is the Taylor factor, α is a geometric constant, G is shear modulus, b is Burgers vector and ρ is dislocation
density. Usually, the contribution of the deformation twins to the strength could be calculated as following equation [24,
45]:
 (7)
where f is the fraction of deformation twins, λ is the mean spacing of the deformation twins, and K is the Hall-Petch
constant for twinning, which is several times of the Hall-Petch constant for slip (K=CkS, where C is a constant) [46].
For the PR specimens with only dislocations, the grain boundaries and dislocations contribute to σYS, and ΔσYS can
be written as below:
      (8)
With decreasing the temperature from 293 K to 77 K, b changes very slightly and can be supposed to be a constant.
Thus, a combined equation for ΔσYS in PR specimens through the Eqs. (6) and (8) gives
     (9)
Here the shear modulus G of CoCrFeMnNi HEA increases slightly as the temperature decreases from 293 K to 77 K
[47, 48]. Hence, the enhanced ΔσYS in PR specimens with increasing the prestrain level results from the increasingly
stored dislocations. But the limited magnitude is induced by the weak dependence of G on temperature.
For the PL specimens with dislocations and deformation twins, combining Eqs. (5), (6) and (7), the ΔσYS can be
expressed as below:
         (10)
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
9
The first two items on the right-hand side of the above equation approximately equal to the two items on the right-
hand side of Eq. (9) due to the roughly same density of dislocations. The Hall-Petch constant for twinning (K=CkS) in
the last item has a strong temperature dependence [43], so it would amplify the strengthening effect caused by
deformation twins in the PL specimens. In summary, introducing more boundaries in the microstructures would lead to
the superior cryogenic strengthening magnitude. In this connection, we plot the difference in the normalized yield
strength (σYS/G) between two temperatures as a function of grain size, as shown in Fig. 11. Data of typical FCC metals,
i.e. copper and aluminum, were also included. Datum points of the CoCrFeMnNi HEA were schematically fitted for
clarification. It is found that ΔσYS increases remarkably with grain refinement.
3.5. Strain-hardening capability of the prestrained HEAs
Fig. 12 shows true stress-strain and strain-hardening curves for the typical PL2 and PR2 HEAs after tensile tests at
293 K and 77 K. It is well known that materials would be necking when strain-hardening rate Θ is equal to true stress
σ, which is described as the Considère’s criterion [53, 54]. The true stress-strain curves and strain-hardening curves
before necking points for each state are shown in Fig. 12. At the same deformation temperature, the strain-hardening
rate of the PL2 is higher than that of PR2 HEAs. At 293 K, PR2 HEA suffers early necking while PL2 HEA possesses
promising ductility, indicating that the prestrain history can be critical for the optimization of strength and ductility [28].
At 77 K, both PR2 and PL2 exhibit ultrahigh σYS of ~1.5GPa along with high εUE of ~0.2, as shown in Fig. 12(a). For
the strain-hardening behavior, it is found that the strain-hardening rate of PL2 HEA is higher than PR2 HEA at 293 K
and 77 K, indicating that imposing prestrain at 77 K is benefinical for the mechanical property optimization. It is noted
that the strain-hardening rate at 77 K is apparently higher than 293 K, as shown in Fig. 12(b). Such a high strain-
hardening ability at 77 K is largely related to the deforamtion twins activated at cryogenic temperature. It is reported
that the twins are very effective in enhancing strain-hardening rate of alloys, such as Mg alloys and TWIP steels, which
results in the hump in strain-hardening curve [32, 55, 56]. Furthermore, lowering the temperature would facilitate the
suppression of dynamic recovery and promote the dislocation storage, which could enhance the work hardening ability
[55, 57]. During the plastic deformation process, the dislocations could be dissociated when interacting with twin
boundaries, which could avoid excessive stress concentration due to piling-up of dislocation, allowing the materials to
deform with retaining ductility [58].
Fig. 12. (a) Typical true stress-strain and (b) strain-hardening curves for PL2 and PR2 HEAs at 293 K and 77 K. The necking points for each
state were indicated in (a).
Generally, the dislocation evolution resulted from the plastic strain could be calculated by the Orowan equation [53]:

 (11)
where ρgen is the density of dislocations, γp is the shear strain, k is a correction parameter, b is the Burger vector, and l is
an average distance that dislocations travel, as referred as the mean free path. When grain and twin boundaries are
considered as the main barriers for dislocation slip, l could be calculated by the following equation [59]:
(12)
where d is the grain size and λ is the twin spacing.
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
10
Substituting Eq. (12) into Eq. (11) gives


(13)
Eq. (12) indicates that the mean free path would be reduced with introducing deformation twins in the grains. As a
result, ρgen increases as calculated by Eq. (13), which would improve the strength. Furthermore, the interactions of
dislocations and deformation twins afford room for storing dislocations efficiently, which would facilitate pronounced
strain-hardening capability [60].
A comparison of the σYS and εUE of the prestrained HEAs with those of reported HEAs is shown in Fig. 13. Although
FCC HEAs are soft, σYS could be enhanced by introducing various defects, i.e. dislocations and deformation twins
through prestrain. It is clear that modulating prestrain history brings a superior balance of σYS and εUE in CoCrFeMnNi
HEA. To our knowledge, the cryogenic σYS of 1.84 GPa and εUE of 13% in PL3 HEA are particularly noteworthy, and
such high cryogenic σYS surpassed those of the existing CoCrFeMnNi HEAs, as shown in Fig. 13(b). Furthermore, the
strength and ductility balance also shows superiority in contrast to other HEAs and MEAs. These results indicate that
the strength and ductility of HEAs can be harmonized by modulating the prestrain history, providing a simple but
efficient strategy to process HEAs for potential applications in cryogenic fields.
Fig. 13. Plot of σYS and εUE for CoCrFeMnNi HEAs at (a) 293 K and (b) 77 K in this work. Data of other HEAs and MEAs as reported in the
previous work [14, 16-18, 20-23, 26, 30, 31, 43, 61-64].
4. Conclusions
Fully recrystallized UFG HEA (ORI) with an average grain size of 0.65 μm was obtained by cold rolling and annealing
treatment. ORI specimens were tensioned to specified strains at 293 K and 77 K and unloaded to obtain PR and PL
specimens. ORI, PR and PL specimens were further tensioned at 293 K and 77 K to investigate the mechanical behavior.
The following conclusions were drawn:
(1) Dislocation features were captured in the PR specimens. In the PL specimens, both dislocations and deformation
twins were stored.
(2) The dependence of cryogenic strengthening magnitude (ΔσYS=σYS,77K-σYS,293K) on prestrain level shows distinct
features in the PR and PL specimens. ΔσYS rose slightly with increasing the prestrain level in the PR specimens,
which was induced by the weak dependence of G on temperature in the dislocation strengthening effect. ΔσYS rose
dramatically with increasing the prestrain level in the PL specimens, which was attributed to the twin boundary
strengthening effect via the remarkable change of Hall-Petch slope with temperature.
(3) Cryogenic strengthening magnitude (ΔσYS=σYS,77K-σYS,293K) is found to be negatively correlated with the grain size
in various FCC alloys.
(4) In contrast to the PR specimens, the PL specimens exhibit superior mechanical properties at 293 K and 77 K due to
the introduction of deformation twins, which reduce the mean free path of dislocations and contribute significantly
to the strength and strain-hardening capability.
(5) Modulating the microstructures by applying prestrain can lead to an ultrahigh σYS of 1.84 GPa and a considerable
εUE of 13% at 77 K in PL3 specimen.
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
11
Acknowledgements
This work was financially supported by the Fundamental Research Funds for the Central Universities (No. N180204015), the fund
of the State Key Laboratory of Solidification Processing in NPU (No. SKLSP201922) and LiaoNing Revitalization Talents Program
(No. XLYC1808027). S.J. Sun, Y.Z. Tian and Z.F. Zhang acknowledge the support of the Chinese Academy of Sciences (CAS) and
Japan Society for the Promotion of Science (JSPS) through the Bilateral Program (No. GJHZ1774). The authors also thank Prof. N.
Tsuji for valuable discussions.
References
[1] J.W. Yeh, S.K. Chen, S.J. Lin, J.Y. Gan, T.S. Chin, T.T. Shun, C.H. Tsau, S.Y. Chang, Nanostructured high-entropy alloys with multiple principal
elements: Novel alloy design concepts and outcomes, Adv. Eng. Mater. 6 (2004) 299-303.
[2] T. Yang, Y.L. Zhao, Y. Tong, Z.B. Jiao, J. Wei, J.X. Cai, X.D. Han, D. Chen, A. Hu, J.J. Kai, K. Lu, Y. Liu, C.T. Liu, Multicomponent
intermetallic nanoparticles and superb mechanical behaviors of complex alloys, Science 362 (2018) 933.
[3] S. Gorsse, D.B. Miracle, O.N. Senkov, Mapping the world of complex concentrated alloys, Acta Mater. 135 (2017) 177-187.
[4] Y. Zhang, T.T. Zuo, Z. Tang, M.C. Gao, K.A. Dahmen, P.K. Liaw, Z.P. Lu, Microstructures and properties of high-entropy alloys, Prog. Mater.
Sci. 61 (2014) 1-93.
[5] G. Qin, R. Chen, P.K. Liaw, Y. Gao, X. Li, H. Zheng, L. Wang, Y. Su, J. Guo, H. Fu, A novel face-centered-cubic high-entropy alloy strengthened
by nanoscale precipitates, Scripta Mater. 172 (2019) 51-55.
[6] S.S. Nene, S. Sinha, M. Frank, K. Liu, R.S. Mishra, B.A. McWilliams, K.C. Cho, Unexpected strength–ductility response in an annealed,
metastable, high-entropy alloy, Appl. Mater. Today 13 (2018) 198-206.
[7] B. Cantor, I.T.H. Chang, P. Knight, A.J.B. Vincent, Microstructural development in equiatomic multicomponent alloys, Mater. Sci. Eng. A 375–
377 (2004) 213-218.
[8] H. Huang, Y. Wu, J. He, H. Wang, X. Liu, K. An, W. Wu, Z. Lu, Phasetransformation ductilization of brittle highentropy alloys via
metastability engineering, Adv. Mater. 29 (2017) 1701678.
[9] Y. Bu, Z. Li, J. Liu, H. Wang, D. Raabe, W. Yang, Nonbasal slip systems enable a strong and ductile hexagonal-close-packed high-entropy
phase, Phys. Rev. Lett. 122 (2019) 075502.
[10] S. Yoshida, T. Ikeuchi, T. Bhattacharjee, Y. Bai, A. Shibata, N. Tsuji, Effect of elemental combination on friction stress and Hall-Petch
relationship in face-centered cubic high / medium entropy alloys, Acta Mater. 171 (2019) 201-215.
[11] N.A.P.K. Kumar, C. Li, K.J. Leonard, H. Bei, S.J. Zinkle, Microstructural stability and mechanical behavior of FeNiMnCr high entropy alloy
under ion irradiation, Acta Mater. 113 (2016) 230-244.
[12] H. Luo, Z. Li, A.M. Mingers, D. Raabe, Corrosion behavior of an equiatomic CoCrFeMnNi high-entropy alloy compared with 304 stainless
steel in sulfuric acid solution, Corros. Sci. 134 (2018) 131-139.
[13] Y.Z. Tian, S.J. Sun, H.R. Lin, Z.F. Zhang, Fatigue behavior of CoCrFeMnNi high-entropy alloy under fully reversed cyclic deformation, J.
Mater. Sci. Technol. 35 (2019) 334-340.
[14] B. Gludovatz, A. Hohenwarter, D. Catoor, E.H. Chang, E.P. George, R.O. Ritchie, A fracture-resistant high-entropy alloy for cryogenic
applications, Science 345 (2014) 1153-1158.
[15] F. He, Z. Wang, Q. Wu, D. Chen, T. Yang, J. Li, J. Wang, C.T. Liu, J.J. Kai, Tuning the defects in face centered cubic high entropy alloy via
temperature-dependent stacking fault energy, Scripta Mater. 155 (2018) 134-138.
[16] F. Otto, A. Dlouhý, C. Somsen, H. Bei, G. Eggeler, E.P. George, The influences of temperature and microstructure on the tensile properties of
a CoCrFeMnNi high-entropy alloy, Acta Mater. 61 (2013) 5743-5755.
[17] G. Laplanche, A. Kostka, O.M. Horst, G. Eggeler, E.P. George, Microstructure evolution and critical stress for twinning in the CrMnFeCoNi
high-entropy alloy, Acta Mater. 118 (2016) 152-163.
[18] D. Li, C. Li, T. Feng, Y. Zhang, G. Sha, J.J. Lewandowski, P.K. Liaw, Y. Zhang, High-entropy Al0.3CoCrFeNi alloy fibers with high tensile
strength and ductility at ambient and cryogenic temperatures, Acta Mater. 123 (2017) 285-294.
[19] J. Liu, X. Guo, Q. Lin, Z. He, X. An, L. Li, P.K. Liaw, X. Liao, L. Yu, J. Lin, L. Xie, J. Ren, Y. Zhang, Excellent ductility and serration feature
of metastable CoCrFeNi high-entropy alloy at extremely low temperatures, Sci. China Mater. 62 (2019) 853-863.
[20] J. Miao, C.E. Slone, T.M. Smith, C. Niu, H. Bei, M. Ghazisaeidi, G.M. Pharr, M.J. Mills, The evolution of the deformation substructure in a
Ni-Co-Cr equiatomic solid solution alloy, Acta Mater. 132 (2017) 35-48.
[21] Z. Wu, H. Bei, G.M. Pharr, E.P. George, Temperature dependence of the mechanical properties of equiatomic solid solution alloys with face-
centered cubic crystal structures, Acta Mater. 81 (2014) 428-441.
[22] B. Gludovatz, A. Hohenwarter, K.V.S. Thurston, H. Bei, Z. Wu, E.P. George, R.O. Ritchie, Exceptional damage-tolerance of a medium-
entropy alloy CrCoNi at cryogenic temperatures, Nat. Commun. 7 (2016) 10602.
[23] Y.H. Jo, S. Jung, W.M. Choi, S.S. Sohn, H.S. Kim, B.J. Lee, N.J. Kim, S. Lee, Cryogenic strength improvement by utilizing room-temperature
deformation twinning in a partially recrystallized VCrMnFeCoNi high-entropy alloy, Nat. Commun. 8 (2017) 15719.
[24] S.J. Sun, Y.Z. Tian, X.H. An, H.R. Lin, J.W. Wang, Z.F. Zhang, Ultrahigh cryogenic strength and exceptional ductility in ultrafine-grained
CoCrFeMnNi high-entropy alloy with fully recrystallized structure, Mater. Today Nano 4 (2018) 46-53.
[25] S.J. Sun, Y.Z. Tian, H.R. Lin, H.J. Yang, X.G. Dong, Y.H. Wang, Z.F. Zhang, Transition of twinning behavior in CoCrFeMnNi high entropy
alloy with grain refinement, Mater. Sci. Eng. A 712 (2018) 603-607.
[26] G. Laplanche, A. Kostka, C. Reinhart, J. Hunfeld, G. Eggeler, E.P. George, Reasons for the superior mechanical properties of medium-entropy
CrCoNi compared to high-entropy CrMnFeCoNi, Acta Mater. 128 (2017) 292-303.
[27] S.J. Sun, Y.Z. Tian, H.R. Lin, X.G. Dong, Y.H. Wang, Z.J. Zhang, Z.F. Zhang, Enhanced strength and ductility of bulk CoCrFeMnNi high
entropy alloy having fully recrystallized ultrafine-grained structure, Mater. Des. 133 (2017) 122-127.
[28] S.J. Sun, Y.Z. Tian, H.R. Lin, S. Lu, H.J. Yang, Z.F. Zhang, Modulating the prestrain history to optimize strength and ductility in CoCrFeMnNi
high-entropy alloy, Scripta Mater. 163 (2019) 111-115.
[29] Y.Z. Tian, S. Gao, L.J. Zhao, S. Lu, R. Pippan, Z.F. Zhang, N. Tsuji, Remarkable transitions of yield behavior and Lüders deformation in pure
Cu by changing grain sizes, Scripta Mater. 142 (2018) 88-91.
[30] S.J. Sun, Y.Z. Tian, H.R. Lin, H.J. Yang, X.G. Dong, Y.H. Wang, Z.F. Zhang, Achieving high ductility in the 1.7 GPa grade CoCrFeMnNi
high-entropy alloy at 77 K, Mater. Sci. Eng. A 740-741 (2019) 336-341.
[31] Y. Tong, D. Chen, B. Han, J. Wang, R. Feng, T. Yang, C. Zhao, Y.L. Zhao, W. Guo, Y. Shimizu, C.T. Liu, P.K. Liaw, K. Ino ue, Y. Nagai, A.
Materials Science and Engineering A 801 (2021) 140398, https://doi.org/10.1016/j.msea.2020.140398
12
Hu, J.J. Kai, Outstanding tensile properties of a precipitation-strengthened FeCoNiCrTi0.2 high-entropy alloy at room and cryogenic temperatures,
Acta Mater. 165 (2019) 228-240.
[32] Y. Li, Y. Lu, W. Li, M. Khedr, H. Liu, X. Jin, Hierarchical microstructure design of a bimodal grained twinning -induced plasticity steel with
excellent cryogenic mechanical properties, Acta Mater. 158 (2018) 79-94.
[33] H.R. Lin, Y.Z. Tian, S.J. Sun, H.J. Yang, Z.F. Zhang, Significant enhancement in cryogenic mechanical properties of Cu–Al alloy via minor
recrystallization, Adv. Eng. Mater. 21 (2019) 1800889.
[34] J.F. Nye, Some geometrical relations in dislocated crystals, Acta Metall. 1 (1953) 153-162.
[35] M.F. Ashby, The deformation of plastically non-homogeneous materials, Phil. Mag. 21 (1970) 399-424.
[36] D. Kuhlmann-Wilsdorf, N. Hansen, Geometrically necessary, incidental and subgrain boundaries, Scripta Metall. Mater. 25 (1991) 1557-1562.
[37] O. Muránsky, L. Balogh, M. Tran, C.J. Hamelin, J.S. Park, M.R. Daymond, On the measurement of dislocations and dislocation substructures
using EBSD and HRSD techniques, Acta Mater. 175 (2019) 297-313.
[38] P.J. Konijnenberg, S. Zaefferer, D. Raabe, Assessment of geometrically necessary dislocation levels derived by 3D EBSD, Acta Mater. 99
(2015) 402-414.
[39] P.O. Guglielmi, M. Ziehmer, E.T. Lilleodden, On a novel strain indicator based on uncorrelated misorientation angles for correlating
dislocation density to local strength, Acta Mater. 150 (2018) 195-205.
[40] R.L. Fullman, Measurement of particle sizes in opaque bodies, JOM 5 (1953) 447-452.
[41] Z.C. Luo, M.X. Huang, Revisit the role of deformation twins on the work-hardening behaviour of twinning-induced plasticity steels, Scripta
Mater. 142 (2018) 28-31.
[42] Y.T. Zhu, X.Z. Liao, X.L. Wu, Deformation twinning in nanocrystalline materials, Prog. Mater. Sci. 57 (2012) 1-62.
[43] S.J. Sun, Y.Z. Tian, H.R. Lin, X.G. Dong, Y.H. Wang, Z.J. Wang, Z.F. Zhang, Temperature dependence of the Hall–Petch relationship in
CoCrFeMnNi high-entropy alloy, J. Alloys Compd. 806 (2019) 992-998.
[44] G.I. Taylor, The mechanism of plastic deformation of crystals. Part I.—Theoretical, Proc. R. Soc. London, Ser. A 145 (1934) 362-387.
[45] I. Gutierrez-Urrutia, D. Raabe, Dislocation and twin substructure evolution during strain hardening of an Fe–22 wt.% Mn–0.6 wt.% C TWIP
steel observed by electron channeling contrast imaging, Acta Mater. 59 (2011) 6449-6462.
[46] M.A. Meyers, O. Vöhringer, V.A. Lubarda, The onset of twinning in metals: a constitutive description, Acta Mater. 49 (2001) 4025-4039.
[47] G. Laplanche, P. Gadaud, O. Horst, F. Otto, G. Eggeler, E.P. George, Temperature dependencies of the elastic moduli and thermal expansion
coefficient of an equiatomic, single-phase CoCrFeMnNi high-entropy alloy, J. Alloys Compd. 623 (2015) 348-353.
[48] S. Huang, W. Li, S. Lu, F. Tian, J. Shen, E. Holmström, L. Vitos, Temperature dependent stacking fault energy of FeCrCoNiMn high entropy
alloy, Scripta Mater. 108 (2015) 44-47.
[49] P. Sathiyamoorthi, J. Moon, J.W. Bae, P. Asghari-Rad, H.S. Kim, Superior cryogenic tensile properties of ultrafine-grained CoCrNi medium-
entropy alloy produced by high-pressure torsion and annealing, Scripta Mater. 163 (2019) 152-156.
[50] C.Y. Yu, P.W. Kao, C.P. Chang, Transition of tensile deformation behaviors in ultrafine-grained aluminum, Acta Mater. 53 (2005) 4019-4028.
[51] D.C.C. Magalhães, A.M. Kliauga, M. Ferrante, V.L. Sordi, Plastic deformation of FCC alloys at cryogenic temperature: the effect of stacking-
fault energy on microstructure and tensile behaviour, J. Mater. Sci. 52 (2017) 7466-7478.
[52] N.V. Isaev, T.V. Grigorova, O.V. Mendiuk, O.A. Davydenko, S.S. Polishchuk, V.G. Geidarov, Plastic deformation mechanisms of ultrafine-
grained copper in the temperature range of 4.2-300K, Low Temp. Phys. 42 (2016) 825-835.
[53] M.A. Meyers, K.K. Chawla, Mechanical behavior of materials, Cambridge University Press Cambridge, 2009.
[54] Y.Z. Tian, L.J. Zhao, N. Park, R. Liu, P. Zhang, Z.J. Zhang, A. Shibata, Z.F. Zhang, N. Tsuji, Revealing the deformation mechanisms of Cu–
Al alloys with high strength and good ductility, Acta Mater. 110 (2016) 61-72.
[55] J. Jain, W.J. Poole, C.W. Sinclair, The deformation behaviour of the magnesium alloy AZ80 at 77 and 293K, Mater. Sci. Eng. A 547 (2012)
128-137.
[56] M. Knezevic, A. Levinson, R. Harris, R.K. Mishra, R.D. Doherty, S.R. Kalidindi, Deformation twinning in AZ31: Influence on strain
hardening and texture evolution, Acta Mater. 58 (2010) 6230-6242.
[57] M. Zha, Y. Li, R.H. Mathiesen, R. Bjørge, H.J. Roven, Microstructure evolution and mechanical behavior of a binary Al–7Mg alloy processed
by equal-channel angular pressing, Acta Mater. 84 (2015) 42-54.
[58] Y. Wei, Y. Li, L. Zhu, Y. Liu, X. Lei, G. Wang, Y. Wu, Z. Mi, J. Liu, H. Wang, H. Gao, Evading the strength–ductility trade-off dilemma in
steel through gradient hierarchical nanotwins, Nat. Commun. 5 (2014) 3580.
[59] O. Bouaziz, N. Guelton, Modelling of TWIP effect on work-hardening, Mater. Sci. Eng. A 319-321 (2001) 246-249.
[60] L. Lu, X. Chen, X. Huang, K. Lu, Revealing the maximum strength in nanotwinned copper, Science 323 (2009) 607-610.
[61] D. Li, Y. Zhang, The ultrahigh charpy impact toughness of forged AlxCoCrFeNi high entropy alloys at room and cryogenic temperatures,
Intermetallics 70 (2016) 24-28.
[62] L.B. Chen, R. Wei, K. Tang, J. Zhang, F. Jiang, J. Sun, Ductile-brittle transition of carbon alloyed Fe40Mn40Co10Cr10 high entropy alloys,
Mater. Lett. 236 (2019) 416-419.
[63] Z. Wu, H. Bei, Microstructures and mechanical properties of compositionally complex Co-free FeNiMnCr18 FCC solid solution alloy, Mater.
Sci. Eng. A 640 (2015) 217-224.
[64] Z. Wu, C.M. Parish, H. Bei, Nano-twin mediated plasticity in carbon-containing FeNiCoCrMn high entropy alloys, J. Alloys Compd. 647
(2015) 815-822.
... An example is shown in the dotted rectangle in the corresponding figure. The KAM maps indicate that the fraction of high values of local lattice misorientations, linked to the local plastic strain, increases with the pre-strain level, as reported in previous studies [52]. Besides, the KAM maps clearly emphasize the strain heterogeneity in the inter-and intragranular regions at the grain scale. ...
... A 33-μm diameter area was analysed inside a raster square of 125 × 125 μm 2 . The analysed elements were 1 H, 2 H, 12 C, 52 Cr, 56 Fe and 58 Ni. The results are given in the form of the measured counts per second as a function of abrasion time. ...
Article
The present study aims at investigating the interactions between hydrogen and deformation-induced defects in a 316L austenitic stainless steel (SS). The studied material was pre-strained up to different levels. The resulting microstructures were characterized by optical and electron microscopy and by X-ray diffraction. Dislocation densities were systematically quantified showing a more than 50-fold increase between the unstrained and the most pre-strained sample. Cathodic charging was employed to introduce deuterium into the material. By means of Thermal Desorption Spectrometry (TDS) and Secondary Ions Mass Spectrometry (SIMS) analyses, deuterium energetic and spatial distributions were evaluated both just after charging and after an additional isothermal aging treatment. Thanks to aging treatments, three different energetic states of deuterium in the 316L SS were evidenced: interstitial 2H, low-energy trapped 2H and high-energy trapped 2H. Probing the low-energy trapping was challenging – probably due to the proximity between the migration enthalpy and the associated detrapping energy – but aging highlighted it. This low-energy contribution, displayed in TDS results, increased with the pre-strain level and the dislocation density, and was associated with dislocations elastic fields. Contrary to our expectations, the high-energy trapping could not be correlated directly neither to the pre-strain level nor to the dislocation density. Trapping at vacancies was suggested.
... The interaction of dislocations with GBs including their absorption or re-emission from the GBs is enhanced due to a large fraction of GBs [20,[25][26][27]. The dislocation-mediated plasticity in nc metals is often characterized by global texture evolution [28,29] due to the rotation of individual grains and their interaction with neighbors as revealed by kernel average misorientation (KAM) [30] and intragranular misorientation (mis2mean) [12,31] local analysis. ...
Article
Nanomechanical on-chip testing is combined with nanoscale in situ digital image correlation and automated crystal orientation mapping in TEM to deliver novel statistically representative quantitative data about the deformation mechanisms in nanocrystalline aluminum films. The films are very ductile, with a rare stable multiple necking process with local strains reaching up to 0.45 and macroscopic elongation up to 0.17. The strain fields with resolution below 100 nm are related to the underlying microstructure and crystallographic orientation maps. This reveals nanoscopic shear bands forming preferentially along GB with high misorientations, tilted at +/− 45° with respect to loading direction. The analysis of these data prove that the strong strain delocalization process is promoted by GB migration and grain rotation, leading to large strain rate sensitivity. The distribution of misorientation angles between grains evolve during deformation. The GBs with misorientation between 20° and 40°, which are the GBs with highest energy, involve the largest strains.
... High KAM values were observed near the grain boundaries in both samples, which agrees with the accumulation of geometrically necessary dislocations (GNDs) near the grain boundaries during the plastic deformation. The relationship between GNDs and KAM is given by [18]: ...
Article
Full-text available
High-strength low-alloy steels are known for their superior mechanical properties and these desired properties are achieved through controlled hot rolling. The cooling rate is one of the most important factors in controlled hot rolling, which plays a major role in the final microstructure and mechanical properties. In this work, the effect of the cooling rate on the final microstructure and mechanical properties of high Ti-V HSLA microalloyed steel was investigated using the Gleeble 1500™ thermomechanical processing simulator. The samples were cooled to room temperature after final pass deformation using two different cooling rates to simulate the 16 and 30 mm thick plate rolling patterns. Polygonal ferrite was found to be a dominant phase in the final microstructures; SEM-EBSD IPF maps also showed grains with random orientation distribution. Yield strength and ultimate tensile stress increased by approximately 14% and 10%, respectively, with increasing cooling rate.
... KAM map is shown before and after the peening for different materials. An increase in plastic strain led to higher KAM values and geometricallynecessary dislocations (GNDs) [37]. Fig. 13(a, b, c) showed the lower KAM values for as-received materials at the grain boundary and within the grains. ...
... The multi-principal element design concept breaks the traditional theory that the alloy structure will become complex with the increase in principal elements [2]. Many experimental studies have shown that reasonable control of the types and ratios of elements can make HEAs form a single solid-solution system [1][2][3][4][5]. ...
Article
Full-text available
The gradient nano-grained (GNG) structures achieve a combination of high strength and ductility compared with homogeneous nano-grained structures. In order to study the mechanical behavior of GNG with extremely small grain size, the systematic tension, compression and cyclic deformation behavior of gradient and homogeneous nano-grained CoCrFeMnNi high-entropy alloy (G-HEA and H-HEA) are studied by the molecular dynamics (MD) simulation. The simulation results show that G-HEA and H-HEA have multiple plastic deformation mechanisms during uniaxial and cyclic loading, such as dislocation slip, deformation twinning, phase transformation from face-centered cubic (FCC) to hexagonal close-packed (HCP), and grain boundary plastic behavior. During uniaxial loading, the heterogeneous deformation makes the yield stress of G-HEA higher than that calculated by the rule of mixture (ROM), indicating an extra strengthening in GNG material. However, the extra strengthening of G-HEA is weakened with increased strain due to the dislocation without piled up in front of the grain boundary and relaxed heterogeneous deformation. During cyclic loading, the sessile dislocations promote the reverse movement of dislocations, resulting in the Bauschinger effect. However, the sessile dislocation density of G-HEA is close to that calculated by the ROM, resulting in the Bauschinger effect of G-HEA close to that calculated by the ROM. Therefore, as an alternative perspective, the MD simulation confirms that dislocation pinning in front of grain boundaries is an important source of extra strengthening for gradient structured materials as revealed by experiments. The results extend the understanding of the deformation mechanism of CoCrFeMnNi GNG HEA with extremely small grain size under tension, compression and cyclic loading.
... It is found that the low-temperature strength of previous alloyed materials is difficult to exceed 2. GPa. Even the new medium-and highentropy alloys [39][40][41], although they can achieve the synergy of strength and ductility, are still difficult to achieve ultra-high strength at LNT. For the CRA600 sample, the yield strength not only exceeded 2.0 GPa, but was close to 2.5 GPa. ...
Article
Full-text available
To obtain ultra-high strength with guaranteed ductility, stable nanolamellar austenite packed with high densities of dislocations and nanotwins is inserted into nanolamellar martensite. An ultra-high yield strengths of 1.6–2.2 GPa at room temperature and 1.94–2.5 GPa at liquid nitrogen temperature were obtained in the steel, which is combined with good uniform elongation of 4–9%. The designed steel is shaped with the combination of solid solution strengthening, lamellar boundary strengthening and dislocation strengthening. The nanostructure is stabilized by C distribution, and in this way the stress–strain carrying capacity of martensite and austenite is increased, thus ensuring ultra-high strength. In addition, the generation of TRIP effect at low temperatures ensures the realization of strain-hardening rate and uniform elongation at ultra-high strength. Therefore, the nanolamellar martenite-austenite microstructure and strong carbon partitioning through martensite and austenite is considered as a fundamental guarantee for the ultra-high strength and toughness. Our results provide an optional strategy for the development of new ultra-strong alloys for low-temperature applications.
Article
Full-text available
This study fabricated four kinds of CoCrMnFeNiX0.1 (X = Al, Cu, Mo, Ti) high-entropy alloys through rolling with subsequent thermal treatment, to investigate the alloying effects on the microstructural evolution, mechanical properties, and electrochemical corrosion performance. High-density parallel twinning bands were observed in the as-rolled CoCrMnFeNiX0.1 alloys, excluding the Cu0.1 variant, signifying that plastic deformation was dominated by twinning. However, detwinning and thickening of twin layers was observed after annealing. This effect was more pronounced in the rolling direction-normal direction (RD-ND) plane due to the stress state induced by rolling. The introduction of Cu altered the deformation mechanism to a combination of twinning and slipping, resulting in fewer twinning bands. Notably, annealing induced the formation of abundant bamboo-shaped twinning bands (width: ~ 10 nm) in the Cu0.1 alloy, which were correlated with nano precipitated particles. Additionally, the incorporation of Mo or Al elements activated more twinning directions, thereby enhancing plasticity. For the Ti0.1 alloy, the primary factors influencing the altered tensile and electrochemical performances were the formation and growth of few precipitated particles during annealing. Consequently, the strengthening effect can be ranked in the order of Mo > Ti > Al > Cu. Therefore, the addition of Al and Ti is recommended for improving the anticorrosive property and thermal stability of recrystallized grains. Graphical Abstract
Article
Full-text available
Single-phase equiatomic CoCrFeMnNi high-entropy alloy (HEA) specimens with three different grain sizes (0.65 μm, 2.1 μm and 105 μm) were processed by cold rolling and annealing treatment. Tensile properties were investigated over a broad temperature range from 77 K to 873 K. Superior strength-ductility balance can be achieved by refining grain size and decreasing temperature. The Hall-Petch relationship was well fitted, and the fitting parameters have a negative relation with temperature. The thermal and athermal contributions to the yield strength were revealed. This work has unveiled the superior mechanical property of HEAs at cryogenic temperatures.
Article
Full-text available
The accumulation of the dislocations and development of dislocation structures in plastically deformed Ni201 is examined using dedicated analyses of Electron Back-Scatter Diffraction (EBSD) acquired orientation maps, and High-Resolution Synchrotron Diffraction (HRSD) acquired patterns. The results show that the minimum detectable microstructure-averaged (bulk) total dislocation density (r T) measured via HRSD is approximately 1E13 m À2 , while the minimum GND density (r G) measured via EBSD is approximately 2E12 m À2 e the EBSD technique being more sensitive at low plastic strain. This highlights complementarity of the two techniques when attempting to quantify amount of plastic deformation (damage) in a material via a measurement of present dislocations and their structures. Furthermore, a relationship between EBSD-measured r G and the size of HRSD-measured Coherently Scattering Domains (CSDs) has been mathematically derived e this allows for an estimation of the size of CSDs from EBSD-acquired orientation maps, and conversely an estimation of r G from HRSD-measured size of CSDs. The measured evolution of r T , and r G is compared with plasticity theory models e the current results suggest that Ashby's single-slip model underestimates the amount of GNDs (r G), while Taylor's model is correctly predicting the total amount of dislocation (r T) present in the material as a function of imparted plastic strain.
Article
Full-text available
In this study, we report the effect of elemental combinations on the friction stress and Hall-Petch relationship in medium entropy alloys (MEAs) and high entropy alloys (HEAs) which are defined as the alloys composed of four or less and five or more principal elements, respectively, with (near-) equi-atomic concentrations. The MEAs (CoCrFeNi, CoCrNi, etc.), which are subsystems of equi-atomic CoCrFeMnNi HEA, were highly deformed by high-pressure torsion (HPT) and subsequently annealed at different temperatures. The specimens with fully-recrystallized microstructures of FCC single-phase with various mean grain sizes down to sub-micrometer scale were obtained. Subsequently, tensile tests were performed at room temperature to obtain precise Hall-Petch relationships and friction stresses of the materials. Co20(CrNi)80 was successfully predicted as the alloy showing the highest strength among the MEAs by the modified Labusch model (so-called mean field Labusch model) for solution hardening. Experimental values of the friction stresses were found to fit with the model very well, indicating that the strength of the alloys was closely related to entirely distorted crystal lattice acting as high-density obstacles for dislocation motion in the alloys. At the same time, values of the average lattice distortion in the alloys was found comparable to those in some dilute alloys, although “severe” lattice distortion had been believed as a reason for the higher strength than dilute systems. Finally, a strengthening mechanism by element-element interaction was proposed as an additional mechanism to enhance the strength in FCC HEAs and MEAs.
Article
Full-text available
CoCrFeMnNi high-entropy alloy (HEA) exhibits excellent combination of strength and ductility, but low yield strength. In order to ameliorate the mechanical properties, prestrain was applied in this work. The HEA prestrained at 77 K possesses higher yield strength and uniform elongation than the HEA prestrained at 293 K, indicating that the trade-off relationship between strength and ductility can be broken by modulating the prestrain history. Furthermore, the yield point phenomenon was disappeared after prestrained at 77 K. This can be related to the density and distribution of dislocations as imposed in the prestrain process at 293 K and 77 K.
Article
Full-text available
Seldom could metals and alloys maintain excellent properties in cryogenic condition, such as the ductility, owing to the restrained dislocation motion. However, a face-centered-cubic (FCC) CoCrFeNi high-entropy alloy (HEA) with great ductility is investigated under the cryogenic environment. The tensile strength of this alloy can reach a maximum at 1,251±10 MPa, and the strain to failure can stay at as large as 62% at the liquid helium temperature. We ascribe the high strength and ductility to the low stacking fault energy at extremely low temperatures, which facilitates the activation of deformation twinning. Moreover, the FCC→HCP (hexagonal close-packed) transition and serration lead to the sudden decline of ductility below 77 K. The dynamical modeling and analysis of serrations at 4.2 and 20 K verify the unstable state due to the FCC→HCP transition. The deformation twinning together with phase transformation at liquid helium temperature produces an adequate strain-hardening rate that sustains the stable plastic flow at high stresses, resulting in the serration feature.
Article
A new single-phase face-centered-cubic (FCC) Co9Cr7Cu36Mn25Ni23 [atomic percent, similar hereinafter] high-entropy alloy (HEA) was prepared by arc melting. A uniform distribution of nanometer-sized precipitates was achieved. The tensile yield strength, ultimate tensile strength, and elongation were 401 MPa, 700 MPa, and 36%, respectively. The energy-dispersive spectrometer results showed that the nano-precipitates were rich in Co and Cr elements. Moreover, the crystal-forming behavior and the nanoscale-precipitates-forming mechanism were revealed.
Article
We study the kinetics of the substructure evolution and its correspondence to the strain hardening evolution of an Fe-22 wt.% Mn-0.6 wt.% C TWIP steel during tensile deformation by means of electron channeling contrast imaging (ECCI) combined with electron backscatter diffraction (EBSD). The contribution of twin and dislocation substructures to strain hardening is evaluated in terms of a dislocation mean free path approach involving several microstructure parameters, such as the characteristic average twin spacing and the dislocation substructure size. The analysis reveals that at the early stages of deformation (strain below 0.1 true strain) the dislocation substructure provides a high strain hardening rate with hardening coefficients of about G/40 (G is the shear modulus). At intermediate strains (below 0.3 true strain), the dislocation mean free path refinement due to deformation twinning results in a high strain rate with a hardening coefficient of about G/30. Finally, at high strains (above 0.4 true strain), the limited further refinement of the dislocation and twin substructures reduces the capability for trapping more dislocations inside the microstructure and, hence, the strain hardening decreases. Grains forming dislocation cells develop a self-organized and dynamically refined dislocation cell structure which follows the similitude principle but with a smaller similitude constant than that found in medium to high stacking fault energy alloys. We attribute this difference to the influence of the stacking fault energy on the mechanism of cell formation.
Article
Existing alternatives for the calculation of geometrically necessary dislocation (GND) densities from orientation fields are discussed. Importantly, we highlight the role of reference frames and consider different sources of error. A well-controlled micro cantilever bending experiment on a copper bicrystal has been analyzed by 3-dimensional electron back scatter diffraction (3D EBSD). The GND density is determined experimentally by two different approaches and assessed theoretically, assuming a homogeneous bending of the cantilever. Experiment and theory agree very well. It is further shown that the deformation is accommodated mainly by GNDs, which carry and store lattice rotation, and not (only) by mobile dislocations that leave a crystal portion inspected, without lattice rotations. A detailed GND analysis reveals a local density minimum close to the grain boundary and a distinct difference in edge to screw ratios for both grains.
Article
Linear defects, referred to as dislocations, determine the strength, formability, and toughness of crystalline metallic alloys. The associated deformation mechanisms are well understood for traditional metallic materials consisting of one or two prevalent matrix elements such as steels or aluminum alloys. In the recently developed high-entropy alloys (HEAs) containing multiple principal elements, the relationship between dislocations and the mechanical behavior is less understood. Particularly HEAs with a hexagonal close-packed (hcp) structure can suffer from intrinsic brittleness due to their insufficient number of slip systems. Here we report on the surprisingly high formability of a novel high-entropy phase with hcp structure. Through in situ tensile testing and postmortem characterization by transmission electron microscopy we reveal that the hcp phase in a dual-phase HEA (Fe 50 Mn 30 Co 10 Cr 10 , at. %) activates three types of dislocations, i.e., hai, hci, and hc þ ai. Specifically, nonbasal hc þ ai dislocations occupy a high line fraction of ∼31% allowing for frequent double cross slip which explains the high deformability of this high-entropy phase. The hcp structure has a c=a ratio of 1.616, i.e., below the ideal value of 1.633. This modest change in the structure parameters promotes nonbasal hc þ ai slip, suggesting that ductile HEAs with hcp structure can be designed by shifting the c=a ratio into regimes where nonbasal slip systems are activated. This simple alloy design principle is particularly suited for HEAs due to their characteristic massive solid solution content which readily allows tuning the c=a ratio of hcp phases into regimes promoting nonbasal slip activation.
Article
Ultrafine-grained materials with nanotwins are expected to produce a remarkable combination of strength and ductility. In the present study, ultrafine-grained CoCrNi medium-entropy alloy with nanotwins is fabricated by high-pressure torsion followed by annealing; and investigated for cryogenic tensile properties. The alloy exhibits superior cryogenic tensile properties with a tensile strength of ~2 GPa and tensile strain of ~27%. The cryogenic tensile strength of ultrafine-grained sample increased by 67% as compared to the cryogenic tensile strength of coarse-grained sample due to fine grain size, annealing nanotwins, residual dislocation density, and strong temperature dependence of yield strength.