ScriptaMater WJF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/309545582

Strengthening effect of nanoscale precipitation and transformation induced


plasticity in a hot rolled copper-containing ferrite-based lightweight steel

Article  in  Scripta Materialia · March 2017


DOI: 10.1016/j.scriptamat.2016.10.025

CITATIONS READS

12 41

6 authors, including:

Xiaodong Wang Xj Jin


Shanghai Jiao Tong University Shanghai Jiao Tong University
72 PUBLICATIONS   1,176 CITATIONS    167 PUBLICATIONS   2,414 CITATIONS   

SEE PROFILE SEE PROFILE

Li Wang
University of Westminster
41 PUBLICATIONS   842 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

High-entropy Alloys Strengthened by Coherent Precipitates View project

Toughening Mechanism and Microstructure Design of Ultrahigh Stength Press Hardening Steels for Automobiles View project

All content following this page was uploaded by Xiaodong Wang on 28 March 2021.

The user has requested enhancement of the downloaded file.


Scripta Materialia 129 (2017) 25–29

Contents lists available at ScienceDirect

Scripta Materialia

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

Regular article

Strengthening effect of nanoscale precipitation and transformation


induced plasticity in a hot rolled copper-containing ferrite-based
lightweight steel
Junfeng Wang a, Zhaoguang Wang a, Xiaodong Wang a,⁎, Qi Yang b,c,⁎⁎, Xuejun Jin a, Li Wang b,c
a
State Key Laboratory of Metal Matrix Composites, School of Materials Science and Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
b
Baoshan Iron & Steel Co., Ltd., Shanghai 201900, China
c
State Key Lab of Development and Application Technology of Automotive Steels (Baosteel), Shanghai 201900, China

a r t i c l e i n f o a b s t r a c t

Article history: Effects of copper addition on microstructure and tensile property of a hot rolled ferrite-based lightweight steel
Received 29 July 2016 are investigated. Copper addition delays bainitic transformation, resulting in higher volume fraction of retained
Received in revised form 20 October 2016 austenite with lower stability in the steel. Addition of 1.0 wt% copper increases the yield strength and tensile
Accepted 20 October 2016
strength of the steel by 17% and 19%, respectively, and maintains high ductility. The yield strength enhancement
Available online xxxx
is predominantly attributed to nanoscale copper precipitation. The higher work hardening rate due to precipita-
Keywords:
tion hardening and the transformation-induced-plasticity mechanism leads to higher tensile strength (805 MPa)
Ferrite-based lightweight steel and remarkable elongation-to-failure (~30%) for the copper-containing steel.
Bainitic transformation © 2016 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Copper precipitation strengthening
Mechanical property

Aluminum-enriched transformation-induced-plasticity (TRIP) undergo continuous nanoscale precipitation during a proper isothermal
steels, also called δ-TRIP steels or ferrite-based lightweight steels, hold or aging process, with the crystal structure of precipitates
could possess lower densities, excellent combinations of strength and transitioned from coherent body-centered cubic (BCC) through
ductility, and appreciable producibility [1–10], and therefore have twinned 9R to face-centered cubic (FCC) [22]. The effects of Cu addition
great potential for automotive applications. As a characteristic constitu- on microstructure as well as strength enhancement without noticeable
ent phase, δ-ferrite forms during solidification and persists during sub- ductility deterioration in TRIP-assisted steels had been studied [23],
sequent thermo-mechanical processing and heat treatment due to the however the microstructural features including the Cu precipitation,
addition of a large amount of aluminum [2,4,6]. Moreover, it is hardly bainitic transformation, and retained austenite (RA) stability need to
refined by the austenite-to-ferrite transformation and recrystallization. be further elaborately clarified to interpret the correlation between mi-
Correspondingly, the δ-ferrite is oftentimes pretty coarse, and softer crostructure and mechanical behavior. This is one of the objectives of
than α-ferrite and fine austenite [11]. Strengths of ferrite-based light- the present investigation which is to be performed on hot rolled
weight steels are therefore relatively low owing to the presence of a ferrite-based lightweight steels containing δ-ferrite. The other objective
large fraction of the δ-ferrite phase in the microstructure. is to confirm the feasibility of utilizing Cu precipitates to strengthen
It is well known that precipitation strengthening is a useful ap- ferrite-based lightweight steels and to attain an excellent combination
proach to improving the strengths of steels. Microalloying with strong of strength and ductility in the meantime.
carbide-forming elements such as Ti, Nb, Mo and V could remarkably in- Two ferrite-based lightweight steels were fabricated for the investi-
crease yield strengths and/or tensile strengths of ferritic steels [12–15] gation. The chemical compositions of the steels are Fe-0.35C-1.1Mn-
and TRIP-assisted multiphase steels [16,17] through precipitation 4.1Al-0.4Si-(0\1.0)Cu (in wt%) with traces of nitrogen, sulfur and phos-
strengthening as well as other microstructural modifications (such as phor. The two steels were referred to as “Cu-free” and “1.0Cu”, respec-
grain refinement, bainitic transformation, etc.), however the ductility tively hereafter for brevity. Steel ingots were homogenized at 1200 °C
of the steels may be deteriorated. It has been another important strate- for 2.5 h, and rough rolled from 180 mm to 40 mm, followed by air
gy to strengthen the above types of steels by utilizing copper (Cu)-rich cooling. The slabs were then reheated to 1200 °C for 60 min and hot-
precipitates [18–21]. Because of limited solubility in Fe, Cu solutes could rolled into plates of 3.6 mm with a finishing temperature of approxi-
mately 880 °C. The rolled plates were immediately water quenched to
⁎ Corresponding author.
450 °C, and then held at this temperature in a furnace for 60 min, and
⁎⁎ Correspondence to: Qi Yang, Baoshan Iron & Steel Co., Ltd., Shanghai 201900, China. finally furnace cooled to room temperature in order to simulate a coiling
E-mail addresses: xdwang77@sjtu.edu.cn (X. Wang), yangqi@baosteel.com (Q. Yang). procedure. After cooling, the hot-rolled plates were pickled in a dilute

http://dx.doi.org/10.1016/j.scriptamat.2016.10.025
1359-6462/© 2016 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
26 J. Wang et al. / Scripta Materialia 129 (2017) 25–29

HCl solution to remove the surface oxide film for microstructural and variation of the volume fraction of RA, XRD measurements were made
mechanical analyses. in the uniformly-deformed regions of fractured specimens and the
Tensile test samples were cut along the rolling direction (RD) with a necked regions which are 3 mm away from the fracture surfaces. Tensile
gauge length and width of 50 mm and 12.5 mm, respectively (according true strains, ε, were calculated as follows:
to the ASTM E8-04 standard). Tests were performed on a computer-  
controlled 8801 Instron machine with a constant crosshead speed of w0 t 0
ε ¼ ln ð2Þ
3.0 mm/min (corresponding to an initial strain rate of 8.8 × 10−4/s) at wt
room temperature. At least three samples were tested to obtain average
tensile property for each steel. where w0 and t0 are the width and thickness of tensile specimens before
Microstructures on the plane of the plate RD and normal direction tensile test, respectively; w and t are the width and thickness of tensile
(ND) were observed using optical microscope (OM) and field- specimens in the uniformly-deformed regions or necked regions after
emission scanning electron microscope (SEM; FEI Versa 3D). Metallo- tensile test, respectively. It should be noted that the necked region is
graphic specimens were mechanically polished and etched in a 4% in the three-dimensional stress state (rather than the uniaxial tensile
nital solution. Electron back-scattered diffraction (EBSD) analysis was stress state as in the uniformly-deformed region), and the strain value
carried out on the RD-ND plane of the hot-rolled steel plates as well, calculated by Eq. (2) roughly represents tensile limit strain.
using Hitachi Su-70 Schottky field emission SEM at an acceleration volt- Fig. 1 shows typical OM and SEM micrographs, and EBSD phase maps
age of 20 kV. EBSD data were post-processed by the HKL CHANNEL 5 fla- of the Cu-free and 1.0Cu steels, respectively. The hot-rolled microstruc-
menco software. Fine microstructures were further characterized by tures are basically banded structures in which lighter bands are aggre-
transmission electron microscope (TEM; JEOL ARM-200F) equipped gates of recovered/recrystallized δ-ferrite grains, and darker bands are
with a Cs corrector, operating at an acceleration voltage of 200 kV. transformation products of prior austenite during the isothermal hold
TEM foils were prepared by twin-jet electropolishing punched disk and furnace cooling [25]. Darker bands show different variants and mor-
specimens (3 mm in diameter and 80 μm in thickness) in a 4 vol% phologies, depending on whether Cu is added. In the Cu-free steel, the
perchloric acid ethanol solution at −30 °C and with a voltage of about darker bands mainly include carbide-free bainite which is composed of
50 V. Phase identification was made by an X-ray diffractometer (XRD, lath-shaped bainitic ferrite and austenite, in addition to polygonal α-
Bruker-AXS D8 Discover) with Cu Kα radiation operated at 40 kV and ferrite grains and some blocky austenite particles (Fig. 1(a, b)). Martensite
40 mA at room temperature. Volume fractions of retained austenite is hardly observed in the microstructure, as indicated by the EBSD phase
(RA) were measured from the Rietveld refinement of XRD profiles to map (Fig. 1(c)). It is thus implied that the RA after bainitic transformation
take account of the texture effect. The carbon concentration of the RA reaction is quite thermally stable. In the 1.0Cu steel, the darker bands are
was determined using the following equation [24], assuming that the mainly composed of blocky austenite/martensite packets (Fig. 1(d, e)).
partitioning effect of substitutional elements can be neglected, Some of prior austenite transform into feathery bainitic structure during
the simulated coiling process, which is illustrated in black regions in
aγ ¼ 3:578 þ 0:033X C þ 0:00095X Mn þ 0:0056X Al þ 0:00157X Si Fig. 1(d) and by the inset magnified view in Fig. 1(e), and will be further
þ 0:0015X Cu ð1Þ confirmed by TEM analysis. The formation of martensite (M/A islands) as
verified by the EBSD phase map (Fig. 1(f)) is ascribed to the transforma-
where aγ is the austenite lattice parameter in Å, and XC, XMn, XAl, XSi and tion of large and relatively thermally-unstable austenite when the tem-
XCu are the concentrations of C, Mn, Al, Si and Cu in steels, respectively, perature drops below Ms. (martensite start temperature) during furnace
in wt%. The austenite lattice parameter (aγ) was determined from the d- cooling. According to EBSD statistical measurements from at least 100
spacing of (220)γ positions. To evaluate the effect of tensile strain on the grains for each phase, the grain sizes for δ-ferrite, α-ferrite and RA are

Fig. 1. OM and SEM micrographs and EBSD phase maps of the hot rolled lightweight steels: (a–c) Cu-free steel; (d–e) 1.0Cu steel. SEM micrographs and EBSD maps show transformation
products of prior austenite in darker bands in more detail. The step size for EBSD mapping is 70 nm. The area in red color depicts RA. (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.)
J. Wang et al. / Scripta Materialia 129 (2017) 25–29 27

11.5 ± 5.8 μm, 2.8 ± 0.8 μm and 2.2 ± 1.1 μm, respectively in the Cu-free and RA in the 1.0Cu steel are finer than those in the Cu-free steel.
steel, and 11.1 ± 5.3 μm, 2.4 ± 0.9 μm and 2.5 ± 1.0 μm, respectively in From TEM analysis, the above fine structure is unlikely to be originated
the 1.0Cu steel. The presence of RA is attributed to ‘incomplete reaction from martensitic transformation due to a thermal effect, considering
phenomenon’ of bainitic transformation when the austenite carbon con- that no twinned substructure is observed within prior austenite having
tent approaches the T0 line [16,26]. XRD measurements show that the vol- such high carbon content. It is unclear that the bainitic structure in the
ume fraction of RA is 17.1% in the 1.0Cu steel, and 9.6% in the Cu-free steel. 1.0Cu steel is finer than that in the Cu-free steel. Imaginably, the coordi-
From Eq. (1), the average carbon concentration of RA is 2.08 wt% in the nation of bainitic transformation and interphase Cu precipitation might
Cu-free steel, and 1.85 wt% in the 1.0Cu steel. slow down the growth of bainitic ferrite.
Cu is an austenite stabilizer. A simple thermodynamic calculation in- Cu clustering/precipitation inside ferrite grains in the 1.0Cu steel is
dicates that the addition of 1.0 wt% Cu does not cause an appreciable investigated.
change in the volume fraction of austenite at high temperatures. Thus, Fig. 3(a) shows the high-angle annular dark-field scanning transmis-
Cu addition increases thermal stability of high-temperature austenite. sion electron microscopy (HAADF STEM) image of δ-ferrite grains. High-
The prior austenite with increased thermal stability could be more read- density ultra-fine particles and sparse dislocation lines are observed in-
ily retained in a blocky shape by delaying bainitic reaction or without side δ-ferrite. The sparse distribution of dislocations inside the δ-ferrite
experiencing substantial bainitic reaction during an isothermal hold grain interior indicates that δ-ferrite undergoes quite adequate recov-
(subsequent to finish rolling and water quenching), although some of ery/recrystallization at high rolling temperature. As seen in Fig.
the austenite with larger grain sizes would eventually transform into 3(b) which depicts a magnified view of the boxed region in Fig. 3(a), a
martensite as the temperature falls below Ms. (Fig. 1(f)). This could be large number of precipitating particles (shown by white dots) at a nano-
the reason why a larger volume fraction of blocky RA is present in the meter scale gather along the dislocation line. Cu precipitates have a
1.0Cu steel. In comparison, the Cu-free steel could undergo adequate closely-packed face-centered cubic (FCC) crystal structure, as demon-
bainitic transformation reaction during an isothermal hold because of strated by the atomic-resolution HAADF STEM image and the inset cor-
the lower thermal stability of high-temperature austenite. Thus, the responding fast Fourier transform (FFT) pattern in Fig. 3(c). There exists
RA is more carbon enriched due to the carbon partitioning from the a N-W orientation relationship between the precipitates and matrix:
neighboring bainitic ferrite, and its volume fraction is lower in the Cu- ð111ÞFCC−Cu ==ð110Þα and [011]FCC − Cu//[001]α. The inverse FFT (IFFT)
free steel. Compared to the Cu-free steel, the 1.0Cu steel possesses image by masking {111} FCC reflection is shown in Fig. 3(d), and statis-
higher volume fraction of RA with lower carbon content after the simu- tical measurements indicate that the sizes of Cu precipitates are mainly
lated coiling process. smaller than 3–5 nm. The densely distributed Cu precipitates inside the
Fine microstructure formed during the bainitic hold in the 1.0Cu ferritic matrix would undoubtedly affect mechanical property of the
steel is further characterized. Fig. 2(a) illustrates a TEM bright-field steel.
image of a bainitic sheaf. As shown in the corresponding center dark- Room temperature tensile stress-strain curves and strain hardening
field images (Fig. 2(b, c)) and by the selected area electron (SAED) pat- rate curves (as an inset) of the two steels are shown in Fig. 4(a). The
tern (Fig. 2(d)), RA films exist in between bainitic ferrite plates, and 1.0Cu steel possesses the yield strength of 563 MPa and tensile strength
there are no carbides precipitating within or between bainitic ferrite of 805 MPa, higher than those of the Cu-free specimen by 83 MPa
plates. This implies that the supersaturated carbon in the bainitic ferrite (i.e., 17% increment) and 130 MPa (i.e., 19% increment), respectively,
diffuses into the surrounding austenite during the transformation reac- while the elongation-to-failure of these two steels are comparable,
tion and then stabilizes the remaining austenite. The crystallographic about 30%. In addition, the 1.0Cu steel maintains higher strain hardening
orientation relationship between bainitic ferrite and RA film follows rate than the Cu-free steel at true strains ranging from 0.03 to 0.24. Me-
the N-W relationship: ð111Þγ ==ð110ÞB F and [011]γ//[001]BF. Bainitic fer- chanical properties of the two steels are listed in Table 1 in detail.
rite plates are less than 200 nm in thickness, and RA films are about a The strain hardening rate is closely related to the transformation be-
few tens of nanometer in thickness. In morphology, this bainitic struc- havior of RA during tensile deformation in TRIP steels. Fig. 4(b) shows
ture is basically carbide-free bainite, except that laths of bainitic ferrite the variation of the volume fraction of RA as a function of true strain

Fig. 2. TEM micrographs of the 1.0Cu steel: (a) bright-field image of bainite; (b) g200Bf dark-field image of bainitic ferrite; (c) g200γ dark-field image of film-like retained austenite; (d) SAED
pattern of bainite.
28 J. Wang et al. / Scripta Materialia 129 (2017) 25–29

Fig. 3. (a) STEM image of δ-ferrite grain; (b) magnified view of the boxed region in (a); (c) atomic resolution HAADF STEM image (FFT pattern inset); (d) IFFT image acquired by masking
{111} FCC reflections.

for the present two steels. As deformation proceeds, in particular ensures continuous martensitic transformation at varied stress levels,
throughout uniform strain, more RA particles are transformed at a faster thus providing continuous strain hardening of TRIP steels [27,28]. That
rate in the 1.0Cu steel. This means that RA is mechanically less stable in is to say, transformation of RA with larger grain sizes and/or a blocky
the 1.0Cu steel, although the transformed volume fraction of RA is shape (i.e., with lower mechanical stability) primarily provides strain
higher than that in the Cu-free steel. The mechanical stability of RA in hardening at a lower stress level, while transformation of RA with
the 1.0Cu steel lower than that of RA in the Cu-free steel could be as- finer grain size and/or a lath shape (i.e., with higher mechanical stabili-
cribed to the joint effect of primarily blocky shape, slightly larger grain ty) provides strain hardening at a higher stress level. This is believed to
size and lower carbon concentration of RA in the 1.0Cu steel. However, be true for the present two steels.
the higher transformed volume fraction together with lower mechani- As a TRIP steel is loaded, yielding starts from the softer phase, ferrite
cal stability of RA, probably assisted by Cu precipitation, causes the when RA has fine grain size and good mechanical stability [3,11,29]. Fur-
1.0Cu steel to maintain a higher strain hardening rate up to high strains thermore, δ-ferrite is believed to be the softer ferrite since it has a larger
(Fig. 4(a)) and thus achieve a remarkably high elongation-to-failure. grain size (compared to α-ferrite) and undergoes quite adequate recov-
Furthermore, as generally accepted, tailored mechanical stability of RA ery/recrystallization (Fig. 1 and Fig. 3). The contributions to the

Fig. 4. (a) Room-temperature engineering stress-strain relationship of the hot-rolled steels (strain hardening rate curves inset); (b) variation of RA volume fraction as a function of true
strain.
J. Wang et al. / Scripta Materialia 129 (2017) 25–29 29

Table 1 Acknowledgment
Mechanical properties of the both steels.

Yield Tensile Uniform Total The authors are grateful for financial support from the National Nat-
strength/MPa strength/MPa elongation/% elongation/% ural Science Foundation of China (grant no. 51271112) and Research
Cu-free 480 675 24 30.5 Fund from Baoshan Iron & Steel Co., Ltd.
1.0Cu 563 805 24.6 29.8
Appendix A. Supplementary data

difference in the yield strengths of the present two steels would only Supplementary data to this article can be found online at doi:10.
arise from the Cu solid-solution strengthening and Cu precipitation 1016/j.scriptamat.2016.10.025.
strengthening of δ-ferrite, since the two steels undergoing adequate re-
covery/recrystallization have comparable δ-ferrite grain sizes (the ef-
fects of grain boundary strengthening and dislocated substructure References
strengthening can thus be ignored). The Cu solubility limit in iron in
[1] H.L. Yi, K.Y. Lee, H.K.D.H. Bhadeshia, Phys. Eng. Sci. 467 (2010) 234–243.
the range of 500–700 °C is given in the following equation [30]: [2] S. Chatterjee, M. Murugananth, H.K.D.H. Bhadeshia, Mater. Sci. Technol. 23 (2007)
  819–827.
log10 ½Cuwt% ¼ 5771323 T2
−15763:84 T
þ 9:944961 ð3Þ [3] H.L. Yi, K.Y. Lee, H.K.D.H. Bhadeshia, Mater. Sci. Eng. A 528 (2011) 5900–5903.
[4] H.L. Yi, K.Y. Lee, H.K.D.H. Bhadeshia, Mater. Sci. Technol. 27 (2011) 525–529.
[5] H.L. Yi, J.H. Ryu, H.K.D.H. Bhadeshia, H.W. Yen, J.R. Yang, Scr. Mater. 65 (2011)
where T is the absolute temperature. By extrapolation, the Cu solubility 604–607.
at 450 °C is about 0.15 wt%, and becomes even less during furnace [6] Y.J. Choi, D.W. Suh, H.K.D.H. Bhadeshia, Phys. Eng. Sci. 468 (2012) 2904–2914.
cooling. Therefore, an increase in the yield strength due to the Cu [7] H.L. Yi, P. Chen, Z.Y. Hou, N. Hong, H.L. Cai, Y.B. Xu, D. Wu, G.D. Wang, Scr. Mater. 68
(2013) 370–374.
solid-solution strengthening (equal to 38 MPa/wt% × Cu solubility in [8] H.L. Yi, P. Chen, H.K.D.H. Bhadeshia, Metall. Mater. Trans. A 45 (2014) 3512–3518.
wt% [31]) is less than 5.7 MPa. From the above analysis, it is clearly dem- [9] S.S. Sohn, B.J. Lee, S. Lee, N.J. Kim, J.H. Kwak, Acta Mater. 61 (2013) 5050–5066.
onstrated that the higher yield strength of the 1.0Cu steel predominant- [10] S.J. Park, B. Hwang, K.H. Lee, T.H. Lee, D.W. Suh, H.N. Han, Scr. Mater. 68 (2013)
365–369.
ly arises from the Cu precipitation. It should be pointed out that the yield
[11] K. Lee, S.-J. Park, Y.S. Choi, S.-J. Kim, T.-H. Lee, K.H. Oh, H.N. Han, Scr. Mater. 69
strength enhancement due to precipitation hardening is theoretically (2013) 618–621.
calculated to be more than 138 MPa according to the dislocation- [12] U. Brüx, G. Frommer, Steel Res. Int. 73 (2002) 534–548.
[13] A. Zargaran, H.S. Kim, J.H. Kwak, N.J. Kim, Scr. Mater. 89 (2014) 37–40.
precipitate cutting model (shown by eq. (3) in reference [20]) and to
[14] S.G. Hong, H.J. Jun, K.B. Kang, C.G. Park, Scr. Mater. 48 (2003) 1201–1206.
be more than 317 MPa according to the dislocation-precipitate bowing [15] H.-W. Yen, C.-Y. Huang, J.-R. Yang, Scr. Mater. 61 (2009) 616–619.
model (shown by Eq. (4) in reference [20]), both much higher than the [16] X.D. Wang, B.X. Huang, L. Wang, Y.H. Rong, Metall. Mater. Trans. A 39 (2008) 1–7.
measured value of about 80 MPa. The fact that the strengthening of Cu [17] G.K. Tirumalasetty, M.A. van Huis, C.M. Fang, Q. Xu, F.D. Tichelaar, D.N. Hanlon, J.
Sietsma, H.W. Zandbergen, Acta Mater. 59 (2011) 7406–7415.
precipitates can not be predicted by the dislocation-precipitate interac- [18] B. Mishra, K. Kumbhar, K.S. Kumar, K.S. Prasad, M. Srinivas, Mater. Sci. Eng. A 651
tion models might be attributed to the modulus difference between Cu (2016) 177–183.
precipitate and ferrite or an interaction between screw dislocation cores [19] J. Takahashi, K. Kawakami, Y. Kobayashi, Mater. Sci. Eng. A 535 (2012) 144–152.
[20] Y.R. Wen, Y.P. Li, A. Hirata, Y. Zhang, T. Fujita, T. Furuhara, C.T. Liu, A. Chiba, M.W.
and precipitates with the sizes of a few nanometers [20,32]. The higher Chen, Acta Mater. 61 (2013) 7726–7740.
tensile strength of the 1.0Cu steel is due to the composite effect of pre- [21] M. Kapoor, D. Isheim, G. Ghosh, S. Vaynman, M.E. Fine, Y.-W. Chung, Acta Mater. 73
cipitation hardening and higher strain hardening rate (Fig. 4). (2014) 56–74.
[22] Y.R. Wen, A. Hirata, Z.W. Zhang, T. Fujita, C.T. Liu, J.H. Jiang, M.W. Chen, Acta Mater.
In summary, the effects of Cu addition on microstructure evolution 61 (2013) 2133–2147.
and mechanical behavior in a hot rolled ferrite-based lightweight steel [23] S.-J. Kim, C.G. Lee, T.-H. Lee, C.-S. Oh, ISIJ Int. 42 (2002) 1452–1456.
are investigated. Cu addition improves thermal stability of high- [24] D. Dyson, B. Holmes, J. Iron Steel Inst. 208 (1970) 469–474.
[25] J. Wang, Q. Yang, X. Wang, L. Wang, 2016. Metall. Mater. Trans. A. , http://dx.doi.org/
temperature austenite and subsequently delays bainitic transformation
10.1007/s11661-016-3752-8.
at 450 °C. As a result, a higher volume fraction of RA primarily in a [26] H.K.D.H. Bhadeshia, J.W. Christian, Metall. Trans. A 21 (1990) 767–797.
blocky shape but with relatively lower thermal/mechanical stability re- [27] Z.H. Cai, H. Ding, X. Xue, J. Jiang, Q.B. Xin, R.D.K. Misra, Scr. Mater. 68 (2013)
865–868.
mains in the ferrite matrix of the Cu-containing steel during furnace
[28] J.H. Ryu, J.I. Kim, H.S. Kim, C.-S. Oh, H.K.D.H. Bhadeshia, D.-W. Suh, Scr. Mater. 68
cooling to room temperature. Addition of 1.0 wt% Cu increases the (2013) 933–936.
yield strength and tensile strength of the steel by 83 MPa (i.e., 17% incre- [29] J.H. Ryu, D.-I. Kim, H.S. Kim, H.K.D.H. Bhadeshia, D.-W. Suh, Scr. Mater. 63 (2010)
ment) and 130 MPa (i.e., 19% increment), respectively, but maintains 297–299.
[30] M. Perez, F. Perrard, V. Massardier, X. Kleber, A. Deschamps, H. de Monestrol, P.
high ductility. The increment of the yield strength is predominantly at- Pareige, G. Covarel, Philos. Mag. 85 (2005) 2197–2210.
tributed to nanoscale Cu precipitation. The cooperation of precipitation [31] Q.L. Yong, Secondary Phases in Steels, Metallurgy Industry Press, Beijing, 2006.
hardening and the TRIP mechanism results in the higher strain harden- [32] M.E. Fine, D. Isheim, Scr. Mater. 53 (2005) 115–118.

ing rate, and therefore higher tensile strength (805 MPa) and remark-
able elongation (~30%) for the 1.0Cu steel.

View publication stats

You might also like