Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

mechanism of plastic behavior (TWIP vs TRIP) has

Communication been by calculation of the intrinsic stacking fault energy


(ISFE).[5,7–10] It has been well established that deforma-
Processing and Properties of Medium- tion twinning will occur if the c-austenite ISFE is in the
Mn TRIP Steel to Obtain a Two-Stage range of 20 to 45 mJ/m2.[7–10] A TWIP-TRIP response,
where c-austenite will mechanically twin followed by a-
TRIP Behavior martensite formation at twin intersections, can be
obtained if the ISFE is 11 to 20 mJ/m2 and the two-
D.M. FIELD, L.G. GARZA-MARTINEZ, and stage TRIP behavior is observed when the ISFE of the c-
austenite <10 mJ/m2. Field et al.[5] recently showed that
D.C. VAN AKEN
in dual-phase microstructures it is important to calculate
Two medium-Mn steels of nominal composition 0.15C- the ISFE based on the composition of the c-austenite
2.0Si-10.5Mn-(0.75, 1.5) Al-0.04Nb-balFe (wt pct) were after annealing due to the enrichment of the c-austenite
processed to produce sub-micron grain sizes of 0.65 and with Mn and C leading to a change in the ISFE.
0.80 lm. Mechanical testing was performed in three In two-stage TRIP alloys, McGrath et al.[1] quantified
successive conditions: hot rolled, intercritically the microstructure of a steel with a composition of
annealed, and cold rolled with subsequent intercritical 0.07C-2.85Si-15.3Mn-2.4Al-0.017N-balFe in wt pct and
annealing. Intercritical annealing was performed at showed that the steel had two distinct work hardening
923 K for 20 hours. Electron-backscattered diffraction rates associated with each TRIP stage. Work hardening
and X-ray diffraction were utilized to characterize the during cfie TRIP was lowest with K  2500 MPa and n
microstructure, consisting of a-ferrite, a-martensite,  0.2, whereas a maximum in work hardening was
e-martensite, and c-austenite. Microstructural con- reached as efia TRIP occurred with K = 8,065 MPa
stituents were tracked during tensile deformation and and n = 1.3. Grässel observed that when TRIP is
it was found that both steels exhibited two-stage TRIP activated, the tensile strength and total elongation were
with c-austenite martensitically transforming first to increased due to the delayed necking from mobile
e-martensite and as strain increased e-martensite trans- dislocation generation.[2]
formed to a-martensite. Dynamic strain aging (DSA) is a significant concern
for implementation of these Mn steels due to the
negative strain-rate dependence on ultimate strength.
https://doi.org/10.1007/s11661-020-05901-2
Early onset of necking was observed in two-stage TRIP
 The Minerals, Metals & Materials Society and ASM
alloys due to DSA[5] as well as alloys where twinning is
International 2020
observed.[11,12] Field and Van Aken[4] showed that
introducing chromium of sufficient quantity to produce
M23(C,N)6 precipitates reduces the concentration of
Twinning-induced plasticity (TWIP) and transforma- dissolved interstitials in the c-austenite and this leads to
tion-induced plasticity (TRIP) steels with 10 to 15 wt pct the dynamic strain aging mechanism transitioning from
Mn have demonstrated properties desirable for auto- an interstitial defect interaction to the trapping of Mn at
motive applications with ultimate tensile strengths of the stacking fault. This work attempts to elucidate the
1200 to 1500 MPa and total ductility of 30 to 25 pct. effect of aluminum on the microstructure of two-stage
These manganese bearing steels obtain their elevated TRIP steels, and how the developed microstructure
strengths and ductility either through the formation of influences the work hardening capability of the alloys.
deformation twins (TWIP steels), single-stage TRIP Two steels of nominal composition 0.15C-2.0Si-
(cfia), or two consecutive martensitic transformations 10.5Mn-(0.75, 1.5)Al-0.04Nb-balFe (compositions given
(c fi e fi a) leading to a two-stage TRIP behavior.[1–6] in wt pct) were induction melted using iron, ferrosilicon,
To date, the most successful method of predicting the electrolytic manganese, pure aluminum, ferroniobium,
and carbon in the form of graphite. An argon cover gas
was used to shield the melt and calcium wire was added
to modify oxide inclusions and remove sulfur. Before
tapping a synthetic vermiculite, slag was added to the
surface of the melt and an argon lance with a porous
D.M. FIELD is with the US Army Combat Capabilities Development plug was submerged into the melt to inject gas bubbles
Command Army Research Laboratory, Aberdeen Proving Ground, MD for inclusion flotation and slag entrapment; this was
21005. Contact e-mail: Daniel.m.field6.civ@mail.mil L.G. GARZA- performed for 60 seconds prior to tapping into the ladle.
MARTINEZ is with AK Steel, Middletown, OH, 45005. D.C. VAN Alloys were tapped into a ladle with a 150 K (150 C)
AKEN is with the Department of Materials Science and Engineering,
Missouri University of Science and Technology, Rolla, MO 65409.
superheat and cast into a phenolic no-bake sand mold to
Manuscript submitted January 28, 2020. form a Y-block with dimensions measuring 176 9 178 9
Article published online July 6, 2020 60 mm3. Chemical analyses were obtained by inductively

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 51A, SEPTEMBER 2020—4427


coupled plasma (ICP) spectrometry after sample disso- annealing with a cool down of typically ~24 hours, and
lution in hydrochloric and nitric acid. Carbon and this material will be designated as the ICA condition.
nitrogen contents were determined using Leco CS600 The main purpose of the ICA treatment was to permit
and TC500 analyzers, respectively. Alloy composition, greater reductions during cold rolling. A Stanat TA 315
ISFE, and martensite start temperatures are listed in in a 4-high roll configuration was used to cold roll the
Table I. Alloys are designated by their nominal alu- ICA material to a reduction of 50 pct in thickness. Cold
minum concentration. Mse and ISFE were calculated rolled alloys were heated to 923 K (650 C) for 20 hours
according to the method of Olson and Cohen[13] using in chemically treated stainless steel bags and furnace
Eq. [1] with an updated regular solution model as cooled to 298 K to simulate an industrial batch anneal-
described by Pisarik and Van Aken.[14] ing process and are designated as BA. A schematic of
  the processing is shown in Figure 1.
ISFE mJ=m2 ¼ nqðDGc!e Þ þ 2rc=e ½1 Tensile bars were milled from the HB, ICA, and BA
strip according to ASTM E8[16] standard for testing
where n is the number of (0001)e layers within the par-
sheet product. All tensile bars had a gage length of 50
ent austenite and q is the molar surface density of the
mm and gage width of 12.5 mm with tensile loading
{111}c for the steel composition using Vegard’s law.
conducted parallel to the rolling direction. The sheet
DGcfie is the driving force for the transformation of c-
surface was not machined nor chemically milled prior to
austenite to e-martensite, and rc/e is the interfacial
testing. Interrupted tensile tests were conducted on the
energy between the c-austenite and e-martensite and
final processed BA condition using displacement control
was held constant at 10 mJ/m2. The martensite start
rate of 0.01 mm/sec with a 245 kN servo-hydraulic test
temperature, MaS , was calculated according to the work
frame with strain measured using a contact extensome-
by Field and Van Aken,[15] where the strain energy of ter. For samples strained to failure, a non-contact laser
transformation, ( DGc!astr ) was balanced against the extensometer was utilized to capture the entire strain to
chemical driving force ( DGc!a
Chem ) according to Eqs. [2] failure.
and [3]. X-ray diffraction (XRD) samples were obtained for
DGc!a c!a all conditions to investigate the microstructural evolu-
Chem þ DGStr ¼ 0 ½2
tion throughout processing. XRD samples were
  mechanically polished to 0.1 lm using a diamond paste
DGc!a
Str
J=mol ¼ EXd2 ð14:8  0:013TÞ ½3 in the Longitudinal-Transverse plane (Rolling Plane
Normal). Diffraction patterns were obtained with a
where DGc!aChem is calculated according to a modified Philips X-pert diffractometer using a Cu Ka radiation
regular solution model, X is the molar volume for iron
of 7.15 9 106 (m3/mol), d is the lattice misfit between
the c-austenite and a-martensite with an approximate
strain of 1.11 9 102 (m/m), T is the temperature in
Kelvin, and E is the modulus in units of Pa calculated
in accordance with Reference 12. The start tempera-
ture for the a-martensite was calculated by determining
the temperature where Eq. [2] is true. It should be
noted that the values calculated in Table I assume a
fully austenitic microstructure with the listed
chemistry.
Castings were milled to a parallelepiped block of
dimensions 173 9 150 9 58 mm3, hot rolled at 1523 K
(1250 C) to a thickness of 2.0 mm, and cooled between
two insulating blankets. Finish roll temperature was 990
± 30 K (717 ± 30 C). Alloys in this condition are
termed hot band (HB) steels. Intercritical annealing Fig. 1—Schematic of the processing history of the alloys under
(ICA) was performed at 923 K (650 C) for 20 hours and investigation.
furnace cooled to room temperature to simulate a batch

Table I. Composition and Calculated Parameters of the Alloys Under Investigation

Composition (Wt Pct) Calculated Parameters*

Mse Msa
Alloy C Mn Si Al Nb N ISFE (mJ/m2) K (C) K (C)
0.75Al 0.15 10.2 2.07 0.71 0.040 0.026 3.0 346 (73) 453 (180)
1.5Al 0.17 10.6 2.08 1.54 0.042 0.013 11.5 286 (13) 360 (87)
*Assuming single phase c-austenite.

4428—VOLUME 51A, SEPTEMBER 2020 METALLURGICAL AND MATERIALS TRANSACTIONS A


source with a Ni filter and equipped with a flat graphite Stress vs strain plots for the two alloys in all three
monochrometer. Phase quantification, including e- conditions are shown in Figure 2. Hot band and
martensite, was calculated using the Rietveld refinement intercritically annealed alloys exhibit rapid work hard-
method as described by Martin et al.[17] for an Fe-16Cr- ening immediately after yielding and reached ultimate
6.8Mn-6.1Ni steel and modified accordingly for the steel tensile strengths greater than 1400 MPa with total
chemistry investigated in this study. strains to failure ranging from 12 to 25 pct. Cold rolled
Specimens for electron-backscattered diffraction and batch-annealed alloys demonstrate strains to failure
(EBSD) were also mechanically polished and finished greater than 38 pct. All specimens exhibited post
using a vibratory polisher with a 0.02 lm colloidal silica necking ductility despite observing dynamic strain aging
solution. EBSD samples were examined in the Longitu- in the BA condition for both alloys as well as the 0.75Al-
dinal-Short plane (perpendicular to both the rolling ICA condition. A summary of tensile properties is
plane normal and rolling direction). Phase image map- provided in Table II.
ping via pattern analysis was performed on a Helios EBSD phase maps of the two alloys in the various
Nanolab 600 using an accelerating voltage of 20.0 kV processed conditions are shown in Figure 3 which
and emission current of 11 nA. Phase identification was illustrates the distribution of three phases: c-austenite
performed using a Nordlys detector and the AZTEC is shown in green, e-martensite is shown in red, and a-
software package. Degree of recrystallization was deter- ferrite/martensite is shown as blue. The mean linear
mined according to the work described by Field et al.[5] intercept (L3) of the BA alloys was measured to be 0.65
and using the post-processing software Channel 5 on ± 0.05 lm for the 0.75Al alloy and 0.80 ± 0.02 lm for
multiple EBSD maps to analyze at least 1000 grains. the 1.5Al alloy. Instantaneous work hardening rate and
Grains containing internal misorientation greater than work hardening exponents were calculated according to
5 deg were considered deformed, and grains with Eqs. [4] and [5] and graphed for the BA alloys with the
misorientation less than 1 deg were considered fully phase quantities as measured according to XRD in
recovered. Grain size was measured according to the Figure 4. Logarithmic smoothing was applied to the
ASTM E112–13[18] using the Heyn–Lineal Intercept curve. Tabulated XRD volume fractions of the three
method. conditions are shown in Table III. The highest rate of

Fig. 2—Stress-strain plot of the (a) 0.75 Al and (b) 1.5Al alloys under investigation throughout processing.

Table II. Mechanical Properties of the Alloys in the Three Conditions and the Measured Degree of Deformation According to
EBSD Analysis with a 95 Pct Confidence Level Shown

YS Recrystallized
Alloy Condition (MPa) UTS (MPa) Total Strain (Pct) Deformed (Vol Pct) (Vol Pct)
0.75Al HB 490 1690 13.0 74 ± 6 26 ± 4
ICA 385 1490 24.6 54 ± 10 46 ± 11
BA 770 1280 38.0 31 ± 3 69 ± 8
1.5Al HB 670 1600 12.1 69 ± 11 31 ± 5
ICA 610 1390 19.5 68 ± 6 32 ± 7
BA 875 1200 47.5 28 ± 1 72 ± 1

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 51A, SEPTEMBER 2020—4429


Fig. 3—EBSD phase maps of the 0.75Al alloy in the (a) HB, (b) ICA, and (c) BA conditions and the 1.5Al steel in the (d) HB, (e) ICA, and (f)
BA conditions. c-austenite is shown in green, e-martensite is red, and a-ferrite/martensite is shown in blue (Color figure online).

Fig. 4—Instantaneous work hardening rate (K), exponent (n), and microstructural components as a function of true plastic tensile strain of the
0.75Al-BA (a), (b). 1.5Al-BA (c) and (d).

work hardening was obtained during the second stage of


ei
TRIP with e-martensite transforming to a- martensite. n¼K ½5
ri
dr
K¼ ½4
de Previous researchers work on two-stage TRIP steels
have shown that there are two distinct martensitic
reactions that occur during straining. McGrath et al.[1]

4430—VOLUME 51A, SEPTEMBER 2020 METALLURGICAL AND MATERIALS TRANSACTIONS A


and Field and Van Aken[4] showed that for the initial documented[5,11,19] that aluminum will lower the work
transformation, Stage I, a low initial work hardening hardening behavior of both TRIP and TWIP steels. The
rate (~2700 MPa) is observed for the two-stage TRIP primary difference between the two alloys being inves-
steel; and is observed in the current study where both tigated here is the aluminum content and the lower work
alloys exhibit a work hardening rate (K) of 1200 to 1600 hardening rate of the 1.5Al alloy appears to be related to
MPa in the initial Stage I regime. The higher initial work the slower rate of a-martensite formation, since both
hardening rate reported by McGrath et al. is potentially steels produced an equivalent amount of transformable
due to a very low initial yield strength associated with a e-martensite prior to the onset of Stage II TRIP.
prior austenite grain size of 30 lm and the starting In the hot bands microstructure, a fraction of grains
microstructure containing a higher fraction of c-austen- appeared to be not fully recrystallized or unrecovered.
ite and e-martensite. The high volume fraction of e- The increase in aluminum concentration leads to an
martensite contributes to TRIP induced refinement of increased volume of recovered grains in the hot band
the microstructure during work hardening. McGrath samples. Prior observations by Field et al.[5] on two-
et al. reports properties in a hot band condition that was stage TRIP steels has demonstrated that the stored
rolled to gauge thickness, reheated to 1173 K (900 C) deformation from hot rolling exhibits a non-linear
for 10 minutes, and quenched to room temperature in relationship to the Si:Al ratio, where a Si:Al > 1 leads
water. The developed microstructure was reported to be to a primarily unrecovered steel in the hot rolled product
a mixture of c-austenite (27 pct), e-martensite (60 pct), (HB).
and a-martensite (13 pct). The BA alloys in this work The 0.75Al alloy also has a higher concentration of e-
contain c-austenite (< 40 vol pct), e-martensite, and a- martensite in both conditions (10 vol pct HB and 23
martensite/ferrite (> 45 vol pct) due to processing by vol pct ICA) compared to the 1.5Al alloy which has a
cold rolling and intercritical annealing. The a-phase will higher volume fraction of c-austenite (18 vol pct HB and
not TRIP and reduces the work hardening capability of 22 vol pct ICA). The increased e-martensite content
these alloys. The total content of transformable product, leads to a consistently lower yield strength during tensile
the volume fraction of c-austenite, and e-martensite is testing (ry 0.75Al < ry 1.5Al), and this is rationalized
much lower (44 ± 11 vs87 vol pct) in the alloys due to the volume change associated with the competing
contained within this study compared to the work by martensitic reactions. The cfie-martensite transforma-
McGrath et al. The maximum work hardening rate and tion is a volume contraction of  0.74 pct as reported
exponent (K  6000 MPa, n  0.90 in Figures 4(a) and for both two-stage TRIP alloys and shape memory
(b)) for the 0.75Al-BA alloy was observed at true plastic steels.[5,20,21] Upon straining, the e-martensitefia trans-
strain value of 0.20 and this corresponds to the final e- formation is a volume expansion of +2.1 pct. as
martensite transforming to a-martensite. A lower work reported in Reference 5 and this volume change is
hardening rate and exponent (K  3700 MPa and n  larger than the direct cfia reaction (+1.4 pct.). A
0.70) is observed for the 1.5Al-BA steel at a true plastic volume expansion (e fi a, DV>0) will accommodate an
strain value of 0.30. A maximum in the strain hardening imposed tensile strain, whereas a volume contraction
occurs when the final e-martensite is exhausted and e- (cfie, DV<0) will resist the imposed tensile strain. Thus,
martensite transform to a-martensite: at true plastic alloys with higher concentrations of e-martensite tend to
strains of 0.07 for 0.75Al and 0.12 for 1.5Al. It is well have lower yield strengths whereas alloys with high c-

Table III. Phase Quantities of the Hot Band, Intercritical Annealed, and Batch-Annealed Processed Alloys Utilizing XRD

c-Austenite e-Martensite a-Martensite/Ferrite


Alloy Condition (Vol Pct)
0.75Al HB 2 4 94
ICA 9 24 67
BA 22 11 67
1.5Al HB 8 1 91
ICA 21 6 73
BA 40 16 44

Table IV. Composition and Calculated Thermodynamic Parameters of the c-Austenite from Multiphase Equilibria Simulation
Using ThermoCalc 2018a

Composition (Wt Pct) Calculated Parameters

SFE Mse
Alloy C Mn Si Al N NbC (mJ/m2) K (C)
0.75Al-BA 0.20 19.8 2.35 0.21 0.026 0.046 4.6 341 (68)
1.5Al-BA 0.23 16.2 2.40 0.90 0.013 0.048 7.6 317 (44)

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 51A, SEPTEMBER 2020—4431


Table V. Calculated Mechanical Properties of the Developed Alloys According to Equations Referenced from Ref. 5

Calculated Values Rel. Error (Pct)

Alloy YS (MPa) UTS (MPa) Pct etot YS UTS Pct etot


0.75Al-BA 665 1315 40  13 3 5
1.5Al-BA 620 1270 50  29 4 5

austenite that first TRIPs to e-martensite will have the yield strength of the alloys compared to the
higher yield strengths assuming each has the same relationship reported by Field et al.; however, it should
content of a-phases. be noted that the alloys under investigation are modified
To better replicate the composition of the a-ferrite with Nb to promote NbC formation during hot rolling.
and c-austenite using the thermodynamic software, a A Nb contribution was not included in the previously
temperature was determined to reproduce the measured proposed empirical equation and more work needs to be
a-ferrite volume fraction. ThermoCalc 2018b was performed to determine where these carbides reside: at
utilized to calculate the thermodynamic equilibrium grain boundaries to provide grain size stability or grain
composition of the c-austenite using the TCFE9 interiors leading to Orowan strengthening. It has been
database and this chemistry is shown in Table IV with reported that NbC as a micro-alloying addition can lead
an equilibrium temperature chosen to replicate the to significant increases in strength of Mn TWIP
quantity of measured a-ferrite. Precipitation of alloys.[22, 23] Bai et al.[22] showed that with a moderate
Nb(C,N), partitioning of Al to the a-ferrite, and Mn addition of niobium (0.09 wt pct) a strength increase of
and C to the c-austenite significantly alters the ISFE up to 160 MPa could be obtained. Scott et al. however
and e-martensite stability of the alloy. The loss of Al showed that Nb had a moderate effect (~60 MPa) on
and increase in Mn and C into the c-austenite decrease increasing the strength of high Mn (‡ 18 wt pct) TWIP
the ISFE and raises the e-martensite start temperature steels[23] after accounting for grain size. If an average
of both alloys. This effect is most significant in the increase of 100 MPa is added to the strengths of the steel
1.5Al steel with a decrease of 4 mJ/m2 compared to the in this investigation a relative error of  6 and  18 pct
bulk composition leading to an increase in the Mse of is obtained. The UTS and total elongation show a good
31 K (31 C) and translates to an Mse > 298 K (25 C). agreement with the calculated properties exhibiting a
This explains the higher volume fraction of e-martensite relative error less than 5 pct.
observed in BA vs both the ICA and HB conditions. Two 10.5 wt pct manganese steels have been pro-
This translates to greater total elongation and a cessed to obtain a two-stage TRIP response during
reduced work hardening rate. tensile testing. A high ultimate tensile strength >1220
Field et al.[5] report a compositionally based empir- MPa and an increased total elongation >38 pct was
ical equation to calculate the ultimate tensile strength obtained after hot rolling, intercritical annealing at
and total elongation of two-stage TRIP steels and are 923 K (650 C), and cold rolling 50 pct and annealing at
shown in Eqs. [6] and [7] where xi represents the weight 923 K (650 C). A sub-micron grain structure comprised
percent alloying element ‘‘i’’; they also showed that for of c-austenite, e-martensite, and a-ferrite was developed
the sub-micron grain structure an inverse root rela- after processing. Strength and ductility were shown to
tionship to the yield strength was obtained for the exceed the DOE targets for future 3rd generation
mean linear intercept distance (L3) calculated according advanced high-strength steels after processing to obtain
to Eq. [8]. a dual-phase microstructure.
UTS ðMPaÞ ¼ 2580ðxC Þ þ 13ðxMn Þ  42ðxSi Þ
 30ðxAl Þ þ 72ðxCr Þ þ 7820ðxN Þ þ 750
½6
This work was supported by the Peaslee Steel
Tot:eðpct:Þ ¼ 77ðxC Þ  2ðxMn Þ  5ðxSi Þ Manufacturing Research Center (PSMRC). Companies
þ 20ðxAl Þ þ 6ðxCr Þ  115ðxN Þ þ 51 directly involved in this work include AK Steel,
ArcelorMittal, Nucor Steel, and US Steel. The FEI
½7 Helios NanoLab dual-beam FIB was obtained with a
Major Research Instrumentation grant from the Na-
 pffiffiffiffiffiffiffi tional Science Foundation under Contract DMR-
390 MPa lm
ry ðMPaÞ ¼ pffiffiffiffiffiffi þ 185ðMPaÞ ½8 0723128. The authors also acknowledge the support of
L3 the Materials Research Center and in particular Dr.
Clarissa Wisner and Dr. Eric Bohannan for guidance
The calculated values and their relative error are and training in using the FIB and performing the
shown in Table V. A high relative error is calculated for XRD work.

4432—VOLUME 51A, SEPTEMBER 2020 METALLURGICAL AND MATERIALS TRANSACTIONS A


REFERENCES 14. S.T. Pisarik and D.C. Van Aken: Met. Trans. A., 2014, vol. 45,
pp. 3173–78.
1. M.C. McGrath, D.C. Van Aken, N.I. Medvedeva, and J.E. 15. D.M. Field, D.S. Baker, and D.C. Van Aken: Met. Trans. A, 2017,
Medvedeva: Metall. Mater. Trans. A, 2013, vol. 44A, pp. 4634–43. vol. 48, pp. 2150–63.
2. O. Grässel, L. Krüger, G. Frommeyer, and L.W. Meyer: Int. J. 16. ASTM E 8/E 8M-08, Standard Test Methods for Tension Testing
Plast., 2000, vol. 16, pp. 1391–1409. of Metallic Materials.
3. D.M. Field and D.C. Van Aken: Met Trans A., 2016, vol. 47A, 17. S. Martin, C. Ullrich, D. Simek, U. Martin, and D. Rafaja: J.
pp. 1912–17. Appl. Crystallogr., 2011, vol. 44, pp. 779–87.
4. D.M. Field and D.C. Van Aken: Met Trans A., 2018, vol. 49, 18. ASTM E112-13, Standard Test Methods for Determining Average
pp. 1152–66. Grain Size.
5. D.M. Field, J. Qing, and D.C. Van Aken: Met. Trans. A., 2018, 19. S.-J. Lee, J. Kim, S.N. Kane, and B.C. De Cooman: Acta Mater.,
vol. 49, pp. 4615–32. 2011, vol. 59, pp. 6809–19.
6. B.C. de Cooman, P. Gibbs, S. Lee, and D.K. Matlock: Met. Trans. 20. S.K. Huang, Y.H. Wen, N. Li, J. Teng, S. Ding, and Y.G. Xu:
A, 2013, vol. 44A, pp. 2563–72. Mater Charact., 2008, vol. 59, pp. 681–87.
7. B.C. De Cooman, Y. Estrin, and S.K. Kim: Acta Mater., 2018, 21. S. Shin, M. Kwon, W. Cho, I.S. Suh, and B.C. De Cooman:
vol. 142, pp. 283–362. Mater. Sci. Eng., 2017, vol. 683, pp. 187–94.
8. L. Remy: A. Pineau Mater. Sci. Eng., 1976, vol. 26, pp. 123–32. 22. D.Q. Bai, F. Hamad, J. Asante, and S. Hansen: Mater. Sci. Forum,
9. S. Allain, J.P. Chateau, and O. Bouaziz: Mater. Sci. Eng., 2004, 2005, vols. 500–501, pp. 481–88.
vol. 387, pp. 143–47. 23. C.P. Scott, B. Remy, L.-J. Collet, A. Cael, C. Bao, F. Danoix, B.
10. T.H. Lee, E. Shin, C.S. Oh, H.Y. Ha, and S.J. Kim: Acta Mater., Malard, and C. Curfs, Int. J. Mater. Res. 2011.
2010, vol. 58, pp. 3173–86.
11. S.-J. Lee, J. Kim, S.N. Kane, and B.C. DeCooman: Acta Mater.,
2011, vol. 59, pp. 6809–19. Publisher’s Note Springer Nature remains neutral with regard to
12. Y.N. Dastur and W.C. Leslie: Mater. Trans. A, 1981, vol. 12, jurisdictional claims in published maps and institutional affiliations.
pp. 749–59.
13. G.B. Olson: M. Cohen Met. Trans. A, 1976, vol. 7, pp. 1897–1904.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 51A, SEPTEMBER 2020—4433

You might also like