Journal Pre-Proof: Materials Science & Engineering A

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Journal Pre-proof

Martensitic transformation and mechanical behavior of a medium-entropy alloy

C. Wang, K.F. Lin, Y.L. Zhao, T. Yang, T.L. Zhang, W.H. Liu, C.H. Hsueh, H.C. Lin,
J.J. Kai, C.T. Liu

PII: S0921-5093(20)30452-4
DOI: https://doi.org/10.1016/j.msea.2020.139371
Reference: MSA 139371

To appear in: Materials Science & Engineering A

Received Date: 11 February 2020


Revised Date: 6 April 2020
Accepted Date: 7 April 2020

Please cite this article as: C. Wang, K.F. Lin, Y.L. Zhao, T. Yang, T.L. Zhang, W.H. Liu, C.H. Hsueh,
H.C. Lin, J.J. Kai, C.T. Liu, Martensitic transformation and mechanical behavior of a medium-entropy
alloy, Materials Science & Engineering A (2020), doi: https://doi.org/10.1016/j.msea.2020.139371.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


Credit Author Statement

C. Wang: Methodology, Data collection and analysis, Manuscript preparation


K. F. Lin: Data collection and analysis
Y. L. Zhao: Data collection and analysis
T. Yang: Data collection
T. L. Zhang: Methodology
W. H. Liu: Methodology
C. H. Hsueh: Writing-Reviewing, Supervision
H. C. Lin: Supervision
J. J. Kai: Supervision
C. T. Liu*: Writing-Reviewing, Supervision
Martensitic Transformation and Mechanical Behavior of a

Medium-Entropy Alloy

C. Wanga, K. F. Linb, Y. L. Zhaoc, T. Yangc, T. L. Zhanga, W. H. Liua, C. H. Hsuehb,


H. C. Linb, J. J. Kaic, C. T. Liua*
a
Department of Materials Science and Engineering, City University of Hong Kong, Hong

Kong, China
b
Department of Materials Science and Engineering, National Taiwan University, Taipei 10617,

Taiwan
c
Department of Mechanical Engineering, City University of Hong Kong, Hong Kong, China

Abstract
Diffusionless martensitic transformation (MT) exerts one of the most significant
influences on the mechanical properties of alloys. However, the application of
martensitic transformation to improve mechanical performance was seldom involved
in the manufacture of high-entropy alloys (HEAs) and medium-entropy (MEAs)
alloys. In this work, an innovative non-equiatomic MEA, Fe42Co42Cr16, was proposed
with incorporation of martensitic transformation during water quenching and plastic
deformation. Water quenching for the alloy in the high-temperature single-phase
region produced a partial MT; i.e., transformation of -FCC austenite phase into an
-HCP martensite phase, responsible for the coexistence of and phases in the
dual-phase (DP) alloy. Another triple-phase (TP) alloy, including -FCC austenite,
-HCP martensite and B2-BCC precipitates, was obtained by quenching the alloy in
+ 2 phase region. Owing to the low intrinsic stacking fault energy ( ), both DP
( =10.9 mJ/m2) and TP ( =12.2 mJ/m2) alloys involved the complete polymorphic
MT process during plastic deformation; i.e., the transformation of the -FCC
austenite phase into the -BCT martensite phase with an intermediate -HCP
martensite phase. Due to the transformation-induced plasticity effect and precipitation
strengthening, the produced TP alloy exhibited a yield strength above 1 GPa with a

1
total elongation of as high as 25%.
Keywords: Medium-entropy alloy; Martensitic transformation; Microstructure;
Mechanical properties

1. Introduction
High-entropy alloys (HEAs) with multicomponent (without base element) have
attracted massive attention from a wide range of investigations over the past decade
[1-6] due to their excellent and unique properties, such as superior ductility [7, 8],
impressive cryogenic fracture toughness [9, 10], high corrosion [11, 12] and wear
resistances [13, 14], novel sluggish diffusion behavior [15, 16] and promising
irradiation stability [17, 18]. Recent investigations motivated a series of technological
progress in HEA materials, contributing to the development of non-equiatomic alloys
and alloys with three or four components (medium-entropy alloy, MEA) [19-21]. The
HEAs or MEAs with face-centered cubic (FCC) single phase structure have aroused
an extensive attention in investigations, primarily because of the remarkable ductility
of these single-FCC HEAs. However, deficiency in their yield strength puts a limit on
their engineering applications [22-24]. It has been widely confirmed that precipitation
strengthening [25-27] or martensitic transformation [28-30] could provide an effective
method to address the strength-ductility trade-off.
Recently, Li et al. [28] reported dual-phase FeMnCoCr HEAs induced by phase
transformation resulted in high tensile strength and superior ductility. The martensitic
transformation was achieved by regulating the content ratio between Fe and Mn,
responsible for the change in the stacking fault energy (SFE) of the alloys. The
deformation mechanism exhibits an evident difference depending on the SFE in the
FCC-alloys. For alloys with a high SFE (> 45 mJ/m2), dislocations are rarely
dissociated, and the deformation process is mainly dependent on the perfect
dislocation glide to accommodate plasticity [31]. The alloys with an intermediate SFE
(20-45 mJ/m2) is deformed by a twinning process [32, 33]. The deformation twins are
created by Shockley partial dislocations dissociated from the perfect dislocation

2
slipping, producing twinning induced plasticity (TWIP). In the case with low SFE (<
20 mJ/m2), the partial dislocations or half partial dislocations slipping are inclined to
the martensitic phase transformation, generating a phase transformation induced
plasticity (TRIP) [34, 35].
Based on these previous findings, martensitic phase transformation can be
achieved by adjusting the SEF in the HEAs through tailoring the composition of the
alloys. In this work, according to the influence of empirical element on SFEs in
austenite stainless steels [36], and considering the unique cocktail-effect in the HEAs
(Fe, Co, Cr with FCC/BCC, HCP and BCC structures, respectively) [37], a
non-equiatomic MEA; i.e., Fe42Co42Cr16, was developed through the manipulation of
martensitic phase transformations by means of composition tuning. The synergy effect
of phase transformation and precipitation hardening were discussed here. Besides, the
phase stability, microstructures and mechanical properties of this MEA were
systematically investigated and discussed.

2. Materials and Methods


2.1 Processing of Fe42Co42Cr16 medium-entropy alloy
Alloy ingots with the nominal composition of Fe42Co42Cr16 were fabricated by
arc melting under pure Ar atmosphere. Each constitutive raw material had a purity
above 99.9 wt%. The chamber of arc melting was first evacuated to a level of
~8 × 10 Pa and then filled with Ar gas to a pressure of ~3 × 10 Pa. Before the
melting of pre-weighted raw metals, a small amount of Ti was sacrificed to remove
residual oxygen and nitrogen in the chamber. Each ingot was flipped and re-melted
for at least five times to guarantee a homogeneous composition, before pouring into a
rectangular oxygen-free copper mold.
The fabricated ingots (50 × 10 × 4.9 mm ) were then annealed in atmosphere at
1200 for 2 h and water quenched (WQ) for homogenization. The ingots were
successfully cold rolled (CR) along the longitudinal direction with a constant
thickness reduction of 0.2 mm for each rolling. The final thickness of the alloy sheet

3
was ~1.6 mm with a total thickness reduction of ~67%. The cold-rolled sheets were
then annealed in atmosphere at 700 , 750 and 800 for 1 h, respectively, and
quenched in water to produce a recrystallization.
The dog-bone-shaped tensile test specimens with a gauge length of 12.5 mm
were manufactured from recrystallized alloy sheets by an electric discharge machine.
The tensile specimens were then ground with abrasive paper (800, 2000, 4000 in
sequence, 3M WetordryTM) on each side to eliminate the oxidation layer on the
surface generated during annealing and wire-electrode cutting. The gauge portion of
final specimens had a thickness and width of ~1.5 mm and ~3.4 mm, respectively.
Uniaxial tensile tests were performed by a universal test machine (MTS Criterion
42.503 Test System) at a strain rate of 10–3 s–1. The elongation of the gauge portion
was monitored by non-contact laser extensometer (MTS LX 500) with two
removeable laser marks stuck on the specimens.

2.2. Characterizations
The structure of the samples was analyzed by X-ray diffraction (XRD, Rigaku
SmartLab) using a Cu-Kα radiation with the scanning angle 2θ from 20° to 100° and a
monochromator was used to remove the fluoresce signals excited from the samples.
The metallographic phases were examined by scanning electron microscopy (SEM,
FEI Quanta 450) equipped with electron backscatter diffraction (EBSD) and
energy-dispersive X-ray spectroscopy (EDS). The samples for XRD, SEM and EBSD
were electropolished using an electrolyte containing 25% nitric acid and 75% methyl
alcohol at temperature of -30 with an applied voltage of ~30V. The microstructures
were then analyzed by a transmission electron microscopy (TEM, JEOL 2100F) with
an EDS detector and a scanning transmission electron microscopy (STEM) with a
high angle annular dark field (HAADF) detector. For the TEM samples preparation,
thin foils were firstly grounded down to the thickness of about 100 μm and discs with
diameter of 3 mm were punched out from the foils. The discs were then grounded to a
thickness of about 40 μm and dimpled (Gatan 656 Dimple Grinder) down to 10 μm.

4
The dimpled discs were subsequently thinned by ion miller (Gatan PIPS ) at -30
with the incident ion beam smaller than 6° and the diameter of the final broken holes
was about 50 μm.

3. Results and Discussion


3.1 Microstructures and crystal structures
The XRD patterns of the alloy with recrystallization at 700 , 750 and 800
are shown in Fig. 1. The structures of the alloys changed from dual-phase ( -FCC and
-HCP) to triple-phase ( -FCC, -HCP and B2-BCC) with a decrease in the
recrystallization temperature from 800 to 700 . Fig. 2 shows the microstructures
of Fe42Co42Cr16 MEAs after cold rolling and annealing. The backscattered electron
images clearly reveal that the DP microstructure was featured with equiaxed grains
with laminate plates spanning invariably over an entire grain in the DP specimens.
When the recrystallization temperature decreased to 700 , as shown in Fig. 2(c), the
third phase embedded in the matrix could be evidently observed, consistent with the
corresponding XRD result shown in Fig. 1. The EDS mapping (Fig. 2(d)) shows the
homogeneous distribution of elements without elemental modulations in the DP
specimen annealed at 750 , and the result is the same for the specimen annealed at
800 . Therefore, the diffusionless martensitic transformation during water quenching
is the only governing mechanism for the formation of the dual-phase in the specimens,
and the B2-BCC precipitates in the TP specimen with recrystallization at 700 will
be discussed in the following section.

5
Fig. 1. XRD patterns of Fe42Co42Cr16 MEAs at different annealing temperatures after
cold rolling.

Fig. 2. Backscattered electron images of the metallographic microstructures of


specimens annealing at (a) 800 , (b) 750 and (c) 700 . (d) Elements distribution
of cold-rolled specimen annealed at 750 .

3.1.1 Microstructure of dual-phase MEAs


We first analyze the microstructure of the DP MEA including -FCC and
-HCP phases. Figure 3(a) and (b) show the bright-field images and selected area
diffraction pattern (SADP), respectively, and reveal the co-existence of -FCC and
-HCP phases for the water-quenched specimen recrystallized at 750 . According to
6
the SADP along the -〈110〉 zone axis from the circled area in Fig. 2(a), two
phases within the -{111} habit planes could be identified, and the orientation
relationship between the phases was confirmed to be -〈1120〉// -〈110〉,
- 0002 // - 111 , which are consistent with the typical
Bogers-Burgers-Olson-Cohen (BBOC) relationship [38-40]. The lattice constants of

the -FCC and -HCP are = 3.6 Å and = 2.5 Å & ! = 4.2 Å, respectively.

Figure 3(c-e) exhibit the dark-field TEM images highlighted by - 111 , "- 0111
and #- 0111 , respectively. Evidently, the hierarchical nanolaminate structure is
featured with -HCP nanolaminates in the -FCC blocks, and -FCC nanolaminates
in -HCP blocks. According to Fig. 3(c), The fine γ nanolaminates had an average
thickness of ~13 nm, while the ε nanolaminates displayed a smaller average thickness
of ~8 nm as shown in Fig. 3(d).
As shown in Fig. 4(b), in order to investigate the relationship among blocks,

" blocks and # nanolaminates in terms of their atomic-scaled orientation, the


region marked by a circle in Fig. 4(a) was further examined with high-resolution TEM
(HRTEM). According to the fast Fourier transform (FFT) result (inset in Fig. 4(b)),
the atomic structure followed the standard BBOC relationship; i.e., -〈1120〉//
-〈110〉. This orientation relationship remained unchanged in micro-scaled ( and
blocks), nano-scaled (nanolaminates) and atomic-scaled structures, contributing to the
interlaced hierarchical structure in the Fe42Co42Cr16 DP MEA. The microstructure of
the water quenched specimen recrystallized at 800 was the same as that at 750 .

7
Fig. 3. (a) Bright-field TEM image of the specimen annealed at 750 . (b) SADP for
the region marked by circle in (a) along 〈111〉 zone axis. (c-e) Dark-field TEM
images for the three corresponding diffraction spots marked in (b).

Fig. 4. (a) Bright-field TEM image with hierarchical nanostructure in the Fe42Co42Cr16
MEA. (b) HRTEM image with the corresponding FFT pattern (inset) of a " block
and # nanolaminate within the block.

3.1.2 Microstructure of triple-phase MEA


The bright-field TEM image in Fig. 5(a) shows an overview of the
microstructure of the Fe42Co42Cr16 MEA recrystallized at 700 after cold rolling. In
addition to the coexisting phases of -FCC and -HCP phases, the third phase,
typically termed as incoherent B2-BCC precipitate, was visibly embedded in the
matrix. According to SADP (Fig. 5(b)), the lattice constant of the B2-BCC phase was
&
$% = 4.1 Å. The HRTEM image along -〈001〉 and the corresponding FFT pattern
of the B2-BCC, shown in Fig. 5(c), suggesting the ordered super lattice atomic
8
structure of this phase. Similar to the DP MEAs (Fig. 3), the orientation relationship
of -FCC and -HCP phases is -〈1120〉// -〈110〉, - 0002 // - 111 , as shown
in Fig. 5(d) and (e). On the contrary, no clear # blocks and nanolaminates exists
accompanied with the - 111 twin occurs in the three phases microstructure as
shown in Fig. 5(a) and (d), which produces a significant difference compared with the
DP specimen. The other discrepancies include a strong reduction in the density of
nanolaminates and its increased average thickness of ~21 nm. The precipitate can be
visualized in the STEM image (Fig. 6(a)). According to the corresponding EDS
mappings for the elements distribution shown in Fig. 6(c-d), the compositions of the
matrix ( -FCC and -HCP) and precipitates (B2-BCC) were obtained as shown in
Table 1. It indicates that the constituents of this B2-BCC phase are mainly Fe and Co
with a small amount of Cr solute. Apparently, besides the third B2-BCC phase
precipitated in the matrix, the TP specimen was composed of the -FCC and
martensitic -HCP phases, whose orientation yielded to the BBOC relationship (Fig.
5(d)).

Fig. 5. (a) Bright-field TEM image Fe42Co42Cr16 MEA recrystallized at 700 ,


suggesting the existence of -FCC, -HCP, B2-BCC and -twin after water
quenching. (b) SADP of the B2-BCC precipitate in (a). (c) HRTEM image and the
9
corresponding FFT pattern (inset) of B2-BCC phase. (d) SADP of the circled region
in (a) along -〈110〉 zone axis. (e) HRTEM image and the corresponding FFT
pattern (inset) of the rectangle region marked in the (a).

Fig. 6. (a) STEM image of the water quenched specimen recrystallized at


700 corresponding to the composition segregation in the B2-BCC phase. (b-d)
EDS mappings for the element distribution of Fe, Co and Cr.

To evaluate the phase stability of the triple-phase Fe42Co42Cr16 MEA, a


simulation was performed using the Thermo-Calc simulation software based on the
CALPHAD method [41], which has been proven as an effective tool for
multicomponent alloy systems [42-44]. Figure 7 shows the calculated equilibrium
phase fractions as a function of temperature with TTNI8 database. According to the
calculation, the single -FCC region spans from 765 to the liquidus, and the
coexistence of -FCC and BCC covers a narrow temperature range from 703 to
765 . In addition, the content of Fe, Co and Cr in the BCC at 705 based on the
calculation was confirmed to be 50.04 at.%, 43.94 at.% and 6.02 at.%, respectively.
Therefore, the general agreement between the calculation results and the experimental
10
data verified the B2-BCC precipitation during recrystallization at 700 after cold
rolling. Certainly, the thermodynamic equilibrium calculation failed to predict the
unstable martensitic -HCP phase transformed from -FCC.

Table 1. Compositions of matrix and precipitation in triple-phase MEA

Elements (at. %) Matrix Precipitations

Fe 40.52 ± 0.08 54.48 ± 0.09

Co 41.86 ± 0.16 43.12 ± 0.11

Cr 17.62 ± 0.05 2.40 ± 0.08

Fig. 7. The mole fraction of equilibrium phases in Fe42Co42Cr16 MEA calculated from
Thermo-Calc software.

3.2 Microstructure difference between dual-phase and triple-phase MEAs


In order to explain the microstructure difference between the DP specimens and
the matrix of TP specimen, both the stacking fault energy and temperature gradient
need to be considered. The stacking fault energy is related to the interruption of the
regular ABCABC… stacking sequence in the -FCC structure and determines the
transformation from -FCC to -HCP. This strain-induced phase transformation is

11
governed by the following dissociation of planar dislocations on the - 111 basal
plane:
" " "
(110) = (211) + (121), (1)
# * *

"
which means a perfect (110) dislocation dissociates into two Shockley partial
#

" "
dislocations with the Burges vectors of (211) and (121), respectively [45]. The
* *

dissociation on every second - 111 plane produced the intrinsic stacking fault; i.e.,
a bilayer -HCP with BCBC… sequence. In such a process, -HCP phase is formed.
Therefore, the intrinsic stacking fault energy ( ), involved with the interruption of
normal stacking sequence, holds the key to evaluate the possibility of the phase
transformation, and it can be expressed for FCC-structured specimens as [46]:
→ →
= 2+, Δ. + 20 , (2)

where Δ. is the molar free energy difference between the -FCC and -HCP

phase, 0 is the coherent interfacial energy between these two phases (15 mJ/m2)
"
and +, = is the planar packing density of the close-packed - 111 plane
√ 23 4

based on the lattice constant of -FCC phase ( = 3.6 Å) and Avogadro’s number

(5 = 6.02 × 10# ). The intrinsic stacking fault energy is determined by the free
energy difference between -FCC and -HCP with the corresponding lattice constant.
Table 2 lists the room-temperature free energy of -FCC and -HCP phases for both
DP and TP specimens calculated by the Thermo-Calc software based on the
CALPHAD method. The matrix composition of TP specimen for calculation is shown
in Table. 1. Accordingly, the intrinsic stacking fault energy of the DP and TP
specimens are 10.9 mJ/m2 and 12.2 mJ/m2, respectively. Since the intrinsic stacking
fault energy could quantify the likelihood of the dislocations to dissociate into two
partials, the DP specimens with smaller intrinsic stacking fault energy are more
susceptible to the phase transformation.
We proceed to discuss the temperature gradient during water quenching.
According to the free energy of the -FCC and -HCP phases (Table. 2), the -HCP

12
phase was more stable than the -FCC at ambient temperature in both DP and TP
specimens. In addition to the thermodynamic analysis, the phase transformation is
considered to be a kinetic process and it requires an activation energy to overcome the
energy barrier for the transformation process. Temperature gradient during water
quenching provided the activation energy for the phase transformation from -FCC
to -HCP. Even though the kinetic process is not the focus of this work, a qualitative
affirmation is certain that DP specimen quenched from 800 or 750 consumed
more energy for phase transformation than TP specimen quenched from 700 .
Conclusively, the stacking fault energy (thermodynamic process) and temperature
gradient (kinetic process) result in the microstructure difference between DP
specimens and the matrix of TP specimen.

Table 2. Free energy and free energy difference of -FCC and -HCP phase in
dual-phase and triple-phase MEAs.

Molar energy (J/mol) DP specimen TP specimen

. -721 -763

. -1047 -1067


6. -326 -304

3.3 Mechanical properties of dual-phase and triple-phase MEAs


Figure 8(a) shows the stress-strain curves of the examined specimens from the
uniaxial tensile test. The yield strength of the specimens annealed at 800 , 750
and 700 were ~600 MPa, ~700 MPa and ~1050 MPa, while the ultimate strength
(0789 ) were 840 MPa, 880 MPa and 1150 MPa, respectively. For DP specimen, both
the yield strength (0: ) and ultimate strength (0789 ) increase with the decrease of

annealing temperature because of grain refinement from 6.1 ± 0.4 μm to 5.4 ± 0.7
μm. Compared to the DP specimens, the mechanical performance of the TP specimen
was impressively improved, with an elongation of ~25%. Figure 8(b) shows the
13
work-hardening effect of the specimens as a function of the true strain. The
work-hardening rate of these specimens decreased sharply at the beginning of the
plastic deformation, followed by a weak descend as the strain further increased.
Because of the precipitation strengthening of the third phase, the work-hardening
effect of TP specimen is stronger than that of DP specimens. Nevertheless, all of these
specimens exhibited a comparatively low level of work-hardening rate, due to the
extremely low stacking fault energy induced phase transformation.

Fig. 8. (a) Engineering stress-strain curve of tensile tests and (b) working hardening
rate versus true strain curve for the DP specimens (recrystallized at 750 and 800 )
and TP specimen (recrystallized at 700 )

3.4 Phase evolution in the Fe42Co42Cr16 MEA

14
Figure 9 shows an overview of the phase evolution in the Fe42Co42Cr16 MEA
under varied heat treatments. As shown in Fig. 9(a) and (b), partial martensitic
transformation of the -FCC into the -HCP phase occurred when the specimens
were quenched from the high-temperature single-phase region (Fig. 7). Owing to the
low stacking fault energy (10.9 mJ/m2), the instability of -FCC phase during water
quenching was responsible for the DP specimens at ambient temperature. Then,
another plastic deformation induced the martensite phase, -BCT (body centered
tetragonal) appeared in the cold rolled DP specimens, according to the EBSD and
XRD results, which provide evidence for the coexistence of -FCC, -HCP and
-BCT (Fig. 9(c)). Specifically, during the complete polymorphic plastic deformation
induced martensitic phase transformation process, the formation of the -BCT
martensite phase was preceded by a prior transformation of the -FCC austenite
phase into -HCP martensite phase. As shown in Fig. 9(d), when the cold rolled
specimens recrystallized at 750 or 800 , the specimens resumed the
high-temperature single phase region of -FCC austenite (Fig. 7) and partial
martensitic transformation of the -FCC to the -HCP phase occurred due to the
subsequent water quenching. During the recrystallization process, the grain size was
refined. In contrast, an equilibrium B2-BCC precipitated in the specimen
recrystallized at 700 , while a partial martensitic transformation occurred in the
matrix during water quenching (Fig. 9(e)). Consequently, the TP specimen, composed
of -FCC austenite phase, -HCP martensite phase and B2-BCC precipitates, was
generated with such a process.
Here, it should be clarified that the EBSD phase mapping produces intolerable
inaccuracy for the analysis of phase fraction and orientation in the current alloy
because of the incongruence between the nano-scaled hierarchical microstructure and
spatial resolution limitation of EBSD. As shown in Fig. 9 (d), the sub-micro-scaled
and blocks with nanolaminates (shown in Fig. 3) cannot be detected by EBSD.
Meanwhile, Transmission Kikuchi diffraction (TKD) method [46], as shown in Fig.
10, offers a phase mapping with spatial resolution up to ~40 nm, depending on the

15
signal collection time and the limitation of SEM resolution. However, a significant
part of the microstructure below 40 nm is unavailable because of the spatial resolution
of SEM. In addition, the detection range of TKD is only able to cover tens of microns
and it’s too limited for the phase fraction counting. Therefore, for Fe42Co42Cr16 MEA,
the EBSD could be only used for qualitative analysis of the phase component, but a
detailed analysis could only be available from the TEM result.

Fig. 9. EBSD phase mapping and corresponding XRD results of (a) as-cast, (b) water
quenched after homogenized at 1200 for 2h, (c) ~67% cold-rolled, and water
quenched after recrystallization at (d) 750 and (e) 700 specimens.

Fig. 10 (a) Phase image and (b) IPF-Z mapping of DP specimen recrystallized at 750
with the spatial resolution of ~40 nm.

3.5 Plastic deformation induced martensitic transformation in Fe42Co42Cr16 MEA


As mentioned above, polymorphic plastic induced martensitic transformation
16
from -FCC→ -HCP→ -BCT was confirmed in the Fe42Co42Cr16 MEA. The
overview of the microstructures of the deformed DP specimen with recrystallization at
750 , as shown in the bright-field TEM image in Fig. 11(a), demonstrates the
coexistence of the -FCC, -HCP and -BCT phases. The deformed TEM sample
was produced from the necking region in the gauge section of the tensile test
specimen. Figure 11(b) exhibits the SADP along the -〈110〉 zone axis for the circle
area in (a). Obviously, the orientation of the three phases yielded the typical
Kurdjumov-Sachs (K-S) [48]; i.e., -〈110〉// -〈1120〉// -〈001〉, - 111 //
- 0002 // - 110 and -=1100>// -=110>. The phase transformation mechanism
from -FCC to -HCP has been mentioned above and the subsequent transition from
-HCP to -BCT could be interpreted according to the model established by Xu et al
[49] in the austenite stainless steel with low stacking fault energy. During -HCP to
-BCT transition, a pair of half Shockley partial dislocations with the Burgers vectors
"" ""
of opposite sign, (112) and − (112), glide on both sides of the every second
#* #*

of the -=0001> plane. The original inclined angle was 70.5° between -=111>

plane and -(112) zone axis. Due to the dipole dislocation slipping on both sides of

the -=0001> plane, the angle was rectified to 90° between -=110> plane and

-(110) zone axis, as schematically shown in Fig. 11(c). Furthermore, to release the
unstable stacking fault energy and stress energy generated by half Shockley partial
dislocations slipping, the involved atoms shuffled, leading to the change in the

d-spacing of -=110> to -=0002> of 2.08 Å. Therefore, the transformation from


-HCP to -BCT was accomplished by the synergy of half partial dislocations
slipping and atoms shuffling.
In summary, the transformation from -FCC to -HCP is mainly governed by a
perfect dislocation dissociation into partial dislocations, while the subsequent
transformation from -HCP to -BCT results from atoms shuffling and the partial
dislocation dissociation into a half partial dislocations. Compared with the parent
austenite -FCC, the -BCT exhibits a compression by 18.4% along -(001), an

17
expansion by 15.3% along -(110) and unchanging dimension in the -(110)
direction [50, 51]. This volume change could provide additional plasticity and strain
hardening.

Fig. 11. Bright-field TEM image of the deformed DP specimen recrystallized at 750 .
(b) SADP along -〈110〉// -〈1120〉// -〈001〉 zone axis from the circled region in
(a). (c) Schematic summarizing diffraction spots and phase orientation from (b).

3.6 Precipitation strengthening in the triple-phase MEA


The bright-field TEM image in Fig. 12(a) shows an overview of microstructures
of the deformed ( ~ 20%) TP specimen recrystallized at 700 . Noticeably, compared
to that of the water quenched TP specimen with a partial martensitic transformation in
the matrix (Fig. 5), plastic deformation provided sufficient energy to accomplish the
complete transformation ( -FCC→ -HCP→ -BCT). The phases orientation
between -FCC, -HCP and -BCT followed the typical K-S relationship
mentioned above. In addition, the hierarchical structure with -FCC, -HCP
nanolaminates was also established during plastic deformation.
According to the tensile test results shown in Fig. 8, the TP specimen exhibits a
dramatical improvement in mechanical performance with an increase in the yield
strength from 600 MPa to 1050 MPa (~75 %) and the ultimate tensile strength from
850 MPa to 1150 MPa (~35%). Even though the reduced grain size and relatively
higher stacking fault energy could contribute to the strengthening, a significant
portion of the strengthening effect results from B2-BCC precipitates. As shown in Fig.
12 (b), a dispersed B2-BCC precipitate impedes the movement of planar dislocations

18
(marked by parallel lines) and accordingly enhances the strength of the alloy.
However, because of the low SFE of the matrix, the impeded planar dislocations are
prone to dissociation into partial dislocations instead of accumulation. It provides an
explanation for the limited improvement of the work-hardening rate for the TP
specimen.

Fig. 12. (a) Bright-field TEM images of deformed TP specimen recrystallized at


700 , including -FCC austenite, -HCP martensite, -BCT martensite phase and
B2-BCC precipitate. (b) Enlarged view for the rectangle region in (a), indicating
slipping dislocations impeded by the B2-BCC precipitate.

4. Conclusion
In this work, an innovative Fe42Co42Cr16 MEA was fabricated. By regulating the
heat treatment process, DP (dual-phase, -FCC and -HCP) with the hierarchical
nanostructure and TP (triple-phase, -FCC, -HCP and B2-BCC) specimen with
19
precipitates were obtained. Partial martensitic transformation; i.e., -FCC⟶ -HCP,
and the complete martensitic transformation; i.e., -FCC⟶ -HCP⟶ -BCT, were
confirmed and the transformation mechanism was discussed with the following main
conclusions.
(1) The Fe42Co42Cr16 MEA exhibits a wide single-phase region from 765 to
the liquidus. The -HCP formed in the DP specimen with recrystallization at
750 and 800 was a non-equilibrium phase caused by the partial
martensitic transformation during water quenching.
(2) The fabricated TP specimen consisted of two equilibrium phases ( -FCC and

B2-BCC) formed during recrystallization at 700 ℃ and the martensitic

phase ( -HCP) generated during water quenching.


(3) Because of the low SFE (10.9 mJ/m2), hierarchical structure with -FCC and
-HCP nanolaminates were generated in DP specimens, which were absent in
the matrix of TP specimen due to the increased SFE (12.2 mJ/m2).
(4) Complete martensitic transformation ( -FCC⟶ -HCP⟶ -BCT) occurred
in both DP and TP specimens during plastic deformation. The -FCC⟶
" "
-HCP transformation was dominated by the (211) and (121) partial
* *

dislocation slipping on every second of -=111> plane; The -HCP⟶


-BCT transformation was ascribed to the slipping of half partial dislocations
"" ""
(# * (112) and − # * (112)) on the every second of -=0001> plane and

changed in d-spacing of -=110> equal to -=0002> due to atomic


shuffling. The orientation between these phases was confirmed to be
-〈110〉// -〈1120〉// -〈001〉.
(5) The mechanical performance of the TP alloy was strikingly improved by the
B2-BCC precipitates, with a yield strength of ~1050 MPa and an ultimate
strength of 1150 MPa.
This work provides an optional base alloy for the further alloy design, especially
for the alloys with TRIP effect. It is clear that the mechanical performance can be

20
improved by optimizing the alloy composition and heat treatment process, or
importing reinforcement phase in the alloys.

Acknowledgments
This research was financially supported by the Hong Kong Government through the
General Research Funds (GRFs) with the grant number of CityU 11202718 and
11209314, and the Ministry of Science and Technology, Taiwan under Contract no.
MOST (108-2218-E-002-062).

Reference
[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(5) (2004) 299-303.
[2] B. Cantor, I.T.H. Chang, P. Knight, A.J.B. Vincent, Microstructural development in equiatomic
multicomponent alloys, Mat. Sci. Eng. A-Struct. 375 (2004) 213-218.
[3] K.C. Hsieh, C.F. Yu, W.T. Hsieh, W.R. Chiang, J.S. Ku, J.H. Lai, C.P. Tu, C.C. Yang, The
microstructure and phase equilibrium of new high performance high-entropy alloys, J. Alloy. Compd.
483(1-2) (2009) 209-212.
[4] O.N. Senkov, G.B. Wilks, D.B. Miracle, C.P. Chuang, P.K. Liaw, Refractory high-entropy alloys,
Intermetallics 18(9) (2010) 1758-1765.
[5] F. Zhang, Y. Wu, H.B. Lou, Z.D. Zeng, V.B. Prakapenka, E. Greenberg, Y. Ren, J.Y. Yan, J.S.
Okasinski, X.J. Liu, Y. Liu, Q.S. Zeng, Z.P. Lu, Polymorphism in a high-entropy alloy, Nat. Commun.
8 (2017) 15697.
[6] 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.
[7] Y.P. Lu, X.Z. Gao, L. Jiang, Z.N. Chen, T.M. Wang, J.C. Jie, H.J. Kang, Y.B. Zhang, S. Guo, H.H.
Ruan, Y.H. Zhao, Z.Q. Cao, T.J. Li, Directly cast bulk eutectic and near-eutectic high entropy alloys
with balanced strength and ductility in a wide temperature range, Acta Mater. 124 (2017) 143-150.
[8] S.R. Reddy, S. Bapari, P.P. Bhattacharjee, A.H. Chokshi, Superplastic-like flow in a fine-grained
equiatomic CoCrFeMnNi high-entropy alloy, Mater. Res. Lett. 5(6) (2017) 408-414.
[9] 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(6201) (2014) 1153-1158.
[10] M.X. Yang, L.L. Zhou, C. Wang, P. Jiang, F.P. Yuan, E.V. Ma, X.L. Wu, High impact toughness of
CrCoNi medium-entropy alloy at liquid-helium temperature, Scripta Mater. 172 (2019) 66-71.
[11] Z. Tang, L. Huang, W. He, P.K. Liaw, Alloying and Processing Effects on the Aqueous Corrosion
Behavior of High-Entropy Alloys, Entropy-Switz 16(2) (2014) 895-911.
[12] H. Luo, Z.M. 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)
21
131-139.
[13] C.Y. Hsu, T.S. Sheu, J.W. Yeh, S.K. Chen, Effect of iron content on wear behavior of
AlCoCrFexMo0.5Ni high-entropy alloys, Wear 268(5-6) (2010) 653-659.
[14] Y.X. Ye, C.Z. Liu, H. Wang, T.G. Nieh, Friction and wear behavior of a single-phase equiatomic
TiZrHfNb high-entropy alloy studied using a nanoscratch technique, Acta Mater. 147 (2018) 78-89.
[15] K.Y. Tsai, M.H. Tsai, J.W. Yeh, Sluggish diffusion in Co-Cr-Fe-Mn-Ni high-entropy alloys, Acta
Mater. 61(13) (2013) 4887-4897.
[16] S.J. Zhao, Y. Osetsky, Y.W. Zhang, Preferential diffusion in concentrated solid solution alloys:
NiFe, NiCo and NiCoCr, Acta Mater. 128 (2017) 391-399.
[17] W.Y. Chen, X. Liu, Y.R. Chen, J.W. Yeh, K.K. Tseng, K. Natesan, Irradiation effects in high
entropy alloys and 316H stainless steel at 300 degrees C, J. Nucl. Mater. 510 (2018) 421-430.
[18] T.F. Yang, S.Q. Xia, W. Guo, R. Hu, J.D. Poplawsky, G. Sha, Y. Fang, Z.F. Yan, C.X. Wang, C.Y.
Li, Y. Zhang, S.J. Zinkle, Y.G. Wang, Effects of temperature on the irradiation responses of
Al0.1CoCrFeNi high entropy alloy, Scripta Mater. 144 (2018) 31-35.
[19] Z.M. Li, D. Raabe, Strong and Ductile Non-equiatomic High-Entropy Alloys: Design, Processing,
Microstructure, and Mechanical Properties, Jom-US 69(11) (2017) 2099-2106.
[20] B. Gludovatz, A. Hohenwarter, K.V.S. Thurston, H.B. Bei, Z.G. Wu, E.P. George, R.O. Ritchie,
Exceptional damage-tolerance of a medium-entropy alloy CrCoNi at cryogenic temperatures, Nat.
Commun. 7 (2016) 10602.
[21] H.W. Yao, J.W. Qiao, M.C. Gao, J.A. Hawk, S.G. Ma, H.F. Zhou, MoNbTaV Medium-Entropy
Alloy (vol 18, 189, 2016), Entropy-Switz 18(8) (2016) 189.
[22] F. Otto, A. Dlouhy, 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(15) (2013)
5743-5755.
[23] B. Schuh, F. Mendez-Martin, B. Volker, E.P. George, H. Clemens, R. Pippan, A. Hohenwarter,
Mechanical properties, microstructure and thermal stability of a nanocrystalline CoCrFeMnNi
high-entropy alloy after severe plastic deformation, Acta Mater. 96 (2015) 258-268.
[24] J.Y. He, C. Zhu, D.Q. Zhou, W.H. Liu, T.G. Nieh, Z.P. Lu, Steady state flow of the FeCoNiCrMn
high entropy alloy at elevated temperatures, Intermetallics 55 (2014) 9-14.
[25] Y.L. Zhao, T. Yang, Y. Tong, J. Wang, J.H. Luan, Z.B. Jiao, D. Chen, Y. Yang, A. Hu, C.T. Liu, J.J.
Kai, Heterogeneous precipitation behavior and stacking-fault-mediated deformation in a CoCrNi-based
medium-entropy alloy, Acta Mater. 138 (2017) 72-82.
[26] 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(6417) (2018) 933-937.
[27] C. Wang, T.H. Li, Y.C. Liao, C.L. Li, J.S.C. Jang, C.H. Hsueh, Hardness and strength
enhancements of CoCrFeMnNi high-entropy alloy with Nd doping, Mat. Sci. Eng. A-Struct. 764 (2019)
138192.
[28] Z.M. Li, K.G. Pradeep, Y. Deng, D. Raabe, C.C. Tasan, Metastable high-entropy dual-phase alloys
overcome the strength-ductility trade-off, Nature 534(7606) (2016) 227-230.
[29] W.J. Lu, C.H. Liebscher, G. Dehm, D. Raabe, Z.M. Li, Bidirectional Transformation Enables
Hierarchical Nanolaminate Dual-Phase High-Entropy Alloys, Adv. Mater. 30(44) (2018) 1804724.
[30] Y.Q. Bu, Z.M. Li, J.B. Liu, H.T. Wang, D. Raabe, W. Yang, Nonbasal Slip Systems Enable a
Strong and Ductile Hexagonal-Close-Packed High-Entropy Phase, Phys. Rev. Lett. 122(7) (2019)
22
075502.
[31] J. Lu, L. Hultman, E. Holmstrom, K.H. Antonsson, M. Grehk, W. Li, L. Vitos, A. Golpayegani,
Stacking fault energies in austenitic stainless steels, Acta Mater. 111 (2016) 39-46.
[32] B. Qin, H.K.D.H. Bhadeshia, Plastic strain due to twinning in austenitic TWIP steels, Mater. Sci.
Tech.-Lond. 24(8) (2008) 969-973.
[33] Y.H. Zhang, Y. Zhuang, A. Hu, J.J. Kai, C.T. Liu, The origin of negative stacking fault energies
and nano-twin formation in face-centered cubic high entropy alloys, Scripta Mater. 130 (2017) 96-99.
[34] C. Herrera, D. Ponge, D. Raabe, Design of a novel Mn-based 1 GPa duplex stainless TRIP steel
with 60% ductility by a reduction of austenite stability, Acta Mater. 59(11) (2011) 4653-4664.
[35] Z. Li, C.C. Tasan, K.G. Pradeep, D. Raabe, A TRIP-assisted dual-phase high-entropy alloy: Grain
size and phase fraction effects on deformation behavior, Acta Mater. 131 (2017) 323-335.
[36] C.G. Rhodes, A.W. Thompson, The composition dependence of stacking fault energy in austenitic
stainless steels, Metall. Trans. A 8(12) (1977) 1901-1906.
[37] J.W. Yeh, Recent progress in high-entropy alloys, Ann. Chim. Sci. Mat. 31(6) (2006) 633-648.
[38] G.B. Olson, M. Cohen, A mechanism for the strain-induced nucleation of martensitic
transformations, J. Less Common. Met. 28(1) (1972) 107-118.
[39] G. Olson, M. Cohen, A general mechanism of martensitic nucleation: Part II. FCC→ BCC and
other martensitic transformations, Metall. Trans. A 7(12) (1976) 1905-1914.
[40] G. Olson, M. Cohen, A perspective on martensitic nucleation, Ann. Rev. Mater. Sci. 11(1) (1981)
1-32.
[41] J.O. Andersson, T. Helander, L.H. Hoglund, P.F. Shi, B. Sundman, THERMO-CALC & DICTRA,
computational tools for materials science, Calphad 26(2) (2002) 273-312.
[42] W.H. Liu, Z.P. Lu, J.Y. He, J.H. Luan, Z.J. Wang, B. Liu, Y. Liu, M.W. Chen, C.T. Liu, Ductile
CoCrFeNiMox high entropy alloys strengthened by hard intermetallic phases, Acta Mater. 116 (2016)
332-342.
[43] C. Ng, S. Guo, J. Luan, Q. Wang, J. Lu, S. Shi, C. Liu, Phase stability and tensile properties of
Co-free Al0. 5CrCuFeNi2 high-entropy alloys, J. Alloy Compd. 584 (2014) 530-537.
[44] D.C. Ma, M.J. Yao, K.G. Pradeep, C.C. Tasan, H. Springer, D. Raabe, Phase stability of
non-equiatomic CoCrFeMnNi high entropy alloys, Acta Mater. 98 (2015) 288-296.
[45] G. Olson, M. Cohen, Dislocation theory of martensitic transformations, Dislocations in solids,
North-Holland (1986) 295-407.
[46] P. Adler, G. Olson, W. Owen, Strain hardening of Hadfield manganese steel, Metall. Mater. Trans.
A 17(10) (1986) 1725-1737.
[47] R.R. Keller, R.H. Geiss, Transmission EBSD from 10 nm domains in a scanning electron
microscope, J. Microsc.-Oxford 245(3) (2012) 245-251.
[48] G. Kurdjumov, G. Sachs, Über den Mechanismus tier Stahlhärtung, Z. Phys. 64 (1930) 325-343.
[49] X.S. Yang, S. Sun, H.H. Ruan, S.Q. Shi, T.Y. Zhang, Shear and shuffling accomplishing
polymorphic fcc gamma → hcp epsilon → bct alpha martensitic phase transformation, Acta Mater. 136
(2017) 347-354.
[50] J.K. Diao, K. Gall, M.L. Dunn, Surface-stress-induced phase transformation in metal nanowires,
Nat. Mater. 2(10) (2003) 656-660.
[51] H. Zheng, A.J. Cao, C.R. Weinberger, J.Y. Huang, K. Du, J.B. Wang, Y.Y. Ma, Y.N. Xia, S.X. Mao,
Discrete plasticity in sub-10-nm-sized gold crystals, Nat. Commun. 1 (2010) 144.

23
Highlights
Partial martensitic transformation, i.e., 𝛾 FCC ⟶ 𝜀 HCP, occurred during water
quenching in the dual phase Fe42Co42Cr16 medium entropy alloy.

Triple phase alloy with 𝛾 FCC austenite, 𝜀 HCP martensite and B2 BCC precipitates,
was obtained by quenching the alloy in 𝛾 + BCC phase region.

Plastic deformation induced complete martensitic transformation, i.e., 𝛾 FCC⟶ 𝜀


HCP⟶ α BCT, were conformed in both dual phase and triple phase Fe42Co42Cr16
medium entropy alloys.

The triple phase medium entropy alloy exhibited a yield strength above 1 GPa with a
total elongation of as high as 25%.
Declaration of interests

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

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like