Kim 2018
Kim 2018
Kim 2018
PII: S0921-5093(18)30751-2
DOI: https://doi.org/10.1016/j.msea.2018.05.083
Reference: MSA36521
To appear in: Materials Science & Engineering A
Received date: 13 February 2018
Revised date: 21 May 2018
Accepted date: 23 May 2018
Cite this article as: Ji Hoon Kim, Eun Jung Seo, Min-Hyeok Kwon, Singon Kang
and Bruno C. De Cooman, Effect of Quenching Temperature on Stretch
Flangeability of a Medium Mn Steel Processed by Quenching and Partitioning,
Materials Science & Engineering A, https://doi.org/10.1016/j.msea.2018.05.083
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Effect of Quenching Temperature on Stretch Flangeability of a
Medium Mn Steel Processed by Quenching and Partitioning
Ji Hoon Kima, Eun Jung Seoa,1, Min-Hyeok Kwona, Singon Kanga,2,*, Bruno C. De Coomana
a
Graduate Institute of Ferrous Technology, Pohang University of Science and
Technology, 77 Cheongam-Ro, Nam-Gu, Pohang, 37673, Republic of Korea
1
Current address: Advanced Steel Processing and Products Research Center,
Colorado School of Mines, 1500 Illinois St. Golden, CO, 80401, USA
2
Current address: Department of Materials Science and Engineering, Dong-A
University, 37, Nakdong-daero 550, Saha-gu, Busan, 49315, Republic of Korea
Abstract
The stretch flangeability of a medium Mn steel prepared by quenching and partitioning was
90 C ≤ TQ ≤ 170 C, at which no secondary martensite is formed upon final quenching. The
hole-expansion test used a 60° conical punch. The hole at the center of the specimens was
prepared by punching or wire cutting. In both sample preparation conditions, hole expansion
ratio (HER) increased as TQ decreased. The dependence of HER on TQ was more pronounced
in the samples prepared by punching than in samples prepared by wire cutting. While there is
no clear correlation between HER and tensile properties, the HER decreased as retained
martensite in the shear-affected zone near the hole edge. This martensite has a negative
1
Keywords: Hole expansion ratio, Quenching and partitioning, Medium Mn steel, Stretch
1. Introduction
Quenching and partitioning (Q&P) steel [1] is a promising candidate for third-
(RT), Q&P steel is composed of a martensitic matrix with retained austenite as a secondary
constituent. The stability of the retained austenite is controlled by partitioning of carbon into
austenite after the formation of martensite [2]. Q&P steels exhibit attractive tensile properties
due to the combination of the high strength of martensitic matrix and deformation-induced
The Q&P heat treatment process was initially applied to conventional transformation-
induced plasticity (TRIP) steels, such as 0.2C-1.6Mn-1.6Si (in wt. %) steel, which contains a
maximum of 13 vol. % of retained austenite after a Q&P process [3, 4]. The Q&P heat
treatment process has since been applied to steels of a broad range of compositions, and has
(in wt. %) medium Mn steel retained up to 33 vol. % austenite after the Q&P process [10]. In
this steel the product of ultimate tensile strength (UTS) and total elongation (TE) was ~25
GPa·%.
The hole expansion ratio (HER) is a representative parameter to quantify the stretch
flangeability of a steel sheet. Although the uniaxial tensile test has a different mode of
2
deformation from the hole expansion test, HER values have been related to tensile properties
such as the UTS [11-13], post uniform elongation (PUE) [14, 15], strain hardening rate [16],
and yield ratio (yield strength (YS)/UTS) [17]. In addition, the size of the shear-affected-zone
(SAZ) generated by hole punching [18-21], the hardness difference between hard phase and
soft phase in multi-phase steel [22, 23], and the volume fraction and stability of the retained
austenite in TRIP steel [24] have been invoked to explain stretch flangeability.
Some of these approaches were used to investigate the stretch flangeability of Q&P
processed steel [25, 26]. De Moor et al. [25] reported that fully-austenitized Q&P steel has
steel after full austenitization and one-step Q&P are positively correlated with yield ratio [26].
In the present study, the stretch flangeability of Q&P processed medium Mn steel
was investigated as a function of quenching temperature TQ, which is the main Q&P
processing parameter. The hole expansion test was performed using samples prepared by
punching; their HER values were compared with the results from specimens with wire-cut
holes to investigate the effect of SAZ formation on the HER. The evolution of HER with TQ
was analyzed by correlating HER with tensile properties, retained austenite fraction, and the
2. Experimental Procedure
thickness of 1.2 mm. The two-step Q&P process was used. The samples were fully
3
170 °C in an oil bath. After the initial quenching, the samples were heated to 450 °C, held for
emission scanning electron microscopy (FE-SEM, FEI Quanta 3D FEG) and electron back-
CH3COOH. The EBSD results with a minimum confidence index (CI) value of 0.1 were used.
The retained austenite fraction was measured using the magnetic saturation method by
specimens was estimated as C = C0 / V, where C0 is the C concentration of the steel, i.e. 0.22
wt. % and V is the measured retained austenite fraction. Prediction of C applied two
assumptions: (1) the phases present during the partitioning are retained austenite and
martensite formed during the initial quenching, and (2) C partitioning from martensite to
emission transmission electron microscopy (FE-TEM, JEOL JEM-2100F). Specimens for the
FE-TEM observations were prepared as thin foil discs with a diameter of 3 mm, and electro-
chemically polished using a twin-jet polisher with a solution of 5% HClO4 + 95% CH3COOH.
The microstructural evolution of the samples during the Q&P process was examined
using dilatometry. A Bähr 805 pushrod dilatometer was used with specimen dimensions of 10
5 1.2 mm. The heating rate was similar to box-furnace heating (20 C/s up to 300 C,
10 C/s up to 600 C, and 2 C/s up to 850 C), and the cooling rate was about 50 C/s. The
Q&P process was simulated in vacuum atmosphere of 3 10-4 mbar; He was injected for
quenching.
4
The tensile tests were carried out using a Zwick/Roell universal tensile tester at a
strain rate of 1 10-3 s-1. Tensile test samples were prepared by wire cutting according to
ASTM E8M for sub-size samples with the gauge length of 25 mm. The tensile properties
were obtained by averaging triplicate measurements. Sheets for the hole expansion test were
prepared with dimensions of 90 mm 90 mm before the Q&P heat treatment process. After
the Q&P heat treatment, 10-mm diameter holes were prepared in the center of the specimens
by either mechanical punching with a clearance of 20%, or by wire cutting. The HER tests of
the two types of samples were conducted using an Erichsen Universal Sheet Metal Testing
Machine (145-60), according to the ISO 16630:2009 standard [27]. A 60° conical punch was
used with a punch velocity of 8 mm/min. The specimens were clamped by a blank holder
force of 200 kN. The tests were stopped when a crack propagated through the full thickness
( ) (1)
where Df is hole diameter after the fracture, and D0 = 10 mm is the initial hole diameter. The
reported HER values were obtained by averaging the diameters in the rolling and transverse
directions.
The SAZ generated by punching was analyzed using a hardness profile in samples
that had been quenched at 90 °C or 170 °C. Hardness was measured using a micro Vickers
hardness tester (Mitutoyo HM-220), with an applied load of 0.2 kgf. The punched hole edge
has roll-over, burnished, and fractured areas (Fig. 1). The hardness was measured every 100
m from the interface between the burnished and fractured parts. Values were obtained by
5
averaging six measurements. Microstructures that included the retained austenite distribution
3. Results
130 °C, or 170 °C show prior austenite grain structure with average size ~6 m. The
morphology of the interior of the prior austenite grains illustrates the complex substructure of
the Q&P processed steels. The corresponding EBSD phase maps (Figs. 2d – f) were used to
characterize their substructure. BCC martensite formed lath-like features; the remainder was
SEM micrographs (Figs. 2g – i) reveal white particles in the martensitic areas; these particles
are carbides formed during partitioning. The carbide density decreased as TQ increased.
6
Fig. 2. (a) – (c) FE-SEM micrographs [yellow dotted lines: prior austenite], (d) – (e)
EBSD phase maps [Green: BCC; red: FCC] and (g) - (i) high-magnification FE-SEM
micrographs [yellow arrows: carbide precipitates in tempered martensite] of 0.22C-
3.79Mn-1.48Si-0.98Cr steel quenched at 90, 130, or 170 °C.
micrographs to reveal the morphologies of retained austenite (Fig. 3). In both specimens, the
original austenite grains were segmented into small areas by the formation of lath martensite.
In the sample quenched at TQ = 90 ºC, film-type retained austenite was dominant. In the
sample quenched at TQ = 170 °C, the retained austenite had a blocky-type morphology.
7
Fig. 3. Bright-field TEM micrographs of film-type retained austenite in specimens
with (a) TQ = 90 °C and blocky-type retained austenite in specimens with (b) TQ = 170
°C, and corresponding dark field TEM micrographs of the austenite with (c) TQ = 90
°C and (d) TQ = 170 °C.
0.25 at TQ = 170 °C (Fig. 4a). When austenite is not fully stabilized, secondary martensite
forms from the C-partitioned austenite during final quenching. This martensite is C-enriched
and generally undesirable due to its brittle characteristics, but it frequently occurs when
specimens are Q&P processed and quenched at high TQ; the result is typically a consequence
The dilatation curves measured for the Q&P process used in the present study confirm
that no secondary martensite formed during cooling to RT. The sample quenched at TQ =
170 °C was fully austenitized at 850 °C (Fig. 4b). During the initial quenching to TQ, no
ferrite transformation was detected, but martensite formed. After the initial martensite
8
formation, the dilatation traces before and after the partitioning are linear and overlap. This
relationship indicates negligible formation of bainite and secondary martensite during and
after partitioning.
Fig. 4. (a) Effect of quenching temperature on the retained austenite fraction of a Q&P
processed 0.22C-3.79Mn-1.48Si-0.98C steel measured by magnetic saturation method
and (b) a dilatometric plot for a Q&P processed sample with a quenching temperature
of 170 °C followed by partitioning at 450 °C for 5 min.
The stability of the retained austenite was evaluated in Q&P processed specimens by
analyzing how TQ affected C, grain size and grain morphology. Calculated C decreased
from 1.69 at TQ = 90 °C to 0.92 at TQ = 170 °C (Fig. 5a). Carbide precipitation limits the full
partitioning of C that was assumed for the calculation. However, in a Q&P steel with a
calculated values.
at TQ = 170 °C (Fig. 5b). The retained austenite morphology changed from film-type to
blocky-type as TQ increased (Fig. 3). In the present study, the retained austenite in the
samples that had been quenched at TQ = 90 °C had high C, reduced grain size, and film-type
morphology, which are all features that indicate increased stabilization of the austenite [28].
9
Fig. 5. (a) Calculated carbon concentrations of retained austenite with different
quenching temperature by assuming full partition of carbon from martensite to
austenite during partitioning heat treatment, and (b) average grain size of retained
austenite with different quenching temperature obtained from EBSD measurements.
The tensile properties of the Q&P processed specimens were also affected by TQ (Fig.
6). The sample quenched at TQ = 90 ºC had YS = 1219 MPa and UTS = 1385 MPa. Both YS
and UTS decreased slightly as TQ increased; this trend was stronger in YS than in UTS (Fig.
maintained as TQ increased (Fig. 6b). The evolutions of the tensile properties are generally
explained by the increase in the soft retained austenite fraction in the hard martensite matrix
as TQ increased, and the contribution of this austenite to strain hardening is related to the
10
Fig. 6. Effect of quenching temperature on (a) yield strength (YS) and ultimate tensile
strength (UTS) and (b) uniform elongation (UE), total elongation (TE), and post-
uniform elongation (PUE) in the Q&P-processed medium Mn steel.
In hole expansion tests, HER was lower in the punched specimens than in the wire-cut
specimens (Fig. 7); this result concurs with earlier studies [14, 15, 23, 29-31]. The formation
of a deformed layer, i.e. the SAZ area, by punching is considered to degrade the HER of
punched specimens. Hardness measurements made in the punched hole edge area were used
to examine the formation of SAZ for the samples treated at TQ = 90 C or 170 C. The matrix
quenched at TQ = 170 C (419 Hv) (Fig. 8). The hardness increased towards the edge; the
maximum hardness at the edge of the punched hole was 530 Hv at TQ = 90 C and 521 Hv at
TQ =170 C. Although hardness in both the matrix and edge were highest in the sample
quenched at TQ = 90 C, the hole-punching process increased the hardness most in the sample
11
Fig. 7. Hole expansion ratio (HER) of the Q&P-processed medium Mn steel with
different quenching temperatures. Black: punched specimens; red: wire cut specimens.
The effect of TQ on the hardness increase can be explained by the difference in the
amount of martensite formed during punching. The phase distributions were examined using
EBSD in areas equivalent to the areas where the hardness measurements were made in the
samples quenched at TQ = 90 C or TQ = 170 C (Figs. 9a, b). In both specimens, the fraction
of areas that had CI < 0.1, (i.e., black areas in Fig. 9) increased near the punched edge; this
12
the retained austenite fraction near the punched region. A sheared microstructure was present
30 μm away from the punched edge (D = 30 μm in Fig. 9); in this region, the retained
Fig. 9. EBSD phase maps of the punched region for the Q&P processed samples with
the quenching temperature of (a) 90 °C and (b) 170 °C. D = distance from the punched
edge.
In both specimens, the retained austenite fraction was reduced near the punched edge
(Fig. 10), but the amount of transformed austenite was clearly smaller in the sample heat
quenched at TQ = 90 °C than in the specimen quenched at TQ = 170 C. This sample had a
larger fraction of retained austenite in the initial stage than the specimen quenched at TQ =
90 °C. The matrix retained austenite fraction as measured by EBSD analysis was lower than
the value obtained using magnetic saturation, because the resolution of the EBSD technique
is limited to detect the small austenite islands or films that exist between the martensite
13
Fig. 10. Dependence of retained austenite volume fraction on distance from the
punched edge obtained from EBSD analysis.
TQ affected the fracture surface of cracks generated by the hole expansion test both in
the fracture surface had a characteristic dimple structure for both punched (Fig. 11a) and
wire-cut (Fig. 11b) specimens. In specimens quenched at TQ = 170 C, the fracture was
14
Fig. 11. FE-SEM fracture surface micrographs generated by the hole expansion tests:
(a) punched and (b) wire-cut specimen with TQ = 90 °C; (c) punched specimen and (d)
wire cut specimen with TQ = 170 °C. Yellow circles: intergranular fracture areas.
4. Discussion
4.1. The dependence of HER on tensile properties and retained austenite fraction
increased, and decreased as TE and retained austenite fraction (Fig. 12). The positive
correlation of HER with PUE (Fig. 12a) concurs with previous reports [14-16]. In this study,
the PUE range was 5.6 to 6.6%, which is quite small; PUE is not a good parameter that can
explain the evolution of HER of the present Q&P steel. The decrease in HER with increase in
TE (Fig. 12b) contradicts a previous report [16], which showed the opposite relationship.
Similarly, the strong increase in HER with increase in UTS increased in the range 1330 MPa
- 1400 MPa (Fig. 12c) contradicts previous HER studies, which reported a decrease in HER
with increasing UTS, with a plateau at high UTS [11, 12, 16]. Furthermore, the decrease in
HER with increase in retained austenite (Fig. 12d) contradicts several previous reports. HER
increased with retained austenite fraction in Q&P-processed 0.25C-2Mn-1.5Si steel [26], and
15
in a 0.39C-(1.2-2.0)Si-1.2Mn TRIP steel processed by intercritical annealing and
austempering [30].
Fig. 12. The evolutions of hole expansion ratio (HER) of punched specimens as a
function of key variables; (a) post-uniform elongation, (b) total elongation, (c) ultimate
tensile strength, and (d) retained austenite fraction.
The HER values obtained in the present study were also compared with HER data
obtained under similar test conditions (60 conical punch, punched hole sample) [13, 14, 16,
21-23, 26, 33, 34]; HER clearly increased as yield ratio (YS/UTS) increased (Fig. 13). HER
values obtained in the present study follow the trend, but were below the reference values.
The dependence of the HER on the tensile properties and the retained austenite fraction in
present Q&P-processed medium Mn steel differ from those reported in the literature. This
16
disagreement indicates that HER of Q&P-processed medium Mn steel is affected by factors
other than those considered here. These factors are discussed in the next section.
Fig. 13. The dependence of the hole expansion ratio (HER) on the yield ratio
(YS/UTS). The results of the punched specimens in the present study were compared
to the reference results [13, 14, 16, 21-23, 26, 33, 34] measured from similar test
conditions.
transformation to martensite was most complete near the punched hole edge where the strain
is concentrated during the hole expansion test. The amount of transformation varied with TQ.
The phase fraction obtained from the EBSD phase map was converted to a fraction of
transformed austenite. The average austenite fraction obtained from an > 1000 m away from
the punched edge was assumed to be the initial austenite fraction prior to punching. The
transformed austenite fraction corresponded with the hardness profiles in the SAZ (Fig. 14).
This similarity implies that the strain-induced martensitic transformation is the main factor
that affects the hardness increase Hv generated by punching in the SAZ.
17
Fig. 14. The comparison of the evolutions of hardness increment and transformed
fraction of retained austenite in the shear affected zone (SAZ) near the punched edge;
(a) specimen quenched to 90 C and (b) specimen quenched to 170 C.
during punching quantified. YS of each phase was predicted using a model proposed [9] for a
medium Mn Q&P steel, then converted to hardness. The alloy composition and processing
conditions in the present study were very similar to those in [9] so, the parameters (Table 1)
for the model were used here. YS of retained austenite (YS) in Q&P steel was calculated as
18
where ss, gb and dis are strengths (Table 2) contributed by solid solution, grain boundary,
specimens were 1.69 wt. % at TQ = 90 C and 0.92 wt. % at TQ = 170 C, assuming full
partitioning from martensite to austenite (Fig. 5a). The other alloying elements such as Mn,
Si and Cr were assumed not to partition. The predicted YS was 1585 MPa for samples
treated at TQ = 90 C, and 945 MPa for samples treated at TQ = 170 C. The predicted YS
Table 1. Model parameters used to calculate the yield strength of retained austenite
and strain-induced martensite adopted from [9].
Retained Strain-induced
Parameters
austenite martensite
Shear modulus, G (GPa) 72.0 81.6
Burgers vector, b (nm) 0.250 0.248
0.35 0.38
Hall-petch parameter, K (MPamm1/2) 9.49 -
19
Table. 2. The contribution of each strength parameter in MPa for both retained
austenite and strain-induced martensite calculated from two different quenching
temperature conditions.
Yield strength of the strain-induced martensite (YS’) that had been transformed
where ’0 is Peierls stress, ’ is solid-solution strengthening, and ’dis is strain hardening
(Table 2). YS’ was calculated assuming that the alloying elements do not diffuse during
transformation. Calculated YS’ was 6072 MPa at TQ = 90 C and 3695 MPa at TQ = 170 C.
170 C. The hardness difference between retained austenite and strain-induced martensite
was 1697 Hv at TQ = 90 C and 1008 Hv at TQ = 170 C. Assuming that the retained austenite
at the very edge of the punched hole is fully transformed to martensite, Hv at the tip of the
punched hole was 221 Hv at TQ = 90 C and 242 Hv at TQ = 170 C; these are both greater
20
than the measured Hv (Fig. 14), but both results show that Hv increased slightly as TQ
precipitation, and the presence of a volume fraction of small retained austenite that was not
detected by EBSD. Nonetheless, the result confirms that Hv measured in the SAZ close to
the hole edge in the punched specimens is related to the strain-induced martensite
transformation.
4.3 Key factors influencing the HER of Q&P processed medium Mn steel
In multi-phase steel, material failure starts with formation of cracks in the harder
phase constituent, and proceeds by crack growth along the interphase boundary [36]. Local
strain near the punched hole is generally much higher than the elongation measured in a
conventional tensile test. In DP steel, the local true strain applied to the very edge of a
punched hole is about 1.17 [36]. Due to the high strain near the punched edge, the punched
specimens in the present study have an elongated grain structure with a large amount of
martensite transformation (Fig. 9). The microstructure near the punched edge is mainly
phase. The two phases have a large hardness difference due to their different C
concentrations, and the strain is localized in the soft C-depleted initial martensite [22, 23, 36].
The cracks are generated preferentially at the hard C-enriched strain-induced martensite, then
prepared by punching can be explained by the fracture mechanism of the multi-phase steels.
Variation in the C concentration between the initial martensite and strain-induced martensite
results in a large hardness difference between the two phases. Thus, the amount of interfacial
21
area affects the dependence of HER on TQ. The severe strain caused by punching near the
edge results in almost full transformation of the retained austenite, and the amount of strain-
retained austenite fraction and decreased stability. In the elongated edge microstructure (Fig.
9), the amount of interphase boundary is proportional to the initial retained austenite fraction.
This observation explains the negative correlation between retained austenite fraction and
5. Conclusions
The evolution of the hole expansion ratio (HER) of a quenching and partitioning
function of quenching temperature TQ. Punched and wire-cut specimens with five TQ ranging
from 90 C to 170 C were subjected to hole expansion tests. The HER of the Q&P-processed
specimens. HER did not show a clear correlation with the tensile properties, but HER
increased as retained austenite fraction decreased. The high fraction and low stability of
retained austenite induced by quenching at high TQ resulted in an extremely low HER in the
punched specimen. At the very edge of the punched hole, the amount of retained austenite
between the high-C strain-induced martensite and the low-C martensite matrix leads to the
formation of cracks at the boundary area. The crack nucleation site density increases with TQ
22
Acknowledgements
The authors gratefully acknowledge the support of the POSCO Technical Research
Data availability
The raw/processed data required to reproduce these findings cannot be shared at this time as
the data also forms part of an ongoing study.
References
[1] J. Speer, A. Streicher, D. Matlock, F. Rizzo, G. Krauss, Quenching and partitioning: a
fundamentally new process to create high strength trip sheet microstructures, Symposium on
the Thermodynamics, Kinetics, Characterization and Modeling of: Austenite Formation and
Decomposition (2003) 505-522.
[2] J. Speer, D.K. Matlock, B.C. De Cooman, J.G. Schroth, Carbon partitioning into austenite
after martensite transformation, Acta Mater. 51 (2003) 2611-2622.
[3] A.J. Clarke, J.G. Speer, M.K. Miller, R.E. Hackenberg, D.V. Edmonds, D.K. Matlock,
F.C. Rizzo, K.D. Clarke, E. De Moor, Carbon partitioning to austenite from martensite or
bainite during the quench and partition (Q&P) process: A critical assessment, Acta Mater. 56
(2008) 16-22.
[4] A.J. Clarke, J.G. Speer, D.K. Matlock, F.C. Rizzo, D.V. Edmonds, M.J. Santofimia,
Influence of carbon partitioning kinetics on final austenite fraction during quenching and
partitioning, Scripta Materialia 61 (2009) 149-152.
[5] M.J. Santofimia, L. Zhao, R. Petrov, J. Sietsma, Characterization of the microstructure
obtained by the quenching and partitioning process in a low-carbon steel, Materials
Characterization 59 (2008) 1758-1764.
[6] E. De Moor, J.G. Speer, D.K. Matlock, J.-H. Kwak, S.-B. Lee, Effect of carbon and
manganese on the quenching and partitioning response of CMnSi steels, ISIJ international 51
(2011) 137-144.
[7] M.J. Santofimia, L. Zhao, R. Petrov, C. Kwakernaak, W.G. Sloof, J. Sietsma,
Microstructural development during the quenching and partitioning process in a newly
designed low-carbon steel, Acta Mater. 59 (2011) 6059-6068.
[8] J. Kähkönen, D.T. Pierce, J.G. Speer, E. De Moor, G.A. Thomas, D. Coughlin, K. Clarke,
A. Clarke, Quenched and Partitioned CMnSi Steels Containing 0.3 wt.% and 0.4 wt.%
Carbon, JOM 68 (2016) 210-214.
[9] E.J. Seo, L. Cho, Y. Estrin, B.C. De Cooman, Microstructure-mechanical properties
relationships for quenching and partitioning (Q&P) processed steel, Acta Materialia 113
(2016) 124-139.
[10] E.J. Seo, L. Cho, B.C. De Cooman, Application of Quenching and Partitioning
Processing to Medium Mn Steel, Metallurgical and Materials Transactions A 46 (2015) 27-
31.
[11] S. Chatterjee, H. Bhadeshia, Stretch-flangeability of strong multiphase steels, Materials
science and technology 23 (2007) 606-609.
23
[12] S. Sadagopan, D. Urban, Formability characterization of a new generation of high
strength steels, AISI/DOE technology roadmap program (2003).
[13] X. Chen, H. Jiang, Z. Cui, C. Lian, C. Lu, Hole expansion characteristics of ultra high
strength steels, Procedia Engineering 81 (2014) 718-723.
[14] J.I. Yoon, J. Jung, H.H. Lee, G.-S. Kim, H.S. Kim, Factors governing hole expansion
ratio of steel sheets with smooth sheared edge, Metals and Materials International 22 (2016)
1009-1014.
[15] J. Lee, S.-J. Lee, B.C. De Cooman, Effect of micro-alloying elements on the stretch-
flangeability of dual phase steel, Materials Science and Engineering: A 536 (2012) 231-238.
[16] S.K. Paul, Non-linear Correlation Between Uniaxial Tensile Properties and Shear-Edge
Hole Expansion Ratio, Journal of Materials Engineering and Performance 23 (2014) 3610-
3619.
[17] B.S. Levy, C.J. Van Tyne, Effect of a Strain-Hardening Rate at Uniform Elongation on
Sheared Edge Stretching, Journal of Materials Engineering and Performance 21 (2012) 2147-
2154.
[18] K.-i. Sugimoto, A. Kanda, R. Kikuchi, S.-i. Hashimoto, T. Kashima, S. Ikeda, Ductility
and Formability of Newly Developed High Strength Low Alloy TRIP-aided Sheet Steels with
Annealed Martensite Matrix, ISIJ International 42 (2002) 910-915.
[19] K.-i. Sugimoto, J. Sakaguchi, T. Iida, T. Kashima, Stretch-flangeability of a High-
strength TRIP Type Bainitic Sheet Steel, ISIJ International 40 (2000) 920-926.
[20] N. Pathak, C. Butcher, M. Worswick, Influence of the sheared edge condition on the
hole expansion of dual phase steel, Proc. IDDRG2013 (Zürich) (2013) 213-218.
[21] O.R. Terrazas, Correlation of microstructure, tensile properties and hole expansion ratio
in cold rolled advanced high strength steels, Master thesis in Colorado School of Mines
(2016).
[22] K. Hasegawa, K. Kawamura, T. Urabe, Y. Hosoya, Effects of Microstructure on Stretch-
flange-formability of 980 MPa Grade Cold-rolled Ultra High Strength Steel Sheets, ISIJ
International 44 (2004) 603-609.
[23] A. Karelova, C. Krempaszky, E. Werner, P. Tsipouridis, T. Hebesberger, A. Pichler,
Hole Expansion of Dual‐phase and Complex‐phase AHS Steels‐Effect of Edge Conditions,
steel research international 80 (2009) 71-77.
[24] K.-i. Sugimoto, A. Nagasaka, M. Kobayashi, S.-i. Hashimoto, Effects of Retained
Austenite Parameters on Warm Stretch-flangeability in TRIP-aided Dual-phase Sheet Steels,
ISIJ International 39 (1999) 56-63.
[25] E. De Moor, D.K. Matlock, J.G. Speer, C. Fojer, J. Penning, Comparison of Hole
Expansion Properties of Quench & Partitioned, Quench & Tempered and Austempered
Steels, SAE Technical Paper, 2012.
[26] R. Johnson, E. De Moor, N. Fonstein, D. Hanlon, A. Pichler, Retained austenite effects
on hole expansion of ultra-high strength steels, Proceedings of the International Symposium
on New Developments in Advanced High Strength Sheet Steels. PA: AIST, 2013.
[27] I.O.f. Standardization, Metallic materials-sheet and strip-hole expanding test, ISO
16630, International Organization for Standardization, Switzerland, 2009.
[28] X.C. Xiong, B. Chen, M.X. Huang, J.F. Wang, L. Wang, The effect of morphology on
the stability of retained austenite in a quenched and partitioned steel, Scripta Materialia 68
(2013) 321-324.
[29] L. Chen, J.K. Kim, S.K. Kim, G.S. Kim, K.G. Chin, B. De Cooman, Stretch‐
Flangeability of High Mn TWIP steel, steel research international 81 (2010) 552-568.
24
[30] O. Matsumura, Y. Sakuma, Y. Ishii, J. Zhao, Effect of retained austenite on formability
of high strength sheet steels, ISIJ international 32 (1992) 1110-1116.
[31] R.G. Davies, Edge cracking in high strength steels, Journal of Applied Metalworking 2
(1983) 293-299.
[32] D. De Knijf, R. Petrov, C. Föjer, L.A.I. Kestens, Effect of fresh martensite on the
stability of retained austenite in quenching and partitioning steel, Materials Science and
Engineering: A 615 (2014) 107-115.
[33] J.I. Yoon, J. Jung, S.-H. Joo, T.J. Song, K.-G. Chin, M.H. Seo, S.-J. Kim, S. Lee, H.S.
Kim, Correlation between fracture toughness and stretch-flangeability of advanced high
strength steels, Materials Letters 180 (2016) 322-326.
[34] I. Pushkareva, S. Allain, C. Scott, A. Redjaïmia, A. Moulin, Relationship between
Microstructure, Mechanical Properties and Damage Mechanisms in High Martensite Fraction
Dual Phase Steels, ISIJ International 55 (2015) 2237-2246.
[35] E.J. Pavlina, C.J. Van Tyne, Correlation of Yield Strength and Tensile Strength with
Hardness for Steels, Journal of Materials Engineering and Performance 17 (2008) 888-893.
[36] B.S. Levy, M. Gibbs, C.J. Van Tyne, Failure during Sheared Edge Stretching of Dual-
Phase Steels, Metallurgical and Materials Transactions A 44 (2013) 3635-3648.
25