You are on page 1of 8

Materials Science & Engineering A 636 (2015) 188–195

Contents lists available at ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

High temperature deformation behavior and dynamic recrystallization


in CoCrFeNiMn high entropy alloy
N.D. Stepanov a,n, D.G. Shaysultanov a, N.Yu. Yurchenko a, S.V. Zherebtsov a, A.N. Ladygin b,
G.A. Salishchev a, M.A. Tikhonovsky b
a
Laboratory of Bulk Nanostructured Materials, Belgorod State University, Belgorod 308015, Russia
b
National Science Center “Kharkov Institute of Physics and Technology” NAS of Ukraine, Kharkov 61108, Ukraine

ar t ic l e i nf o a b s t r a c t

Article history: Microstructure evolution in high-entropy alloy CoCrFeNiMn during uniaxial compression to a height
Received 25 February 2015 reduction of true strain of E1.4 in the temperature interval 600–1100 1C was studied. Although some
Received in revised form differences was observed in the mechanical behavior of the alloy and the activation energy of
23 March 2015
deformation in warm (below 800 1C) and hot (above 800 1C) temperature intervals, microstructure
Accepted 25 March 2015
Available online 3 April 2015
evolution at all studied temperatures was found to be accompanied by discontinuous dynamic
recrystallization (dDRX). During hot deformation recrystallization was primary associated with nuclea-
Keywords: tion of new grains on the initial grain boundaries, while in the warm interval dDRX was mainly observed
EBSD in shear bands. The volume fraction of the recrystallized structure was respectively 0.085 and 0.95 at 600
Mechanical characterization
and 1000 1C and the recrystallized grain size was found to be 0.2 and 40.4 mm for 600 and 1100 1C,
High entropy alloys
respectively.
Thermomechanical processing
Recrystallization & 2015 Elsevier B.V. All rights reserved.
Shear bands

1. Introduction CoCrFeNi and CoCrFeNiMn alloys at elevated temperatures [17]. It


should be noted that the majority of the published data, especially at
So-called high entropy alloys (HEAs) are the new class of materials elevated temperatures, was obtained during tensile testing [15–17,19].
defined as alloys consisting of 5 or more principal elements with nearly This deformation scheme limits the maximum attained strain thereby
equimolar fractions [1–3]. They have received considerable attention in focusing the main attention of the researchers to microstructure
recent years due to their unusual structures and properties. High evolution associated with low strain. Meanwhile the available data on
entropy of mixing associated with multiple principle components is compressive test is mainly limited to room temperature whereas this
thought to prevent formation of intermetallic phases [1] and promote method is very useful for investigation of microstructure evolution
formation of simple solid solution structures. However, among the and mechanical behavior during large deformation in a wide tem-
equiatomic alloys based on transition metals only a couple have truly perature interval.
single solid solution structures, namely CoCrFeNi and CoCrFeNiMn Although thermo-mechanical treatment via warm or hot
alloys [4–6]. These alloys consist of single fcc solid solution. Adding deformation is rather rarely applied to HEAs, it can significantly
other elements like Al, V, Ti, Mo or Nb to CoCrFeNi and CoCrFeNiMn refine the microstructure and enhance the mechanical properties.
alloys results in formation of ordered or intermetallic phases [7–13]. For example, hot multiaxial forging of the multiphase AlCoCrCu-
That is why the CoCrFeNi and CoCrFeNiMn alloys have attracted FeNi alloy has enabled formation of structure with the average
considerable attention of the researchers. grain/particle size of 1.5 mm and promote superplastic behavior of
Remarkable mechanical properties of these alloys were reported the forged alloy at elevated temperature [20–22].
in [14–19]. It was demonstrated that they have relatively low yield It has been already reported that grain size has a very pronounced
stress, excellent strain hardening capability and exceptional ductility effect on mechanical properties of the CoCrFeNiMn alloy [15]. To
at ambient and cryogenic temperatures [14,15]. The CoCrFeNiMn alloy produce recrystallized grains in the CoCrFeNiMn alloy combination of
also has promising fracture toughness at cryogenic temperature [16]. cold working with subsequent annealing i.e. static recrystallization is
Much less attention has been paid to mechanical behavior of the usually used [14,15]. Static recrystallization behavior of the alloy has
been studied quite thoroughly [23–26]. However, no detailed studies
on dynamic recrystallization behavior of the CoCrFeNiMn alloy or any
n
Corresponding author. Tel.: þ 7 4722 585416.
other HEA are available in the literature. It might be supposed that
E-mail addresses: stepanov@bsu.edu.ru, such features as sluggish diffusion [27] or low stacking fault energy
stepanov.nikita@icloud.com (N.D. Stepanov). [28] could influence on dynamic recrystallization behavior of the

http://dx.doi.org/10.1016/j.msea.2015.03.097
0921-5093/& 2015 Elsevier B.V. All rights reserved.
N.D. Stepanov et al. / Materials Science & Engineering A 636 (2015) 188–195 189

CoCrFeNiMn alloy. Therefore the aim of the current study is to


examine the compressive mechanical behavior and corresponding
microstructural response of the CoCrFeNiMn alloy at elevated tem-
peratures in the interval 600–1100 1C.

2. Experimental procedures

The equiatomic alloy with the composition of CoCrFeNiMn was


produced by arc melting of the components in high-purity argon
inside a water-cooled copper cavity. The purities of the alloying
elements were above 99.9(at.)%. To ensure chemical homogeneity,
the ingots were flipped over and re-melted a least 5 times. The
produced ingots of the CoCrFeNiMn alloy had dimensions of about
6  15  60 mm3. Homogenization annealing was carried out at
1000 1C for 24 h in accord with regime used previously for the
alloy [9,10,14]. Prior to homogenization, the samples were sealed Fig. 1. Stress–strain curves obtained during compressive testing of the CoCrFeNiMn
alloy at strain rate of 10  3 s  1 and temperatures of 600–1100 1C.
in vacuumed (10  2 Torr) quartz tubes filled with titanium chips to
prevent oxidation. After annealing, the tubes were removed from
the furnace and the samples were cooled inside the vacuumed Table 1
tubes down to room temperature due to heat exchange with Compressive mechanical properties of the CoCrFeNiMn high entropy alloy at strain
rate of 10  3 s  1 and temperatures of 600–1100 1C.
surrounding air.
Compressive tests were performed on rectangular specimens Temperature (1C) σ0.2 (MPa) σss (MPa)
with dimensions of 7  5  5 mm3 using the Instron 300LX machine
equipped with radial heating furnace. Testing was performed at 600 125 –
700 100 –
temperatures of 600 1C, 700 1C, 800 1C, 900 1C, 1000 1C and 1100 1C.
800 87 –
The initial strain rate was of 10  3 s  1. To estimate activation energy, 900 68 156
additional tests were performed at temperatures of 700–1100 1C and 1000 66 100
strain rates of 10  2 s  1 and 10  4 s  1. All samples were compressed 1100 36 50
to height reduction of 75% corresponding to true strain of E1.4. In
addition, specimens were compressed to height reductions of 25%
and 50–60% (true strain of E0.3 and E0.7–0.9) at temperatures of strengthened during deformation till the maximum strain. Strain
700 1C and 1000 1C and strain rate of 10  3 s  1. hardening capability increases with decreasing deformation tem-
The microstructure of compressed specimens was examined on perature. Serrations are observed at two minimum temperatures of
the plane parallel to the compression axis on the mid-thickness of deformation; the number of serrations decreases with temperature
the specimen. Electron backscattered diffraction (EBSD) was increasing.
mainly employed for microstructure characterization. A FEI Nova- To establish whether the difference between the deformation
NanoSEM 450 scanning electron microscope (SEM) equipped with behavior in two temperature intervals is associated with a change in
a Hikari EBSD detector was used to produce EBSD maps. Samples the mechanism of microstructure evolution, activation energy of
were carefully mechanically polished and then underwent elec- deformation for these temperatures was determined. To this end
tropolishing in mixture of 90% acetic and 10% perchloric acids at compression tests at strain rates of 10  4 s  1 and 10  2 s  1 in the
room temperature and applied voltage of 27 V for 15 s. TSL OIM interval 700–1100 1C were performed. The deformation behavior of
Analysis 6 was used to process EBSD data and generate inverse the alloy under these conditions was found to be followed the above
pole figure (IPF) maps. On the presented maps high angle described trend, i.e. decrease in strain rate and increase in temperature
boundaries (HABs) are indicated with black lines and low angle decreased propensity to strain hardening to almost zero at 1100 1C
boundaries (LABs) are indicated with white lines. Points having and 10  4 s  1. The values of yield strength σ0.2 and steady state flow
low confidence index (CI r0.1) were excluded from analysis and stress σss obtained for strain rates of 10  4–10  2 s  1 and temperatures
are depicted as black dots in presented IPF maps. 700–1100 1C are summarized in Table 2. For low temperatures (700
and 800 1C) flow stresses at e¼ 0.4 were taken as σss.
The relation between strain rate and temperature during high-
3. Results temperature deformation can generally be described with the
well-known Zener–Hollomon parameter Zðε_ Þ [29–31]:
3.1. Mechanical behavior 
Zðε_ Þ ¼ Aε_ n ¼ ε_ exp Q =RT ; ð1Þ

True stress–true strain curves obtained during compressive test- where Q is the activation energy, n is the stress exponent, A is a
ing of the CoCrFeNiMn alloy are shown in Fig. 1. In addition values of constant sensitive to the deformation mechanism, and R is the gas
some important characteristics (yield strength σ0.2, and steady state constant. The parameters n and Q can be determined as
flow stress σss) are summarized in Table 1.
∂ ln ε_
It should be noted that all samples have been compressed to the n¼ : ð2Þ
∂ lnðσ ss =GÞ T
maximum height reduction (E75%, corresponding to true strain of
E1.4) without any signs of fracture. The mechanical behavior of the ∂ ln ε_ ∂ lnðσ ss =GÞ
alloy significantly depended on deformation temperature. In the Q ¼ R  nR ; ð3Þ
∂ð1=TÞ σ ∂ð1=TÞ ε_
interval 900–1100 1C the alloy demonstrated deformation curves
with a well-defined steady state flow stage following a short hard- where ε_ is the deformation strain rate, σss denotes the steady state
ening stage in the very beginning of deformation. In contrast the flow stress, n the stress exponent, R the gas constant, and G is the
alloy compressed in the interval 600–800 1C continuously shear modulus of the CoCrFeNiMn alloy at given temperature,
190 N.D. Stepanov et al. / Materials Science & Engineering A 636 (2015) 188–195

Table 2
Compressive mechanical properties of the CoCrFeNiMn high entropy alloy at temperatures of 800–1100 1C and strain rates of 10  4–10  2 s  1.

Temperature (1C) 700 800 900 1000 1100

ε_ (s  1) 10  4 10  3 10  2 10  4 10  3 10  2 10  4 10  3 10  2 10  4 10  3 10  2 10  4 10  3 10  2
σ0.2 (MPa) 74 100 72 46 87 74 66 68 84 36 66 75 18 36 48
σss (MPa) 282 375 382 174 264 305 93 156 207 54 100 130 31 50 80

Fig. 2. Logarithmic plot of dependence of state strain rate on the normalized stress (a) and Arrhenius plot of logarithmic steady-state flow stress vs the inverse of
temperature (b).

Fig. 3. The XRD pattern (a) and inverse pole figure EBSD map (b) of the homogenized CoCrFeNiMn alloy.

calculated using the equation below [32]: as a linear function with the sloped dependent on strain rates in
  the temperature interval 1100–800 1C. At lower temperature the
G ¼ 85–16= e448=T  1 ; ð4Þ slopes decrease suggesting transition to deformation with quite
weak temperature and strain rate dependence [33]. In accordance
To determine the stress exponent n, logarithmic plot of strain with Eq. (3) the value of activation energy was found to be 291 kJ/
rate against steady state stress (corresponding values were taken mole and 213 kJ/mole for the high-temperature and the low-
from Table 2) at different temperatures was plotted (Fig. 2a). temperature intervals, respectively.
Experimental data can be approximated rather well by a linear
dependence with the slope n ranged from E 5.18 to E11.4 for 3.2. Microstructure evolution during compression test
temperatures 1100–700 1C.
The dependence of logarithm of the steady state stress values In the initial (homogenized) condition CoCrFeNiMn alloy has an fcc
on inverse absolute temperature, 1/T, at three different strain rates single-phase microstructure (Fig. 3a). An EBSD inverse pole figure (IPF)
is given in Fig. 2b. The observed relationships can be approximated map shows rather inhomogeneous microstructure with the grain size of
N.D. Stepanov et al. / Materials Science & Engineering A 636 (2015) 188–195 191

100–500 mm. Many grains have elongated shape which most likely compression at 600 1C the initial grains were found to be sliced
was produced during solidification. Some annealing twins and residual by a network of shear bands (Fig. 4a). The shear bands having
porosity (black dots in Fig. 3b) are observed in the microstructure. width of about several microns were inclined predominantly at
Inverse pole figure (IPF) maps of the CoCrFeNiMn alloy after  451 to compression direction. Sometimes shear bands coincided
compression at different temperatures to true strain of E1.4 are with initial grain boundaries. Very fine equiaxed grains can be
shown in Fig. 4. The microstructure of the deformed alloy detected inside the shear bands. In those parts of grains which
considerably depended on the deformation temperature. After were not involved in shear deformation substructure developed.

Fig. 4. EBSD IPF maps of the CoCrFeNiMn alloy after compression at strain rate of 10  3 s  1 and compressive true strain of E1.4 at different temperatures: (a) 600 1C;
(b) 700 1C; (c) 800 1C; (d) 900 1C; (e) 1000 1C; (f) 1100 1C.
192 N.D. Stepanov et al. / Materials Science & Engineering A 636 (2015) 188–195

Fig. 5. TEM bright field images of the microstructure of shear bands in specimens of CoCrFeNiMn alloy strained to E 75% at strain rate of 10  3 s  1 and temperatures 600 1C
(a) and 700 1C (b).

Increase in deformation temperature to 700, 800 and 900 1C


(Fig. 4b, c and d) resulted in the increase both the width of the shear
bands and the size of new grains in the shear bands. At these
temperatures shear bands almost completely consisted of small
equiaxed grains that most probably rose due to development of
dynamic (or post-dynamic) recrystallization in the most severely
deformed areas. Increase in the width of the shear bands results in
increase of the volume fraction involved in shear deformation and
therefore recrystallization and consumption of initial coarse grains.
Some recrystallized grains contained annealing twins. Inside the
initial coarse grains formation of elongated subgrains was observed.
Increase in deformation temperature to 1000 1C and 1100 1C
(Fig. 4e and f) resulted in formation of almost completely recrys-
tallized structure. The size of the recrystallized grains increased
noticeably with temperature. Recrystallized grains contained
numerous annealing twins.
Further insight into the microstructure of the shear bands at the
lowest temperatures (600 and 700 1C) was also obtained from TEM
Fig. 6. The dependence of volume fraction of recrystallized grains and their mean
analysis (Fig. 5). The microstructure of the shear bands arising during
size on temperature in specimens of CoCrFeNiMn alloy strained to E1.4 at strain
strain at 600 1C consisted of grains or subgrains with clear thin rate of 10  3 s  1. Grain size for the interval 800–1100 1C was determined by EBSD,
boundaries and a mean size of  0.2 mm (Fig. 5a). It is worth to note for 700 1C grain size was determined both by TEM and EBSD and for 600 1C the size
that a lot of grains are dislocation-free or contains only few dislocations. of grain was determined by TEM solely.
Appearance of new small grains and rather low dislocation density
suggests development of dynamic recrystallization in the shear bands. changes in orientation are observed in these grains. However, a
Increase in temperature to 700 1C resulted in the formation quantita- pronounced bulging of the initial grain boundaries is also detected
tively similar microstructure however the size of the recrystallized (higher magnification insert in Fig. 7a). Due to a strain localization
grains was found to be increased to  0.6 μm (Fig. 5b). effect and high internal stresses the initial boundaries tend to become
The quantitative evaluation of the effect of deformation tempera- poorly resolved by EBSD analysis after moderate strain of E0.7 at
ture on the volume fraction and the size of recrystallized grains is 700 1C (Fig. 7b). However at higher magnification fine recrystallized
shown in Fig. 6. The dependence of the volume fraction of recrys- grains were revealed in the initial grain boundaries. These grains
tallized grains on temperature can be described with a usual S-curve obviously developed into a necklace-like structure observed in Fig. 4b.
that is typical for the most saturable processes. Before and after rapid During deformation at 1000 1C recrystallized grains are seen
increase in the interval of 700–1000 1C from 15.1% to 92.7%, the after strain of E0.3 (Fig. 7c). They are located mainly along the
volume fraction of recrystallized grains changes slower both in low- initial grain boundaries and in triple junctions. Most of the initial
temperature (4.5% at 600 1C) and high-temperature (95.5% at 1100 1C) grain boundaries are bulged. The substructure development in
intervals. On the other hand, two-stage dependence of recrystallized initial coarse grains is relatively weak. At moderate strain (Fig. 7d),
grain size on deformation temperature is observed. At the first stage bands of recrystallized grains around cores of the initial grains are
the grain size increases slowly from 0.2 mm to 2.5 μm in the interval observed. Recrystallized grains contain many annealing twins.
of deformation temperatures 600–900 1C. At the second stage corre- There is still no pronounced development of substructure inside
sponding to higher temperatures, rapid increase of grain size to the remnants of the initial grains. Finally, at strain of E1.4, the
11.9 mm at 1000 1C and to 40.4 mm at 1100 1C is observed. widening of the recrystallized structure bands resulted in almost
In addition to the foregoing data, the microstructure evolution of complete consuming of the initial grains (Fig. 5e).
the CoCrFeNiMn alloy was studied during compressive testing at The volume fraction of the recrystallized grains and their size as a
700 1C and 1000 1C and 10  3 s  1. The microstructure of the alloy after function of strain at 1000 1C is shown in Fig. 8. The kinetics of
compression to different strains is shown in Fig. 7. At small strain at recrystallization development can also be described by S-curve. The
700 1C initial grains flattened in the metal flow direction; slight local fraction of recrystallized grains increases to only 2.2% at strain of E0.3
N.D. Stepanov et al. / Materials Science & Engineering A 636 (2015) 188–195 193

Fig. 7. EBSD IPF maps of the CoCrFeNiMn alloy after compression at strain rate of 10  3 s  1 and at temperatures of 700 1C (a, b) and 1000 1C (c, d) to true strain E0.3 (a, c),
E0.7 (b) and E 0.9 (d).

Fig. 8. The volume fraction of recrystallized grains and their mean size as a
function of strain at temperature of 1000 1C and strain rate of 10  3 s  1.

and then increases rapidly to 92.7% at E1.4 of strain. Grain size


increases almost linearly from 4.5 mm to 11.9 mm after strain of E0.3
and E1.4, respectively. Fig. 9. The relationship of flow stress at apparent steady state corresponding to
e 0.7 in log (σ) vs grain size.

4. Discussion reсent overview was published in [34]). Basically there are two types
of dynamic recrystallization which can be distinguished by the
The development of recrystallization during deformation was characteristics of deformation behavior and microstructure evolution
described comprehensively for various metals and alloys (one of the under hot/warm working: (i) discontinuous dynamic recrystallization
194 N.D. Stepanov et al. / Materials Science & Engineering A 636 (2015) 188–195

(dDRX) that is associated with the formation of nuclei via bulging approximately 0.7 or 0.4 for dDRX or cDRX, respectively [34].
mechanism which then grow out consuming the deformed matrix Meanwhile the relationship of flow stress vs grain size in the
thereby providing a grain structure with decreased dislocation present case can be approximated by a straight line with the
density [35,36] and (ii) continuous dynamic recrystallization (cDRX) constant slope thereby indicating the invariance of the recrystalli-
related to the formation of a stable network of low-angle boundaries zation mechanisms (Fig. 9).
followed by their gradual transformation into high-angle grain The character of microstructural response to warm deformation
boundaries upon straining [35,37]. The same material can demon- also shows quite obviously that the microstructure evolution in the
strate either dDRX or cDRX under different deformation conditions. interval 600–800 1C is mainly associated with dDRX. Intensive bulging
Usually the transition of the mechanism of dynamic recrystallization of the initial grain boundaries in the beginning of deformation (Fig. 7a)
is observed with decrease in deformation temperature e.g. [34,38]. and then formation of new small grains (Figs. 5 and 7b) clearly
The results of the mechanical behavior (Fig. 1) and the activation indicate the development of dynamic recrystallization via discontin-
energy analysis (Fig. 2b) can suggest the presence of two temperature uous mechanism. However there is an apparent difference distin-
intervals which can be assigned as hot (above 800 1C) and warm guishing warm deformation from hot one; at the lower temperatures
(below 800 1C) intervals. In the hot deformation interval the CoCrFe- microstructure evolution is associated with formation of the shear
NiMn alloy demonstrates typical of DRX stress–strain curves with the bands (Fig. 4a and b). This effect is reflected in deformation curves as
evident steady state flow stage while deformation in the warm multiple serrations. The intensity of shear deformation decreases
interval resulted in continuous strengthening of the alloy. Besides the while the volume fraction of the material involved in the shear
microstructure evolution during deformation within these two straining decreases with temperature. Local heating in adiabatic shear
intervals is controlled by processes with noticeably different activa- bands can result in formation of very small recrystallized grains [43].
tion energies (291 and 213 kJ/mol for the hot and warm intervals, Therefore it can be suggested that the influence of dislocation
respectively). The value of Q for the hot deformation interval is activity increases at lower temperature thereby decreasing activation
comparable with the activation energy (321.7 kJ/mol) for grain energy of the processes and increasing strain hardening ability. How-
growth in the same alloy [24]. The latter process can be controlled ever dDRX remains the main mechanism controlling the microstruc-
in turn by diffusion of Ni which was found to be the slowest element ture evolution of CoCrFeNiMn in the whole investigated temperature
in CoCrFeMn0.5Ni alloy (the activation energy of this processes was interval 600–1100 1C. The occurrence of dDRX in the CoCrFeNiMn
found to be 317.5 kJ/mol) [27]. The obtained value of activation alloy during hot and warm working instead of cDRX agrees well with
energy for hot deformation also demonstrates reasonable agreement low stacking fault energy of the alloy [28,35].
with activation energy value of 330 kJ/mol determined during tensile
testing of the same alloy at temperature of Z750 1C and strain rate
of 410  5 s  1 [17]. 5. Conclusions
These findings match pretty well with the microstructural
analysis. Deformation at temperatures above 800 1C is associated The microstructure evolution in high-entropy alloy CoCrFeNiMn
with the intensive bulging of initial grain boundaries (Fig. 7b), during uniaxial compression to a height reduction of 75% corre-
formation of small grains along the initial grain boundaries and sponding to a true strain of E1.4 in the temperature interval 600–
gradual consuming of the “old” deformed grains by new grains 1100 1C was studied. The alloy shows different mechanical behavior
(Figs. 4d–f and 7с and d) thereby suggesting development of dDRX and the activation energy of deformation in warm (below 800 1C)
in the alloy during hot deformation. In some metals and alloys and hot (above 800 1C) temperature intervals. Microstructure evolu-
(including CoCrFeNiMn) migrating grain boundaries can leave tion at all studied temperatures was found to be accompanied by
twins behind [25,39]; therefore the presence of annealing twins discontinuous dynamic recrystallization (dDRX). However during the
in the majority of grains is also a mark of dDRX [40]. hot deformation a classical dDRX associated with nucleation of new
The kinetics of microstructure evolution at 1000 1C in terms of grains via migration of the initial grain boundaries and further
the fraction of recrystallized grains having a classical S-curve is growth of nuclei was observed. In the warm interval dDRX was
also typical of dDRX (Fig. 8). However the increase in the size of mainly occurred in adiabatic shear bands. The volume fraction of the
recrystallized grains during deformation (Fig. 8) is less common recrystallized structure at 600 and 1000 1C was 0.085 and 0.95,
for dDRX. Fluctuation in the size of recrystallized grains can be respectively. The recrystallized grain size was found to be 0.2 and
found in case if stress–strain curve exhibits several peaks [41]. In 40.4 μm for 600 and 1100 1C, respectively.
current study, we have not observed distinguished peaks on
stress–strain curves (Fig. 1). However, it might be supposed that
due to sluggish diffusion in the studied HEA [27] the strain interval Acknowledgment
of each peak is quite prolonged. As the maximum strain during
compression testing in the current study is relatively low (true The authors gratefully acknowledge the financial support from
strain of E1.4), one can suggest that the observed grain growth the Russian Scientific Foundation Grant no. 14-19-01104. The authors
can be associated with one half-cycle of grain growth/refinement. are grateful to the personnel of the Joint Research Centre, Belgorod
As it was mentioned above lower activation energy in the State University, for their assistance with the instrumental analysis.
warm interval of deformation (Fig. 2b) and continuous strength-
ening during compression (Fig. 1) could suggest change in the References
main mechanism that controls the microstructure evolution with
decreasing temperature. One of the possible changes can be [1] J.-W. Yeh, S.-K. Chen, S.-J. Lin, J.-Y. Gan, T.-S. Chin, T.-T. Shun, C.-H. Tsau,
associated with the transition of the dynamic recrystallization S.-Y. Chang, Adv. Eng. Mater. 6 (8) (2004) 299–303.
mechanism (dDRX-cDRX) that was observed in a numerous [2] Y. Zhang, T.T. Zuo, Z. Tang, M.C. Gao, K.A. Dahmen, P.K. Liaw, Z.P. Lu, Prog.
Mater. Sci. 61 (2014) 1–93.
investigations for various metals and alloys [34,38,42]. For each
[3] J.W. Yeh, JOM 35 (12) (2013) 1759–1771.
type of DRX (either dDRX or cDRX) the size of recrystallized grains [4] B. Cantor, I.T.H. Chang, P. Knight, A.J.B. Vincent, Mater. Sci. Eng. A 375–377
(D) at the steady-state flow can be expressed by a power law (2004) 213–218.
function of the flow stress (σs) as follows: σs ¼ KD  N, where K and [5] M.S. Lucas, G.B. Wilks, L. Mauger, J.A. Munoz, O.N. Senkov, E. Michel,
J. Horwath, S.L. Semiatin, M.B. Stone, D.L. Abernathy, E. Karapetova, Appl.
N are constants [35]. However grain size exponent N being Phys. Lett. 100 (2012) 251907.
dependent on the deformation conditions takes on the values of [6] F. Otto, Y. Yang, H. Bei, E.P. George, Acta Mater. 61 (2013) 2628–2638.
N.D. Stepanov et al. / Materials Science & Engineering A 636 (2015) 188–195 195

[7] W.-R. Wang, W.-L. Wang, S.-C. Wang, Y.-C. Tsai, C.-H. Lai, J.-W. Yeh, Inter- [24] W.H. Liu, Y. Wu, J.Y. He, T.G. Nieh, Z.P. Lu, Scr. Mater. 68 (7) (2013) 526–529.
metallics 26 (2012) 44–51. [25] F. Otto, N.L. Hanold, E.P. George, Intermetallics 54 (2014) 39–48.
[8] J.Y. He, W.H. Liu, H. Wang, Y. Wu, X.J. Liu, T.G. Nieh, Z.P. Lu, Acta Mater. 62 [26] Z. Wu, H. Bei, F. Otto, G.M. Pharr, E.P. George, Intermetallics 46 (2014) 131–140.
(2014) 105–113. [27] K.Y. Tsai, M.H. Tsai, J.W. Yeh, Acta Mater. 61 (2013) 4887–4897.
[9] G.A. Salishchev, M.A. Tikhonovsky, D.G. Shaysultanov, N.D. Stepanov, A. [28] A.J. Zaddach, C. Niu, C.C. Koch, D.L. Irving, JOM 65 (2013) 1780–1789.
V. Kuznetsov, I.V. Kolodiy, A.S. Tortika, O.N. Senkov, J. Alloy. Compd. 591 [29] G.E. Dieter, Mechanical Metallurgy, third ed., McGraw-Hill, Inc., New York,
(2014) 11–21. 1986.
[10] N.D. Stepanov, D.G. Shaysultanov, G.A. Salishchev, M.A. Tikhonovsky, [30] H.J. McQueen, J.J. Jonas, Treatise on Materials Science and Technology, Vol. 6:
E.E. Oleynik, A.S. Tortika, O.N. Senkov, J. Alloy. Compd. 628 (2015) 170–185. Plastic Deformation of Materials, Academic Press, New York (1975) 393–493.
[11] M.-H. Chuang, M.-H. Tsai, W.-R. Wang, S.-J. Lin, J.-W. Yeh, Acta Mater. 59 (2011) [31] J.J. Jonas, C.M. Sellars, W.J.Mc.G. Tegart, Met. Rev. 14 (1969) 1–24.
6308–6317. [32] G. Laplanche, P. Gadaud, O. Horst, F. Otto, G. Eggeler, E.P. George, J. Alloy.
[12] Y.L. Chou, J.W. Yeh, H.S. Shih, Corros. Sci. 52 (2010) 2571–2581. Compd. 628 (2015) 348–353.
[13] W.H. Liu, J.Y. He, H.L. Huang, H. Wang, Z.P. Lu, C.T. Liu, Intermetallics 60 (2015) [33] A. Belyakov, H. Miura, T. Sakai, Mater. Sci. Eng. A255 (1998) 139–147.
1–8. [34] T. Sakai, A. Belyakov, R. Kaibyshev, H. Muira, J.J. Jonas, Prog. Mater. Sci. 60
[14] A. Gali, E.P. George, Intermetallics 39 (2013) 74–78. (2014) 130–207.
[15] F. Otto, A. Dlouhyґ, Ch. Somsen, H. Bei, G. Eggeler, E.P. George, Acta Mater. 61 [35] F. Humphreys, M. Hatherly, Recrystallization and Related Annealing Phenom-
(2013) 5743–5755. ena, second ed., Elsevier, Oxford, 2004.
[16] B. Gludovatz, A. Hohenwarter, D. Catoor, E.H. Chang, E.P. George, R.O. Ritchie, [36] T. Sakai, J.J. Jonas, Acta Metall. 32 (2) (1984) 189–209.
Science 345 (2014) 1153–1158. [37] N. Dudova, A. Belyakov, T. Sakai, R. Kaibyshev, Acta Mater. 58 (2010)
[17] J.Y. He, C. Zhu, D.Q. Zhou, W.H. Liu, T.G. Nieh, Z.P. Lu, Intermetallics 55 (2014) 3624–3632.
9–14. [38] A. Belyakov, S. Zherebstov, G. Salishchev, Mater. Sci. Eng. A 628 (2015)
[18] Z. Wu, H. Bei, G.M. Pharr, E.P. George, Acta Mater. 81 (2014) 428–441.
104–109.
[19] N. Stepanov, M. Tikhonovsky, N. Yurchenko, D. Zyabkin, M. Klimova,
[39] S. Mahajan, Scr. Mater. 68 (2013) 95–99.
S. Zherebtsov, A. Efimov, G. Salishchev, Intermetallics 59 (2015) 8–17.
[40] M. Tikhonova, A. Belyakov, R. Kaibyshev, Mater. Sci. Eng. A 564 (2013)
[20] A.V. Kuznetsov, D.G. Shaysultanov, N.D. Stepanov, G.A. Salishchev,
413–422.
O.N. Senkov, Mater. Sci. Eng. A 533 (2012) 107–118.
[41] A. Dehghan-Manshadi, P.D. Hodgson, ISIJ Int. 47 (12) (2007) 1799–1803.
[21] A.V. Kuznetsov, D.G. Shaysultanov, N.D. Stepanov, G.A. Salishchev,
[42] H. Beladi, P. Cizek, P.D. Hodgson, Metall. Mater. Trans. A 40 (5) (2009)
O.N. Senkov, Mater. Sci. Forum 735 (2013) 146–151.
1175–1189.
[22] D.G. Shaysultanov, N.D. Stepanov, A.V. Kuznetsov, G.A. Salishchev,
[43] B. Wang, Z. Liu, B. Wang, S. Zhao, J. Sun, Mater. Sci. Eng. A 611 (2014) 100–107.
O.N. Senkov, JOM 35 (2013) 1815–1828.
[23] P.P. Bhattacharjee, G.D. Sathiaraj, M. Zaid, J.R. Gatti, C. Lee, C.W. Tsai, J.W. Yeh,
J. Alloy. Compd. 587 (2014) 544–552.

You might also like