ArticlePDF Available

Abstract and Figures

Comparison on the kinetics of two different phase transformations, including phase transformation after deformation and phase transformation during deformation (i.e. dynamic transformation, DT), reveals a new discovery that the transformation kinetics can be significantly enhanced in DT even under low driving forces. DT enables continuous generation of defects (e.g. dislocations) near the phase boundary, which can act as short-circuiting diffusion paths for atoms. The diffusivity of atoms is enhanced and the activation energy for the atom jump across the phase boundary is lowered under stress during DT, resulting in more pronounced grain growth as well as accelerated transformation kinetics.
This content is subject to copyright. Terms and conditions apply.
MATER. RES. LETT.
2018, VOL. 6, NO. 11, 641–647
https://doi.org/10.1080/21663831.2018.1527787
ORIGINAL REPORT
Deformation-assisted diffusion for the enhanced kinetics of dynamic phase
transformation
Lijia Zhaoa,b, Nokeun Parka,c, Yanzhong Tiana,d,e, Akinobu Shibata a,dand Nobuhiro Tsuji a,d
aDepartment of Materials Science and Engineering, Kyoto University, Kyoto, Japan; bAdvanced Steel Processing and Products Research Center,
Department of Metallurgical and Materials Engineering, Colorado School of Mines, Golden, CO, USA; cSchool of Materials Science and
Engineering, Yeungnam University, Gyeongsan, Republic of Korea; dElements Strategy Initiative for Structural Materials (ESISM), Kyoto
University, Kyoto, Japan; eShenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences,
Shenyang, People’s Republic of China
ABSTRACT
Comparison on the kinetics of two different phase transformations, including phase transformation
after deformation and phase transformation during deformation (i.e. dynamic transformation, DT),
reveals a new discovery that the transformation kinetics can be significantly enhanced in DT even
under low driving forces. DT enables continuous generation of defects (e.g. dislocations) near the
phase boundary, which can act as short-circuiting diffusion paths for atoms. The diffusivity of atoms
is enhanced and the activation energy for the atom jump across the phase boundary is lowered under
stress during DT, resulting in more pronounced grain growth as well as accelerated transformation
kinetics.
IMPACT STATEMENT
Deformation-enhanced grain growth is revealed in dynamic phase transformation, which will pro-
mote microstructure and property design of structural materials where phase transformations
occur.
ARTICLE HISTORY
Received 4 June 2018
KEYWORDS
Dynamic phase
transformation; deformation;
diffusion; kinetics; grain
growth
Solid-state phase transformation commonly exists in
materialsandplaysamajorroleincontrollingtheir
microstructures and properties [15]. Diusional phase
transformation generally incorporates nucleation and
grain growth. During the nucleation, a new interface
separating the product phase from the parent phase is
generated. The interface migrates into the surrounding
parent phase through jumps of atoms across the phase
boundary (growth of nuclei) [6,7]. As in a typical diu-
sional phase transformation, e.g. austenite (face-centered
cubic, FCC) to ferrite (body-centered cubic, BCC) trans-
formation in steels, interstitial carbon atoms experience
partitioning from ferrite to austenite by long-range dif-
fusion. Simultaneously, individual jumps of iron and
CONTACT Lijia Zhao zhaolj618@gmail.com Department of Materials Science and Engineering, Kyoto University, Yoshida-honmachi, Sakyo-ku, Kyoto
606-8501, Japan, Advanced Steel Processing and Products Research Center, Department of Metallurgical and Materials Engineering, Colorado School of Mines,
Golden, CO 80401, USA
Supplemental data for this article can be accessed here. https://doi.org/10.1080/21663831.2018.1527787
substitutional alloying elements across the interface lead
tothereconstructionoftheBCCcrystalfromtheparent
FCC austenite.
The kinetics of diusional phase transformation can
be inuenced by plastic deformation, which is often
applied during processing of materials. The diusivity of
either interstitial or substitutional atoms could be aected
by lattice defects (dislocation, vacancy, etc.) introduced
by plastic deformation, since the defects may lower the
planar density of atoms and provide extra free volumes
that can be favorable pathways for atomic diusion [8].
Basically, there are two dierent phase transformations
in high-temperature processes including plastic defor-
mation: (1) phase transformation after deformation of
© 2018 The Author(s). Published by Informa UK Limited, trading as Taylor& Francis Group.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use,
distribution, and reproduction in any medium, provided the original work is properly cited.
642 L. ZHAO ET AL.
the parent phase (i.e. static transformation, ST) [9,10]
and (2) phase transformation during deformation (i.e.
dynamic transformation, DT) [4,5,1115]. The deforma-
tion applied before or during transformation determines
how and where the defects are introduced, and therefore
could lead to dierent scenarios of transformation kinet-
ics aected by nucleation density and growth rate related
to atomic diusion.
In this work, the eect of plastic deformation on the
kinetics of ferritic transformation in a low-carbon steel
was compared between the route-ST (i.e. phase transfor-
mation after deformation) and the route-DT (i.e. phase
transformation during deformation). A surprising dis-
covery was found: the diusivity of atoms as well as the
transformation kinetics was signicantly enhanced even
underlowdrivingforcesintheroute-DT.Thisstudygives
a new perspective to tailor the microstructures in a wide
range of materials where phase transformations occur.
The material used in this study is an Fe–10Ni–0.1C
alloy (C: 0.111, Ni: 10.08, Mn: 0.01, P: 0.001, Si:
0.006, Al: 0.33, S: 0.0017, Fe: bal. (wt.%)). Para-
equilibrium temperature (Ap3) of the Fe–10Ni–0.1C
alloy is 583°C calculated by Thermo-Calc software
(time–temperature–transformation diagram is shown in
supplementary Figure S1). Cylindrical specimens with a
height of 12 mm and a diameter of 8 mm were machined
from a homogenized plate and used for simulating
the thermomechanical controlled process (TMCP) on
Thermecmastor-Z (Fuji Electronic Industrial Co. Ltd.).
All the specimens were austenitized at 1000°C for 300s.
In the route-ST, the austenitized specimens were cooled
to 520°C at a rate of 30°Cs1, held for 60 s at 520°C to
homogenize the temperature in the specimens, uniaxially
compressed to a strain of 0.916 at a strain rate of 10°s1
and then isothermally held at 520°C for dierent periods
of time followed by water quenching. In the route-DT,
the austenitized specimens were cooled to 520°C at a rate
of 30°C s1, held for 60 s at 520°C, and then uniaxially
compressed to a strain of 0.916 at dierent strain rates
from 10° s1to 103s1followed by water quenching.
Microstructures at the sections parallel to the compres-
sion axis were characterized by optical microscopy (OM),
and a eld-emission type scanning electron microscope
(FE-SEM, FEI XL30S FEG) equipped with the electron
back-scattering diraction (EBSD) system. The point-
counting method was used for measuring the volume
fractions of ferrite on OM images. The area measured to
get the ferrite fraction in the present study was 33,232.1
(μm)2for each condition.
The comparison on transformation kinetics in dif-
ferent processing routes is summarized in Figure 1(a),
and corresponding microstructures are shown in Figure
1(b–e). More relevant microstructures can be seen in
supplementary Figures S2, S3 and S4. Here, the transfor-
mation time incorporates deformation time and isother-
mal holding time (after deformation). Compared to the
sluggish kinetics of static ferrite transformation without
deformation (indicated as ‘ST without Def.’), the kinetics
of ferrite transformation in the route-ST is signicantly
accelerated by the deformation of austenite which could
introduce a high density of defects as the nucleation sites
for the subsequent ferrite transformation [9,10]. The vol-
ume fraction of ferrite in the route-ST is 0.5% at 9 s (i.e.
1 s of deformation time and 8s of holding time after
the deformation, Figure 1(b)), and reaches 79% at 916 s
(Figure 1(c)). In the route-DT, the volume fraction of
ferrite is around 31% at a transformation time of 9 s
(corresponding to the time required for the compres-
sion deformation to a true strain of 0.916 at a strain
rate of 101s1,Figure1(d)), and the fraction of fer-
rite dramatically increases to 86.2% at 916s (the time for
deformation at a strain rate of 103s1,Figure1(e)). It
is of great interest that within the same transformation
time, the volume fraction of ferrite in the route-DT is
much higher than that in the route-ST, suggesting that
the kinetics is signicantly accelerated when the trans-
formation occurs during deformation compared to that
occurring after deformation.
The kinetics can be principally aected by the driving
forceforphasetransformation|Gαγ |,whichisgivenby
|Gαγ |=|Gchem
αγ |+|Gdef
γ|
−|Gdef
αγ (fα)+Gint
αγ (fα)|(1)
where Gchem
αγ (<0) is the chemical Gibbs energy dif-
ference between ferrite (α)andaustenite(γ)atthe
same temperature, Gdef
γ(<0) is the stored energy in
deformed γ,Gdef
αγ (fα)(>0)is the summation of elas-
tic and plastic mist energies between αand γat the
interface, and Gint
αγ (fα)(>0)isthefreeenergyofα/γ
interface. Gdef
αγ (fα)+Gint
αγ (fα)includes the area den-
sity of interfaces per volume and is a function of ferrite
fraction. The stored energy in austenite Gdef
γcan be cal-
culatedbythefollowingformula[16], assuming that the
stored energy is mostly due to dislocations accumulated
in the crystal during the plastic deformation,
Gdef
γ=μb2ρ(2)
where μis the shear modulus, ρis the dislocation
density, and bis the magnitude of the Burgers vector.
The relation between ow stress σof deformed austen-
ite and dislocation density follows the Bailey–Hirsch
MATER. RES. LETT. 643
Figure 1. (a) Variation in volume fraction of ferrite with transformation time at 520°C. (b–e) OM images of the corresponding microstruc-
tures obtained in the two routes (ST and DT). ‘F’ and ‘M’ indicate ferrite (light area) and martensite (dark area), respectively. Compression
axis is parallel to the vertical direction in the images.
equation [17],
σ=Mαμbρ1/2(3)
where Mis the Taylor factor of polycrystalline austenite,
and αis a numerical constant. Thus, combining Equa-
tions (2) and (3), the stored energy in austenite could be
described as
Gdef
γ=σ2M2α2(4)
In Equation (4), μM2α2isaconstant,sothatGdef
γis
proportional to the square of the ow stress σ2.Forthe
steel used in the present study, the values of M,μand α
are 3.08, 6.1 ×1010 Jm
3and 0.2, respectively [18].
Figure 2(a)showsthetruestress-truestraincurves
of the specimens deformed at 520°C. The stored ener-
gies in γcalculated from Equation (4) are plotted against
the deformation time in Figure 2(b). It is clear that the
Gdef
γis lower in the route-DT than that in the route-
ST.Itshouldbenotedthatintheroute-DT,austenite
644 L. ZHAO ET AL.
Figure 2. (a) True stress-true strain curves of specimens
deformed to a strain of 0.916 in the route-ST (at a strain rate of
10° s1) and the route-DT (at strain rates from 101to 103s1)
at 520°C. (b) Stored energy in austenite (γ)inthetworoutes,
calculated by Equation (4).
to ferrite transformation occurred during the deforma-
tion. The transformed ferrite was further after its nucle-
ation, which means that less plastic strain than the total
strain was applied to γ. The actual stored energy in
deformed γis even less that that calculated by Equation
(4). Results of the comparison on all the parameters in
Equation (1) are summarized as follows: the chemical
driving force Gchem
αγ is equal in both routes at 520°C,
since it only depends on the temperature; the Gdef
γis
lower in the route-DT, as derived from Figure 2(b); the
term [Gdef
αγ (fα)+Gint
αγ (fα)] increases with increasing
the ferrite fraction [19], so it is higher in the route-DT
than in the route-ST according to Figure 1(a). Therefore,
the total driving force Gαγ should be smaller in the
route-DT.Thisisarathersurprisingconclusionthatthe
transformation kinetics in the route-DT is faster than that
in the route-ST even under lower driving forces.
Figure 3(a,b) shows grain average misorientation
(GAM) maps of microstructures obtained at a transfor-
mation time of 916 s in (a) the route-ST and (b) the route-
DT. The GAM map obtained by EBSD measurement can
be used to evaluate the degree of misorientation inside
each grain [2022]. A higher GAM value (dark red color)
indicates higher misorientation within the grain. The fer-
rite statically transformed after plastic deformation in
the route-ST incorporates mostly equiaxed grains sur-
rounded by high-angle boundaries (HABs). The ferrite
dynamically transformed in the route-DT incorporates
mostly elongated coarse grains containing a large amount
of low-angle boundaries (LABs). The average GAM val-
ues of ferrite transformed in the route-ST and route-DT
are 0.34° and 0.63°, respectively (Figure 3(c)), indicat-
ing a more deformed ferrite structure in the route-DT.
Figure 3(d) shows variations of the apparent nucleation
density of ferrite in the two routes. The apparent nucle-
ation density of ferrite was calculated by dividing the
number of ferrite grains by the total area of ferrite [23,24].
As shown in Figure 3(d), the apparent nucleation den-
sity decreases with increasing the transformation time in
both routes due to the growth of nucleated ferrite. Within
the same transformation time at 520°C, the apparent
nucleation density in the route-DT is much lower than
that in the route-ST. This is reasonable since the higher
driving force (due to higher dislocation density) in the
route-ST enhanced the nucleation of ferrite. According to
Figure 1(a), the ferrite fraction in the route-DT is always
higher than that in the route-ST. Therefore, it can be
concluded that the faster transformation kinetics in the
route-DT is mainly due to the enhanced growth of ferrite
during deformation.
The dierence in ferrite grain growth behavior
between the two routes is further discussed here. In the
present study, all the experiments were conducted at
520°C, which is below the para-equilibrium temperature
(583°C) of the steel, so the substitutional element Ni can-
notfullydiuseinausteniteatthislowtemperature.Since
the transformation kinetics of static transformation with-
out deformation is slow (shown as ‘ST without Def.’ in
Figure 1(a)), carbon could be partitioned between ferrite
and austenite and its long-range diusion is considered to
dominate the growth of ferrite. It is worth noting that the
interaction between Ni and carbon or moving interfaces
is negligible due to the low binding energies between
them [25,26]. Thus, possible segregation of Ni in the
interfaceandthesolutedrageect,ifany,isexpectedto
be small in the present study.
The diusion of carbon in austenite and ferrite can be
aected by plastic deformation in terms of two aspects:
strain and stress. In the route-ST, the strain of 0.916 is
totally applied on austenite at a high strain rate of 1 s1.
The static ferrite transformation occurs during isother-
mal holding after the deformation. In the route-DT at
lower strain rates (101to 103s1), on the other hand,
the plastic deformation is rstly applied on austenite at
MATER. RES. LETT. 645
Figure 3. (a) and (b) are GAM maps obtained by EBSD measurements of specimens at a transformation time of 916 s in the route-ST and
route-DT, respectively. Non-ferrite phases (i.e. martensite and/or retained austenite) are painted in black. LABs with misorientation of
2–15° and HABs with misorientation above 15° are drawn in red and blue lines, respectively. C.A. indicates the compression axis. (c) and
(d) show distributions of GAM and variations in apparent nucleation density of ferrite grains transformed in two different routes (route-ST
and route-DT), respectively.
early stages and then on dual phases after the formation
of ferrite. It is, therefore, reasonable to consider that the
density of the lattice defects (dislocations, etc.) in austen-
ite, which provide potential nucleation sites for ferrite,
is higher in the route-ST. Therefore, the nucleation den-
sity of ferrite was much higher in the route-ST than in
theroute-DTasshowninFigure3(d). Figure 4schemat-
ically illustrates microstructural evolutions in the two
routes. Grain boundaries of austenite are strong obstacles
for dislocation slips due to the crystallographic discon-
tinuity. It is also expected that various slip systems are
activated near grain boundaries for satisfying compati-
bility between neighboring grains. As a result, dislocation
densities near grain boundaries of austenite are expected
to be higher than that at grain interior (Figure 4(a)). In
the route-ST, ferrite grains prefer to nucleate near grain
boundaries of austenite (and along deformation bands, if
any: Figure 4(b)), and consume dislocations accumulated
nearby the grain boundaries (Figure 4(c)). The ferrite
grains could experience early impingement due to the
higher nucleation density and thus the growth of ferrite
is suppressed near the austenite grain boundaries in the
route-ST.Thenucleatedferritegrainscouldgrowintothe
grain interior of austenite with less nucleation sites and
impingement (Figure 4(c)). In the route-DT, on the other
hand, the austenite is gradually strained at a slower strain
rate, leading to the lower density of nucleation sites for
ferrite at early stages of transformation (Figure 4(d)). The
austeniteandferritearebothdeformedatlaterstages,and
dislocations are preferentially stored near austenite grain
boundaries as well as austenite/ferrite phase boundaries
(Figure 4(e)). The dislocations can act as eective pipes
along which atoms can diuse faster [27]. Therefore, the
diusion of carbon atoms can be enhanced through the
fast routes, and thus the migration rate of austenite/ferrite
phaseboundariesisincreased.Thisistheessentialdier-
ence between DT and ST in ferrite growth, i.e. there are
more dislocations continuously generated near the mov-
ing phase boundaries during DT, and the diusion of car-
bon in ferrite and austenite can be enhanced due to these
additional ‘express-ways’, resulting in more pronounced
grain growth (Figure 4(f)).
The stress may also aect the growth of ferrite grains.
For a single atom, the atom must pass through a ther-
mally activated state with an activation energy to achieve
an eective jump over the interface. It has been simulated
646 L. ZHAO ET AL.
Figure 4. Schematic illustrations of (a–c) static ferrite transformation from deformed austenite (route-ST) and (d–f ) dynamic ferrite
transformation (route-DT). Orange and blue lines depict austenite and ferrite grains, respectively. Black lines represent defects (mainly
dislocations). The areas where atom diffusion is enhanced are highlighted by green color in (e and f).
[8] that under a compressive stress, the presence of lat-
tice defects (vacancy, etc.) near austenite/ferrite inter-
faces could reduce the planar density of atoms, provid-
ing extra free volume at moving interfaces, and thereby
lowers the extent of local rearrangement necessary to
move the interface. Dierent from the route-ST where
the transformed ferrite is almost free of defects (Figures
3(a) and 4(c)), the ferrite in the route-DT is continu-
ously deformed during transformation (Figures 3(b) and
4(f)). The introduced lattice defects could enhance the
accommodation of jumping atoms over the phase bound-
ary under stress, and then increase the migration rate of
the phase boundary (i.e. growth rate of ferrite).
In summary, our ndings reveal a signicant enhance-
ment of transformation kinetics in dynamic phase
transformation even under low driving forces. This is
attributed to the enhanced diusivity of atoms in the
migration of phase boundary (i.e. growth of ferrite)
under strain and stress applied by continuous deforma-
tion. The strain-enhanced diusivity of atoms results pri-
marily from dislocation-assisted diusion through short-
circuiting paths (pipes) near austenite/ferrite interfaces.
Stress may lower the activation energy for atoms to jump
across phase boundaries. This discovery could lead to
new strategies for tailoring required microstructures and
properties in many materials where phase transformation
occurs.
Disclosure statement
No potential conict of interest was reported by the authors.
Funding
This study was nancially supported by the Elements Strategy
Initiative for Structural Materials (ESISM) and the Grant-in-
Aid for Scientic Research (S) (No.15H05767), both through
the Ministry of Education, Culture, Sports, Science and Tech-
nology (MEXT), Japan.
ORCID
Akinobu Shibata http://orcid.org/0000-0001-8577-6411
Nobuhiro Tsuji http://orcid.org/0000-0002-2132-1327
References
[1] Miyamoto G, Karube Y, Furuhara T. Formation of grain
boundary ferrite in eutectoid and hypereutectoid pearlitic
steels. Acta Mater. 2016;103:370–381.
[2] Oerman SE, Van Dijk NH, Sietsma J, et al. Grain nucle-
ation and growth during phase transformations. Science.
2002;298:1003–1005.
MATER. RES. LETT. 647
[3] JonesSJ,BhadeshiaHKDH.Kineticsofthesimultaneous
decomposition of austenite into several transformation
products. Acta Mater. 1997;45:2911–2920.
[4] Ghosh C, Aranas C, Jonas JJ. Dynamic transformation of
deformed austenite at temperatures above the Ae3.Prog
Mater Sci. 2016;82:151–233.
[5] Zhao LJ, Park N, Tian YZ, et al. Realizing ultrane
grained steel by simple hot deformation using dynamic
transformation and subsequent dynamic recrystallization
mechanisms. IOP Conf Series Mater Sci Eng. 2017;194:
012026.
[6] Hillert M. Critical limit for massive transformation. Met-
all Mater Trans A. 2002;33A:2299–2308.
[7] Liu YC, Sommer F, Mittemeijer EJ. The austenite–ferrite
transformation of ultralow-carbon Fe–C alloy; transi-
tion from diusion- to interface-controlled growth. Acta
Mater. 2006;54:3383–3393.
[8] Biglari M Jr, Mittemeijer EJ. Simulation of the massive
austenite–ferrite transformation under uniaxial loading.
Comput Mater Sci. 2013;77:214–223.
[9] Adachi Y, Wakita M, Beladi H, et al. The formation of
ultrane ferrite through static transformation in low car-
bon steels. Acta Mater. 2007;55:4925–4934.
[10] Hosseini SMK, Zaeri-Hanzaki A, Yue S. Eect of
austenite deformation in non-recrystallization region on
microstructure development in low-silicon content TRIP-
assisted steels. Mater Sci Eng A. 2014;618:63–70.
[11] Eghbali B, Abdollah-Zadeh A. Strain-induced trans-
formation in a low carbon microalloyed steel dur-
ing hot compression testing. Scr Mater. 2006;54:1205–
1209.
[12] Zhao LJ, Park N, Tian YZ, et al. Dynamic transformation
mechanism for producing ultrane grained steels. Adv
Eng Mater. 2018;20:1701016.
[13] Park N, Zhao L, Shibata A, et al. Dynamic ferrite
transformation behaviors in 6Ni-0.1 C steel. JOM.
2014;66:765–773.
[14] Zhao L, Park N, Tian Y, et al. Mechanism of dynamic for-
mation of ultrane ferrite grains during high temperature
processing in steel. Adv Eng Mater. 2017;19:1600778.
[15] Zhao L, Park N, Tian Y, et al. Novel thermomechanical
processing methods for achieving ultragrain renement
of low-carbon steel without heavy plastic deformation.
Mater Res Lett. 2017;5:61–68.
[16] HanlonDN,SietsmaJ,VanderZwaagS.Theeectofplas-
tic deformation of austenite on the kinetics of subsequent
ferrite formation. ISIJ Inter. 2001;41:1028–1036.
[17]BaileyJE,HirschPB.Therecrystallizationprocessin
some polycrystalline metals. Proc R Soc A. 1962;267:11–30.
[18] Ghosh C, Basabe VV, Jonas JJ, et al. The dynamic transfor-
mation of deformed austenite at temperatures above the
Ae3.ActaMater.2013;61:2348–2362.
[19] Liu YC, Sommer F, Mittemeijer EJ. Austenite–ferrite
transformation kinetics under uniaxial compressive stress
in Fe–2.96 at.% Ni alloy. Acta Mater. 2009;57:2858–2868.
[20] Zhao L, Park N, Tian Y, et al. Combination of dynamic
transformation and dynamic recrystallization for real-
izing ultrane-grained steels with superior mechanical
properties. Sci Rep. 2016;6:39127.
[21] TianY,GaoS,ZhaoL,etal.Remarkabletransitionsof
yield behavior and Lüders deformation in pure Cu by
changing grain sizes. Scr Mater. 2018;142:88–91.
[22] Zhao L, Park N, Tian Y, et al. Microstructural evolution
of ferrite grains during dynamic transformation in 10Ni-
0.1C steel. In: TMS Society editors. TMS 2014 Annual
Meeting Supplemental Proceedings; 2014 Feb 16–20; San
Diego, CA. Switzerland: Springer International Publish-
ers; 2014. p. 919–926.
[23] Beladi H, Timokhina IB, Mukherjee S, et al. Ultrane
ferrite formation through isothermal static phase trans-
formation. Acta Mater. 2011;59:4186–4196.
[24] Beladi H, Kelly GL, Shokouhi A, et al. The evolution
of ultrane ferrite formation through dynamic strain-
induced transformation. Mater Sci Eng A. 2004;371:343–
352.
[25]HutchinsonCR,ZurobHS,BrechetY.Thegrowthof
ferrite in Fe-C-X alloys: the role of thermodynamics, dif-
fusion, and interfacial conditions. Metall Mater Trans A.
2006;37A:1711–1720.
[26] Van Landehem HP, Langelier B, Gault B, et al. Inves-
tigation of solute/interphase interaction during ferrite
growth. Acta Mater. 2017;124:536–543.
[27] Cohen M. Self-diusion during plastic deformation.
Trans JIM. 1970;11:145–151.
... Those defects remain in the microstructure after deformation. In addition, the dislocationassisted diffusion through short-circuiting paths near α/β interfaces may enhance the β-to-α transformation [13]. Figure 5 shows that the acicular α phase precipitated and penetrated β grains more at 30% β deformation than at 0% and 15% β deformations under constant (α + β) forging ratios. ...
... Metals 2021, 11, x FOR PEER REVIEW 4 of 14 may enhance the β-to-α transformation [13]. Figure 5 shows that the acicular α phase precipitated and penetrated β grains more at 30% β deformation than at 0% and 15% β deformations under constant (α + β) forging ratios. ...
... However, recrystallization scarcely occurs during forging and cooling in the (α + β) region due to α precipitation at β gain boundaries even at a moderate cooling rate [14]. may enhance the β-to-α transformation [13]. Figure 5 shows that the acicular α phase precipitated and penetrated β grains more at 30% β deformation than at 0% and 15% β deformations under constant (α + β) forging ratios. ...
Article
Full-text available
The effect of grain-interior α precipitation on the β texture evolution of the near-β Ti-6246 alloy during through-transus forging was investigated in two-step sequential forgings. The microstructure and texture were analyzed using scanning electron microscopy, electron-backscatter diffraction, and X-ray diffraction. The previous β forging was performed at 1253 K at 0.01/s, while the subsequent forging in the (α + β) region was conducted at 1073 K at 0.01/s. The forging in the β region facilitated the penetration of the interior α phase into β grains and reduced the formation of grain boundary α. The {001} texture intensity increased during the forging in the single β region. By contrast, the increase in the {001} texture intensity was moderate at a lower temperature (1073 K) because the Schmid factor (SF) value of the {110}<111> slip system drastically decreased, but those of the {112}<111> and {123}<111> slip systems increased before α precipitation. During α precipitation for all β forging ratios, the {110}<111> slip system was activated, resulting in a lowering of the {001} texture intensity. The lower the forging temperature before interior α precipitation under a constant total forging ratio, the more the {001} texture intensity was suppressed in the final β texture, accompanied by interior α precipitation.
... This can also be seen from the Ta-Zr (Fig. 12 (e)) and Nb-Zr binary phase diagrams (Fig. S4). A similar phenomenon was reported in AlCoCrCuFeNi HEA, in which a Cu-rich phase emerged after annealing by segregation due to the high positive values (X = Fe, Cr, Co, Ni) [50]. Additionally, Zr atoms are much larger than Ti, Ta, and Nb atoms. ...
... Hence, the nanocuboidal structure disappeared in the samples after compression at ≥ 800°C. Conversely, compression at 600°C, which is below 677°C, retained the nanocuboidal structure, but due to the severe plastic deformation and stress-enhanced diffusion [50], the retained nanocuboidal structure exhibited distortion and coarsening ( Fig. 7 (a,b)). ...
Article
This study investigates the effect of annealing treatment on the phase transformation and mechanical properties of the equiatomic TiZrNbTa MEA from room temperature to 1200 °C. After annealing at 1200 °C for 24h, the single solid-solution body-centred cubic (BCC) phase in the as-cast Ta25Zr25Nb25Ti25 transformed into an extremely high number density (∼10³/μm²) of Ta–Nb-rich BCC nanocuboidal phase (28 ± 10 nm square, BCC1) and a nanostrip-like Zr-rich BCC phase (3 ± 2 nm thick, BCC2). The phase separation from BCC to BCC1 and BCC2 arises from two primary reasons: (i) the high positive mixing enthalpy of both Ta–Zr and Nb–Zr (strong tendency to separation between each pair), and (ii) the 3–4 orders of magnitude higher mobility of Zr than Ta, Nb and Ti in these MEAs (kinetically driven). Detailed CALPHAD simulations of phase formation in this MEA agreed with experiments and provided insightful phase transformation details. The calculated diffusion distance of Zr (∼4.1 nm) from the CALPHAD data corresponds to the measured Zr-rich nanostrip thickness (3 ± 2 nm). The nanocuboidal BCC1-BCC2 structure exhibited 112<111>-type of twinning deformation under compression at room temperature. The Ta25Zr25Nb25Ti25 MEA retained yield strength of ∼410 MPa at 1000 °C and ∼210 MPa at 1200 °C. The phase transformation during cooling after annealing and the microstructural evolution during compression at temperatures from 600 °C to 1200 °C were characterized and discussed in detail.
... Amounts of defects (e.g. dislocations) caused by severe deformation tend to interact with each other especially near grain boundaries, presumably acting as short-circuiting diffusion paths owing to lattice expands [37]. Therefore, Q calculated from electrical conductivity decreases with the increasing reduction rate. ...
Article
In this study, microstructures, mechanical properties, and electrical conductivities of powder metallurgy (PM) Cu-0.28Cr-0.19 Mg alloy after cold-rolling and aging were investigated in detail. PM Cu-Cr-Mg alloy exhibits a combination of both higher mechanical properties and superior electrical conductivity compared to the corresponding casting Cu-Cr-X alloys. Such excellent performances are closely related to the more uniform and finer microstructures generated in PM process. Strengthening from second phase precipitation, grain boundary, and dislocation in the present PM Cu-Cr-Mg alloy were calculated to be approximately 43.65, 24.50, and 23.12% of yield strength, respectively. The grain boundary strengthening is the main reason for the higher strength of the present PM Cu-Cr-Mg alloy and is approximately 325% higher than that of the corresponding casting Cu-Cr-Mg alloy. The gradually decreased activation energy of PM Cu-Cr-Mg alloy with increasing reduction rate suggests that deformation significantly accelerates the precipitation behavior during the subsequent aging process.
... The phase transformation kinetic is significantly accelerated when the transformation occurs under continuous deformation. The lattice defects can enhance the atom jumps (diffusivity) through the interphase boundaries, and consequently increase the migration rate of the austenite/ferrite boundaries [72,73]. Therefore, the increase in the ferrite phase was caused by the strain accumulation in both phases since the austenite / ferrite transformation is diffusion-controlled. ...
Article
Full-text available
Duplex stainless steels can be used for applications in the nuclear industry endorsing its choice as an external shield for Type B packages. The microstructure, crystallographic texture, and the strain hardening behavior of UNS S32304 duplex stainless steel was investigated in hot tensile tests by Electron Backscattering Diffraction technique. Interrupted hot tensile tests with strain rates ε˙ = 10⁻² s⁻¹ and ε˙ = 10⁻¹ s⁻¹ were carried out at 700 °C using a Thermomechanical Physical simulator. The α parameter of gamma distribution was associated for the first time with strain in ferrite and austenite phases through the Grain Orientation Spread parameter (GOS). The ferrite phase did not show the dynamic equilibrium between the stored energy and recovery rates. The average GOS in the ferrite phase was larger for ε˙ = 10⁻¹ s⁻¹ due to a smaller time available for dislocation annihilation and rearrangement in dynamic recovery. The grain rotation sequence <100> → <100> - <101> → <101> was found in ferrite phase. In the austenite phase, the dynamic recrystallization was not observed and the <111> and <100> fiber textures parallel to the tensile direction were strengthened. The austenite → ferrite phase transformation occurred during the hot tensile tests and showed the Kurdjumov-Sachs orientation relationship. The modified Crussard-Jaoul analysis showed six-strain hardening stages which were associated with the average GOS and austenite → ferrite phase transformation.
... Cementite particles and ferrite grain boundaries in the initial cold-rolled sample (see the inset of Fig. 1(a)) could also provide nucleation sites for austenite formation. As a result of the plenty of potential nucleation sites, the number density of austenite grains in the FA sample, calculated from EBSD results by dividing the number of austenite grains by the total area [26], is about 8*10 12 /m 2 , which is~3 times of that in the ART sample. Due to the reduced time for grain growth during FA, RA grains were refined to~185nm ( Fig. 1(c 2 )), which should have a positive effect on the thermal stability of austenite grains. ...
Article
Full-text available
Austenite reversion treatment (ART) is usually used to obtain retained austenite (RA) in medium Mn steel (MMS). In this contribution, we proposed a flash-annealing strategy to produce MMS, which allows us to efficiently obtain a large amount of RA in MMS via the “explosive” nucleation of austenite and the fast C partitioning. Furthermore, flash-annealing can also render a heterogeneous ferritic matrix consisting of recrystallized/non-recrystallized ferrite grains, leading to heterogeneous deformation-induced hardening. As compared to the conventional ART processed MMS, the unique heterogeneous microstructure in the flash-annealed MMS leads to a higher yield and ultimate tensile strength but a marginal ductility loss. Flash-annealing provides an alternative route to tailor microstructure and optimize the mechanical performance of steels.
... The accumulation of strain (i.e., strain energy) during hot rolling intensifies the diffusion of alloying elements (through pipe diffusion) and provides the driving force for the DIF transformation. [39,40] During cold rolling, the defects generated by deformation act as the preferred nucleation sites for ferrite formation through structure-sensitive diffusion. The carbon atoms diffuse during cold rolling by dislocation-assisted pipe diffusion [41] near the austenite/ferrite interfaces. ...
Article
The present study investigates the effect of annealing time and temperature on the microstructural restoration (by recovery and recrystallization), texture evolution and tensile properties of a dual-phase TWIP steel. The samples, which were initially hot rolled-air cooled followed by being in solution treated condition, subsequent 50 pct cold rolled condition and different annealing conditions (temperatures ranging from 500 °C to 1000 °C for 5 minutes to 2 hours), were subjected to microstructural characterization using optical and transmission electron microscopy, electron backscatter diffraction analysis, tensile testing and fractographic study. The deformation-induced ferrite (DIF) transformation due to cold deformation, the evolution of the grain structure and texture in both austenite and ferrite, and the change in the DIF fraction (accompanying the recrystallization annealing treatment) are critically analyzed. The optimum combination of strength and ductility is achieved in the partially recrystallized samples annealed at 700 °C for 30 minutes (UTS: 859 MPa, total elongation: 37 pct) and 900 °C for 30 minutes (UTS: 708 MPa, total elongation: 63 pct).The effects of DIF on the evolution of microstructure and texture during annealing and the final properties are discussed.
Article
As metal forming processes move toward high speed, high throughput, high precision and small scale with temperature dependence, clarifying the fundamental nano-deformation behavior of metals is critical for the optimization of manufacturing processes, and the control of nano-optical, electrical, mechanical or surface properties. Unfortunately, limited by the time scale and sample size, the effect of temperature on the deformation behavior of nano-metals during the ultrahigh-strain-rate forming process remains largely unexplored. This study demonstrates the nonlinear effect of temperature on the formability of nano-metals for the first time. Temperatures below 673 K facilitated the formability of nano-metals benefiting from the temperature-promoted dislocation proliferation process, whereas temperatures above 673 K weakened the plasticity of the nano-metal due to the activation of phase transformation. Frequent phase transition activation and accelerated dislocation annihilation at high temperatures reduced interstitial transport channels and delayed atomic transfer. Based on the temperature response of nano-metals in deformation mechanisms, defect evolution behavior and formability, the constitutive model and nano-deformation mechanism map of nano-metals in ultrahigh-strain-rate forming processes are proposed. The objective of this work is to provide basic support for the reasonable matching of nano-forming technology and processing temperature, and the determination of the optimal process window through fundamental nano-deformation behavior exploration.
Article
The effect of the amount of stored energy on the cementite spheroidization rate of 0.45 pct C–0.75 pct Mn (wt pct) steel was investigated. This was accomplished by using recently proposed thermomechanical conditions to control the microstructural characteristics of medium carbon cold-heading-quality steel. Wire milling was conducted right above the A3 temperature, after which the specimens were cooled at rapid cooling (TMRC) and slow cooling (TMSC) rates. The proeutectoid ferrite and pearlite fractions, cementite and grain size, grain boundary characteristics, and the energy stored in the specimens were quantitively measured to determine the effect of the manufacturing conditions. The spheroidization time was defined as a standard of the cementite aspect ratio of 8. The spheroidization times for the three specimens at 923 K (650 °C) were 3.8 hours (TMRC), 5.1 hours (TMSC), and 8.0 hours (conventional wire, CW), respectively. The amounts of recovered energy of the CW, TMRC, and TMSC specimens were 3.86, 5.75, and 5.00 J/g. The temperatures at which recovery started were 352 °C, 332 °C, and 337 °C for the CW, TMRC, and TMSC specimens. Thus, accelerated spheroidization times that were 1.5 to 2.1 times faster than those of the conventional CW sample could be obtained by tuning the initial microstructures with their large amounts of stored energy. The results of the microstructural analysis using FE-SEM and HR-EBSD showed that spherical cementites were located at the high- and low-angle grain boundaries. The different amounts of stored energy gave rise to differences in the cementite dissolution rates occurring at the beginning of spheroidization. A relatively large amount of stored energy (TMRC) seemed to promote an increase in the rate of self-diffusion of Fe, which in turn accelerated the diffusion rate of C (cementite growth and shape change), and finally resulted in the rapid spheroidization behavior.
Article
Full-text available
Ultrafine grained (UFG) steels with grain sizes around 1 micron exhibit an excellent strength-ductility combination and have been extensively studied worldwide. Among the different grain refinement strategies, thermomechanical controlled processing (TMCP) employing dynamic transformation (DT), that is, ferrite transformation during deformation of austenite, is considered as the simplest and commercially exploitable approach to produce ultrafine ferrite (UFF) with grain size of a couple of microns or below. The present paper reviews the research history of DT and highlights the major aspects of continuous interest including the methods and evidences for identifying DT, thermodynamics and kinetics of DT, mechanism for UFF formation and the effects of some key thermomechanical parameters on DT (and UFF formation), together with an outlook for the future research, and new TMCP design for industrial application. This paper also discusses some areas remaining under debate such as the diffusional or displacive mechanism, thermodynamic modeling, and the mechanism for UFF formation, etc.
Article
Full-text available
Fully recrystallized ultrafine-grained (UFG) pure Cu specimens were fabricated by high-pressure torsion (HPT) and controlled annealing. The recrystallized UFG Cu with a minimum mean grain size of 0.51 μm showed high yield strength, good ductility, obvious yield drop and large Lüders strain during tensile test. The mechanical behavior of the Cu specimen became sensitive to the change in the grain size from 0.2 μm to 4.2 μm. The continuous transitions of yield behavior and Lüders deformation with grain sizes were discussed.
Article
Full-text available
We have found a new strategy for ultra grain refinement without high strain deformation by combining dynamic transformation (DT) and dynamic recrystallization (DRX) mechanisms. Through simple thermomechanical processes using a total plastic strain of 0.92 at elevated temperatures, ultrafine grained microstructures having mean grain sizes down to 0.35 microns could be obtained in a 10Ni-0.1C steel. DRX phenomenon occurring in the dynamically transformed ferrite significantly reduced the strain necessary for the formation of ultrafine grains. The DRX of DT ferrite showed an unconventional temperature dependence, which suggested an optimal condition for grain refinement. The obtained UFG steel exhibited superior mechanical properties, for example, the tensile strength of 970 MPa and the total elongation over 20% at the same time.
Article
Full-text available
The grain refinement mechanism of ferrite dynamically transformed (DT) from austenite is investigated in a 10Ni–0.1C steel. For decades, it has been debated whether dynamic recrystallization (DRX) contributes to the grain refinement of DT ferrite. Here, the authors show that the role of DRX has been previously underestimated in grain refinement of ferrite and intrinsically it is the DRX that leads to ultra grain refinement of DT ferrite. The DT ferrite shows an unconventional grain growth behavior, that is, grain coarsening at an early stage of transformation with subsequent grain refinement by DRX, which strongly depends on the strain rate. The high strain concentration on fine ferrite grains within the austenite + ferrite structure produced by DT mainly contributes to the occurrence of DRX to form ultrafine-grained ferrite structures.
Article
Full-text available
Dynamic recrystallization (DRX) is an important grain refinement mechanism to fabricate steels with high strength and high ductility (toughness). The conventional DRX mechanism has reached the limitation of refining grains to several microns even though employing high-strain deformation. Here we show a DRX phenomenon occurring in the dynamically transformed (DT) ferrite, by which the required strain for the operation of DRX and the formation of ultrafine grains is significantly reduced. The DRX of DT ferrite shows an unconventional temperature dependence, which suggests an optimal condition for grain refinement. We further show that new strategies for ultra grain refinement can be evoked by combining DT and DRX mechanisms, based on which fully ultrafine microstructures having a mean grain size down to 0.35 microns can be obtained without high-strain deformation and exhibit superior mechanical properties. This study will open the door to achieving optimal grain refinement to nanoscale in a variety of steels requiring no high-strain deformation in practical industrial application.
Article
Full-text available
Two novel two-step thermomechanical routes were developed to produce ultrafine-grained ferrite microstructures in a 10Ni–0.1C steel without high-strain deformation. Homogeneous ultrafine ferrite (UFF) structures having mean grain sizes down to 460 nm were fabricated by a total equivalent strain of 0.92 and exhibited high yield strength of 820 MPa with large uniform elongation of 10% and total elongation of 29%. The formation of UFF was attributed to two different phenomena, i.e. dynamic transformation from austenite to ferrite and dynamic recrystallization of ferrite.
Article
Attention is directed to the uncertain state-of-affairs surrounding the influence of plastic deformation on concommitant tracer-diffusivity in metals. We now offer some clarification of this elusive solid-state transport problem. It is concluded that strain-enhanced diffusion is a real phenomenon, and in fact, it can be of substantial magnitude if the strain-rate is high enough and/or the temperature low enough. However, the dynamic-diffusivity measurements can be obscured by spurious mass movements in the diffusion zone arising from inadvertent specimen roughening or nonuniform bulk flow. The diffusivity enhancement per se occurs during compressive or tensile plastic deformation in poly- or mono-crystals, and in face-centered or body-centered cubic metals. Arguments are presented against the vacancy model as an explanation of the observed deformation/diffusion effects; instead, the dynamic diffusivity is accounted for by the presence of moving dislocation which provide high-diffusivity paths for the tracer atoms. The concentration profile remains Gaussian, indicating that the overall process is essentially one of random-walk under the prevailing conditions. Then, the temperature-dependence of the strain-enhanced diffusivity simply becomes the activation enthalpy for normal lattice diffusion minus that for dislocation-pipe diffusion. This interpretation can be checked by independent diffusion measurements under static conditions. The “window” on the experimental determinations of strain-enhanced diffusivities is limited by (a) the requirement for a steady-state structure during the deformation/diffusion run, in order to achieve a constant dynamic diffusivity, (b) the selection of strain-rate/temperature combinations which will permit the dynamic diffusivity to be significantly larger than the static diffusivity, and (c) the avoidance of recrystallization which serves to eliminate the operative dislocation substructure.
Article
Knowledge of solute interaction with the interface during the transformation of austenite into ferrite is fundamental in predicting its kinetics in multicomponent steel. This interaction notably translates in segregation, or depletion, of the solutes at the transformation interface. Here, this segregation was successfully quantified by atom probe tomography (APT) in four ternary Fe-X-C systems involving substitutional solutes commonly found in modern steel grades (X = Cr, Mn, Ni, Mo). Controlled decarburization was used to grow a uniform, planar and incoherent ferrite layer at the surface of fully austenitic samples. In the case of Fe-Cr-C and Fe-Mo-C, the interfacial concentrations permitted the evaluation of the binding energy of each substitutional solute to the interface, which was found to be comparable to its respective grain boundary binding energy. In the case of Fe-Mn-C and Fe-Ni-C, undesirable motion of the interface during the quench of the samples could not be avoided, preventing a reliable estimation of their binding energy since temperature and interface velocity were unknown.
Article
Formation of grain boundary ferrite (GB-α) in high carbon pearlitic steels deteriorates ductility of steel wires. Since it is generally believed that fully pearlitic structure forms in eutectoid steels, GB-α formation in eutectoid or even hypereutectoid steels is curious. Therefore, effects of temperature and carbon content on the formation of GB-α in Fe-1mass%Mn-(0.75 and 1.05) mass%C alloys transformed isothermally at temperatures ranging from 873 K (600 °C) to 973 K (700 °C) were investigated to clarify the formation mechanism of GB-α. It was found that volume fraction of GB-α increases with decreasing transformation temperature, carbon content and prior austenite grain size. Pearlite nucleated at prior austenite grain boundary usually grows into only one of austenite grains separated by the grain boundary, and GB-α forms on the other austenite grain. Orientation analyses revealed that GB-α and pearlitic-α hold the same orientation, indicating that one α grain grows as pearlite into one of austenite grains, and as GB-α into the other austenite grain. It was shown that such morphology difference is caused by the difference in orientation relationship (OR) between ferrite and austenite such that near K-S OR and non K-S OR correspond to GB-α and pearlite, respectively. Consequently, it was proposed that suppression of cementite nucleation at ferrite/austenite boundary holding near K-S OR is a reason for the formation of GB-α in the transformation at low temperature. Furthermore, degenerated pearlite (DP) and Widmanstatten ferrite (WF) is formed as well as GB-α at lower temperature, and OR dependency of GB-α, WF, DP and pearlite were clarified. © 2015 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.