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

Materials Characterization 153 (2019) 108–114

Contents lists available at ScienceDirect

Materials Characterization
journal homepage: www.elsevier.com/locate/matchar

Microstructure and mechanical properties of heavily cold drawn pearlitic T


steel wires: Effects of low temperature annealing

Dasheng Weia, Long Lib, Xuegang Minb, Feng Fanga, , Zonghan Xiec,d, Jianqing Jianga
a
Jiangsu Key Laboratory of Advanced Metallic Materials, Southeast University, Nanjing 211189, China
b
Jiangsu Baosteel Fine Wire&Cord Co., Ltd., Haimen 226114, China
c
School of Materials and Energy, Southwest University, Chongqing 400715, China
d
School of Mechanical Engineering, University of Adelaide, SA 5005, Australia

A R T I C LE I N FO A B S T R A C T

Keywords: Heavily cold drawn pearlitic steel wires prepared with drawing strains (ε) of 4.5, 5.1 and 5.3 were annealed at
Pearlitic steel wires low temperatures for different durations. Microstructural characteristics and mechanical properties of the an-
Low temperature annealing nealed wires were investigated and determined. The tensile strength of cold drawn pearlitic steel wires was
Tensile strength found to increase, due to low temperature annealing treatment. For example, for the cold drawn wires (ε = 5.3)
Ductility
annealed at 60 °C for 1 h, the tensile strength is enhanced from about 4920 MPa to 5020 MPa. Counterintuitively,
Carbon segregation
the ductility shows little decrease. With the increase of annealing temperature and time, the tensile strength
decreases. For the annealed wires, carbon atoms were observed to segregate at subgrain boundaries (SGBs), and
particle precipitation can also be observed. Substitutional solid solution atoms, such as Cr, Mn and Si, remain
locked in their original positions. Exothermic peaks in differential scanning calorimetry (DSC) curves were found
to shift to higher temperatures for the annealed wires.

1. Introduction During cold drawing process, cementite plates first turn into nano-
filaments, before forming amorphous cementite [21,22]. Under heavy
Ultrahigh strength pearlitic steel wires have been used for a wide drawing, chemical decomposition of cementite would occur [5,23,24].
range of engineering applications, such as automotive tyre cords, sus- It was reported that under cold drawing (ε > 3.5), pearlitic or lamellar
pension bridge cables and sawing wires [1]. They are routinely pro- structure evolved into a subgrain structure divided by dislocation
duced by cold drawing, during which a reduction in interlamellar bundles [2,17]. As the drawing strain ε is > 5.1, most of cementite in
spacing [2,3], ferrite deformation accompanied by an increase in dis- cold drawn pearlitic steel wires decomposed [25]. As heavily cold
location density [5,6] and/or cementite decomposition [4,5] would drawn pearlitic steel wires were annealed at lower temperatures
take place, resulting in a significant increase in the strength of steel (< 150 °C), carbon atoms from decomposed cementite segregated to
wires. Relationships between the microstructural evolution and me- dislocations in ferrite and/or at ferrite/cementite interfaces, leading to
chanical properties of cold drawn pearlitic steel wires have been stu- the increases of the wire strength [10,19]. However, a comprehensive
died extensively [2–20]. Through the control of the microstructural understanding of the evolution of microstructure and mechanical
development, the tensile strength of the wires can reach up to 7 GPa properties of heavily cold drawn pearlitic steel wires during the low
[17]. temperature annealing process remains elusive.
Cold drawn pearlitic steel wires are sensitive to heat treatment. For In this work, the microstructural features and mechanical properties
instance, cementite lamellae can assume a spherical shape when being of cold-drawn pearlitic steel wires were investigated, following low
heated to above 400 °C for a certain period of time. As a consequence, temperature annealing. Strengthening mechanisms of cold drawn
the strength of the heat-treated wires decreases with an increase in pearlitic steel wires during low temperature annealing treatment was
plasticity [13,16]. Annealing at relatively lower temperatures revealed and discussed.
(200–400 °C) however could transform amorphous cementite in cold
drawn steel wires to nano-crystalline grains. This would lead to an in-
crease in both the strength and ductility of the steel wires [20].


Corresponding author at: School of Materials Science and Engineering, Southeast University, Jiangning District, Nanjing 211189, China.
E-mail address: fangfeng@seu.edu.cn (F. Fang).

https://doi.org/10.1016/j.matchar.2019.05.003
Received 26 February 2019; Received in revised form 3 May 2019; Accepted 3 May 2019
Available online 05 May 2019
1044-5803/ © 2019 Elsevier Inc. All rights reserved.
D. Wei, et al. Materials Characterization 153 (2019) 108–114

Table 1 accelerated voltage of 200 kV.


Chemical composition of high carbon steel rods used in this work. Atom probe tomography (APT) experiments were performed using a
Element C Mn Si Cr P S Cu Fe local electrode atom probe (LEAP 4000XHR, Cameca Instruments) at a
residual pressure of 3 × 10−9 Pa, a pulse-voltage to dc-voltage ratio of
wt.% 0.92 0.30 0.17 0.20 0.009 0.009 0.006 Bal. 15%, a pulse repetition frequency of 200 kHz, and a specimen tem-
perature of −223 °C. The specimens for APT were prepared using a
standard two-stage electro-polishing method at room temperature [26].
2. Materials and methods APT data analysis was conducted using Imago Visualization and Ana-
lysis Software (IVAS 3.6.14) software. Carbon atoms were collected as
High carbon steel rods used in this study were produced by Nippon single (C+, C2+) and molecular ions (C 32+, C 2+, C 42+, C 3+). All
Steel Corporation, Japan. Their chemical compositions are shown in these different ions were displayed within reconstructed volumes.
Table 1. The steel rods with the initial diameter of 0.66 mm were drawn Carbon quantification was made according to a previous study [27].
successively to 0.07 mm, achieving a true drawing strain of ε = 4.5. In The mean carbon composition of reconstructed volumes is
addition, the steel rods of 0.58 mm in diameter were drawn successively 4.15 ± 0.03 at.%, which is in agreement with the nominal composition
to final diameters of 0.045 mm and 0.040 mm, yielding respective of 4.14 at.% (i.e., 0.92 wt%). The isosurfaces were drawn using IVAS
drawing strains of about 5.1 and 5.3. The cold drawn wires were an- software to identify the existence of precipitates in steel [26,28].
nealed at 60 °C and 100 °C for different periods of time (0.5 h, 1 h, 2 h,
6 h and 12 h) in oil bath. The annealed wires were cooled to room
temperature in air. 3. Results and discussion
The tensile strength of the wires was determined using a universal
tensile testing machine (INSTRON LEGEND 2344) at a strain rate of 3.1. Mechanical properties
3.33 × 10−3 s−1. Torsion tests were conducted with a torsion tester
(KNTEST EZ-1). The torque load and rotational speed was set at 5 N and Fig. 1a shows the tensile strength of cold drawn pearlitic steel wires
1 round/s, respectively. X-ray diffraction (XRD) analysis was performed annealed at 60 °C for different durations. The tensile strength of as-
(Rigaku D-max 2100) using a diffractometer with Cu target at a scan- drawn wires (ε = 4.5) is also given as about 4243 MPa. After being
ning speed of 0.5°/min and step size of 0.001°. Background reduction, annealed at 60 °C for 1 h, the tensile strength increased to about
Kα2 stripping, and data analysis were performed using the MDI Jade 6.0 4385 MPa. With the increase of annealing time (up to 12 h), the tensile
software. The microstructure of the wires was characterized using FEI strength of the wires remained unchanged. For wires prepared from
Tecani G2 T20 transmission electron microscope (TEM) at an higher drawing strain (ε = 5.1 and 5.3), the tensile strength of the as-

Fig. 1. Mechanical properties of cold drawn pearlitic steel wires (ε = 4.5, 5.1 and 5.3) subjected to annealing treatment. a) Tensile strength (60 °C); b) Torsional
cycles (60 °C); c) Tensile strength (100 °C); d) Torsional cycles (100 °C), as a function of annealing time.

109
D. Wei, et al. Materials Characterization 153 (2019) 108–114

Fig. 2. TEM micrographs and SAED patterns of cold drawn pearlitic steel wires (ε = 4.5) before and after annealing treatments. a) As drawn; b) annealed at 60 °C for
12 h; c) annealed at 100 °C for 12 h; d) dark field image of c).

drawn wires is about 4780 MPa and 4920 MPa, respectively. After being
annealed at 60 °C for about 2 h, the tensile strength of the annealed
wires (ε = 5.1 and 5.3) increased to about 4900 MPa and 5020 MPa.
With the increase of annealing time, the tensile strength of the steel
wires (ε = 5.3) slightly decreased. Fig. 1b shows the torsional proper-
ties of pearlitic steel wires (ε = 4.5, 5.1 and 5.3) annealed at 60 °C for
different durations. Torsional cycle-to-failure of cold drawn pearlitic
steel wires (ε = 4.5) is about 49 cycles. After being annealed at 60 °C for
1 h, torsional cycle-to-failure decreased to about 42 cycles. For wires
prepared with higher drawing strains (ε = 5.1 and 5.3), the torsional
cycles also decreased slightly after being annealed at 60 °C. Fig. 1c
shows the tensile strength of cold drawn pearlitic steel wires after an-
nealed at 100 °C for different durations. After being annealed at 100 °C
for 0.5 h, the tensile strength of wires (ε = 4.5) increased from
4243 MPa to 4385 MPa. The strength increase is about the same as the
wires annealed at 60 °C, though the annealing time taken to reach the
highest tensile strength decreases. For wires fabricated with higher
drawing strains, the maximum tensile strength of wires (ε = 5.1 and
Fig. 3. X-ray diffraction patterns of cold drawn pearlitic steel wires before and 5.3) subjected to annealing is about 4880 MPa and 4973 MPa, respec-
after annealing treatment. As-patented and as-drawn refer to the wires of ε = 0 tively. These values are slightly lower than those obtained at 60 °C.
and ε = 4.5, respectively.
Furthermore, the tensile strength of wires (ε = 5.1 and 5.3) annealed at
100 °C for about 12 h decreased to about 4769 MPa and 4862 MPa.
Fig. 1d shows the torsional properties of pearlitic steel wires (ε = 4.5,

110
D. Wei, et al. Materials Characterization 153 (2019) 108–114

5.1 and 5.3) annealed at 100 °C for different durations. The torsional
cycles of all three cold drawn pearlitic steel wires are slightly reduced.

3.2. TEM observation

Fig. 2 shows the microstructure of heavily drawn pearlitic steel


wires (ε = 4.5) with different durations of annealing. Fig. 2a presented
the microstructure of wires preserved in as drawn state. The interface
between ferrite and cementite in heavily cold drawn pearlitic steel
wires cannot be clearly defined. High density dislocation zones can
however be observed in ferrite lamellae. After being annealed at 60 °C
for 12 h, the microstructure in annealed wires was kept unchanged, as
showed in Fig. 2b. Fig. 2c and d show bright field (BF) and dark field
(DF) microstructures of wires annealed at 100 °C for 12 h, respectively.
The selected site in the DF image was subsequently used for the dif-
fraction analysis of cementite in Fig. 2d, which indicates the presence of
nano-precipitates. It is worth noting that the selected aperture also
covers some part of ferrite in the DF image, from which the diffraction
spots of distorted ferrite lattice are observed.

Fig. 4. Carbon atom maps defined by isoconcentration surfaces (with a voxel 3.3. XRD analysis
size of 1.0 nm and a delocalization distance of 2.0 nm) of 1, 8 and 17 at.% C in
cold drawn pearlitic steel wires (ε = 4.5). Carbon atoms and isoconcentration Fig. 3 shows X-ray diffraction patterns of cold drawn pearlitic steel
surfaces are shown in red and peacock blue, respectively. The reconstructed
wires (ε = 4.5) after different durations of annealing treatments. Two
volume is 63 × 63 × 244 nm3 containing 843,000 carbon atoms. (For inter-
diffraction peaks, corresponding to α-Fe (110) and (200), can be
pretation of the references to colour in this figure legend, the reader is referred
to the web version of this article.)
identified. The diffraction peaks of (110) α-Fe and (200) α-Fe show sig-
nificant broadening, which are understood to result from nano-struc-
turing and residual internal stress. It indicates that a refinement of
ferrite lamellae and an increase of dislocation density took place during
the cold drawing process [29,30]. The (110) α-Fe and (200) α-Fe dif-
fraction peaks of the wires annealed at 60 °C and 100 °C remain un-
changed. The (211) cementite and (221) cementite diffraction peaks can be
observed in as-patented pearlitic steel rods. After severe cold drawing
(ε = 4.5), the (211) cementite and (221) cementite diffraction peaks dis-
appeared, indicative of the decomposition of the cementite platelets. No
diffraction peaks of (211) cementite and (221) cementite can be observed in
the annealed wires, suggesting that the low temperature annealing
treatment (i.e., annealed at 60 °C and 100 °C) has little effect on the
microstructure of cold drawn wires.

3.4. APT analysis

3.4.1. Carbon isoconcentration surfaces


Fig. 4 shows the spatial distribution of carbon atoms in cold drawn
pearlitic steel wires (ε = 4.5). The isoconcentration surfaces present a
range of carbon concentration levels. As carbon concentration reaches
17 at.%, the isoconcentration surfaces of carbon atoms almost dis-
appear, as shown in Fig. 4.
Fig. 5 shows the spatial distribution of carbon atoms in cold drawn
pearlitic steel wires (ε = 4.5) annealed at 100 °C for 12 h. Carbon atoms
are found to regroup and form layers of carbon-enriched regions par-
allel to each other. It suggests that carbon atom segregation may take
place along the original cementite positions [15]. Small isosurfaces (as
Fig. 5. Carbon atom maps defined by isoconcentration surfaces (with a voxel shown by the blue arrow) with a maximum carbon concentration of
size of 1.0 nm and a delocalization distance of 2.0 nm) of 1, 8 and 17 at.% C for about 17 at.% appear as lamellar carbon enrichment zones in Fig. 5. It
cold drawn pearlitic steel wires (ε = 4.5) annealed at 100 °C for 12 h. Carbon indicates that the nano-precipitates may form. Therefore, the low an-
atoms and isoconcentration surfaces are shown in red and peacock blue, re- nealing treatment can cause a significant level of carbon segregation
spectively. The reconstructed volume is 62 × 62 × 336 nm3 containing and precipitation of intermediate carbides.
787,000 carbon atoms. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)
3.4.2. 3D atomic maps of components and 1D concentration profiles
Fig. 6a shows 3D atomic map of the elements (C, Cr, Mn and Si) in
cold drawn pearlitic steel wires (ε = 4.5) before and after the annealing
treatment. Cr, Mn and Si exhibit a relatively uniform distribution as a
result of mechanical alloying. A region of interest (ROI) (marked by the
pink box) is selected from the 3D atomic map for an in-depth in-
vestigation. Fig. 6b shows the corresponding 1D concentration profiles

111
D. Wei, et al. Materials Characterization 153 (2019) 108–114

Fig. 6. (a, c) A 7 at.% carbon (red) isoconcentration surface (peacock blue), solute atom (Cr (blue), Mn (light yellow), Si (purple)) maps and (b, d) concentration
profiles of pearlitic steel wires (ε = 4.5) (a and b for the as-drawn; c and d for the annealed at 100 °C for 12 h), together with selected region of interest (ROI). Size of
the analysis box: (a) 63 × 63 × 244 nm3 and (c) 62 × 62 × 336 nm3. Size of the ROI box of cold drawn steel wires before and after the annealing treatment:
37 × 17 × 26 nm3 and 12 × 30 × 22 nm3. The red arrow represents the direction of concentration analysis.

for C, Cr, Mn and Si within the ROI. The 1D concentration curves of Cr, reduction of free energy as carbon atoms were segregated toward SGBs
Mn, and Si show very limited fluctuation. The highest carbon con- [17,33], forming layer-like and parallel carbon-rich zones. Carbide was
centration is about 12 at.% in the carbon-rich region. Red arrows point also found to nucleate at 100 °C in Fig. 5 and partially depletes the
to the location of analysis in Fig. 6a. Within 10–20 nm, carbon con- carbon level in the matrix.
centration shows a large fluctuation especially moving from carbon-rich Notably, the maximal carbon concentration extracted from con-
toward carbon-poor regions. It shows that there is a large variation of centration profiles across the lamellae is below Fe3C stoichiometry
carbon content (1–4 at.%) in ferrite in as-drawn wires. A carbon con- (25 at.%). Since carbon-poor and carbon-rich regions have very dif-
centration-strain relation has also been reported in earlier work ferent evaporation fields, a local magnification effect might be expected
[31,32]. [19,34]. This effect would reduce the integrity of 3D reconstruction.
Fig. 6c and d show 3D atomic map and corresponding 1D con- Specifically, the carbon distribution in carbon-rich regions could be
centration curves of the elements in steel wires (ε = 4.5) annealed at smeared out. Consequently, carbon concentrations lower than the ac-
100 °C for 12 h. There is no apparent redistribution of Cr, Mn and Si tual values might be derived from the carbon enrichment zones. As
atoms. However, the heat treatment at 100 °C caused a significant re- such, the presence of intermediate carbides can be identified through
distribution of carbon atoms. The maximal carbon concentration in- APT isoconcentration surfaces, while the type of carbides (such as ε-
creases to about 14 at.% in the lamellar carbon-rich regions. The red carbide and η-carbide) cannot be readily determined.
arrows point to the direction of concentration analysis in Fig. 6c, where
there is a less fluctuation in carbon-poor regions (carbon concentration 3.5. DSC analysis
of 1–2 at.%). After being annealed at 100 °C, the carbon concentration
in carbon-poor regions was slightly decreased. It might be driven by the Fig. 7 shows the DSC curves of cold drawn pearlitic steel wires

112
D. Wei, et al. Materials Characterization 153 (2019) 108–114

longer annealing time or higher temperatures may promote the trans-


formation of these small coherent intermediate carbides into larger and
incoherent precipitates [19], leading to the decrease of the tensile
strength in the steel wires.
The ductility of cold drawn pearlitic steel wires can be evaluated by
torsional tests [36,37]. A steel wire having a good ductility can be
twisted homogeneously over its whole length for a large number of
torsional cycles before failure. The torsional cycles of the annealed
wires were found to decrease slightly in Fig. 1b and d. The decrease in
plasticity might result from localised strengthening and longitudinal
cracking (i.e., delamination) [19]. Delamination usually occurred along
the direction of drawing at a very early stage [38], which typically
originated from local shear stress concentration in the hardened areas
during torsion tests. It might be linked to a high local carbon con-
centration within the wires [31]. In this work, the low temperature
annealing caused a significant segregation of carbon atoms and pre-
cipitation of intermediate carbides at SGBs, as shown in Fig. 5. The
segregated carbon atoms and intermediate carbides are believed to
Fig. 7. DSC patterns of cold drawn pearlitic steel wires (ε = 4.5) before and inhibit the mobility of the dislocations and grain boundaries migration
after annealing. [16], consequently reducing the ductility of the treated wires.

(ε = 4.5) before and after the annealing treatment. Two broad exo- 4. Conclusions
thermic peaks appear over a temperature range from about 30 °C to
465 °C. Microstructural characteristics and mechanical properties of heavily
For all types of steel wires, a very weak broad exothermic peak cold drawn pearlitic steel wires are investigated before and after low
appears over a temperature range from about 115 °C to 225 °C. This temperature annealing. The following conclusions can be drawn:
exothermic peak may be related to carbon atoms dissolving into ferrite
lattice defects [7,13]. In addition, the exothermic peak appears to shift 1) The tensile strength of cold drawn pearlitic steel wires increased
toward higher temperature for the annealed wires. following the low temperature annealing treatment. For the wires
A relatively broad exothermic peak is evident from about 330 °C to (ε = 4.5) annealed at 60 °C for 1 h, the tensile strength increased
450 °C for all types of steel wires. This peak might correspond to ce- from about 4243 MPa to 4385 MPa. For the cold drawn wires
mentite re-precipitation [19] and spheroidization of lamellar cementite (ε = 5.3) annealed at 60 °C for 1 h, the tensile strength rose from
in cold drawn wires [9,18]. By comparison, the exothermic peak of the about 4920 MPa to 5020 MPa. However, the ductility decreased
as-drawn wires spans from about 333 °C to 451 °C, while the exothermic slightly based on the torsion tests.
peaks of the annealed wires shift toward higher temperature, i.e., from 2) During the low temperature annealing treatment, carbon atoms
about 335 °C to 465 °C. The stored energy is understood to serve as the were found to segregate toward SGBs and particle precipitation was
driving force for spheroidization of cementite lamellae. During the also observed in pearlitic steel wires. Substitutional solid solution
annealing treatment, part of stored energy in wires was released, re- atoms, such as Cr, Mn and Si, remained at the original positions.
sulting in higher exothermic peak in the DSC curves. 3) Exothermic peaks in DSC curves shifted toward higher temperature
During severe cold drawing process, cementite was deformed and for the annealed wires.
even decomposed [21,34]. Deformation-driven cementite dissolution
increased carbon concentration in ferrite, leading to the increase of the Data availability
steel wire strength [5,17,19]. However, by combining the APT results
and mechanical properties in this work, it reveals that the tensile All data included in this study are available upon request by contact
strength of the annealed wires had limited increase (the highest is about with the corresponding author.
140 MPa), since carbon atoms diffused from oversaturated sites within
ferrite to subgrain boundaries (SGBs). This result raises an intriguing Disclosure statement
question: what is the primary factor responsible for the increase of
strength in the cold drawn pearlitic steel wires subjected to low tem- No potential conflict of interest was reported by the author.
perature annealing?
During such a low temperature annealing treatment (under 100 °C), Acknowledgement
dislocation climbing was difficult to occur [15]. As a result, the size of
subgrain structure, formed from the cold drawing, is expected to retain. This work is supported by the Natural Science Foundation of China
Moreover, it is impossible to alter the residual stress under such con- (grant no. 51371050), the Science and Technology Advancement
dition [12], let alone the microstructure and lamellar spacing of pear- Program of Jiangsu Province (BA2017112) and the 333 projects of
lite. However, carbon atoms in over-saturated ferrite may diffuse into Jiangsu Province, China (BRA2018045). The study was also partly
SGBs and form carbon-rich regions during low temperature annealing supported by Industry-University Research Cooperation Project of
process. As the carbon atoms segregated into the SGBs, the interfacial Jiangsu Province, China (BY2018194). Authors thank X. Shen, X. Y. Jia,
energy would be reduced, so that the subgrain boundary can be stabi- B. B. Shen, Y. Huang for assistance with preparation the samples. Z Xie
lized and serve as barriers against dislocation movement [17]. In ad- thanks for the support provided by the Australian Research Council
dition, the increased strength may also attribute to the precipitation of Discovery Projects.
intermediate carbides. These precipitates are usually coherent or semi-
coherent and very small in size (i.e., few nanometers). Such inter- References
mediate carbides could limit interfacial dislocation sources and create
additional hardening effect for the steel wires [19,35]. However, the [1] D. Raabe, P.P. Choi, Y.J. Li, A. Kostka, X. Sauvage, F. Lecouturier, K. Hono,
R. Kirchheim, R. Pippan, D. Embury, Metallic composites processed via extreme

113
D. Wei, et al. Materials Characterization 153 (2019) 108–114

deformation: toward the limits of strength in bulk materials, MRS Bull. 35 (12) Scr. Mater. 120 (2016) 5–8, https://doi.org/10.1016/j.scriptamat.2016.04.002.
(2010) 982–991, https://doi.org/10.1557/mrs2010.703. [21] F. Fang, Y.F. Zhao, P.P. Liu, L.C. Zhou, X.J. Hu, X.F. Zhou, Z.H. Xie, Deformation of
[2] J.D. Embury, R.M. Fisher, The structure and properties of drawn pearlite, Acta cementite in cold drawn pearlitic steel wire, Mater. Sci. Eng. A 608 (2014) 11–15,
Metall. 14 (1966) 147–159, https://doi.org/10.1016/0001-6160(66)90296-3. https://doi.org/10.1016/j.msea.2014.04.050.
[3] G. Langford, A study of the deformation of patented steel wire, Metall. Mater. Trans. [22] C. Borchers, T. Al-Kassab, S. Goto, R. Kirchheim, Partially amorphous nano-
B Process Metall. Mater. Process. Sci. 1 (1970) 465–477, https://doi.org/10.1007/ composite obtained from heavily deformed pearlitic steel, Mater. Sci. Eng. A 502
BF02811557. (2009) 131–138, https://doi.org/10.1016/j.msea.2008.10.018.
[4] M.H. Hong, W.T. Reynolds, T. Tarui, K. Hono, Atom probe and transmission elec- [23] V.G. Gavriljuk, Comment on “effect of interlamellar spacing on cementite dissolu-
tron microscopy investigations of heavily drawn pearlitic steel wire, Metall. Mater. tion during wire drawing of pearlitic steel wires”, Scr. Mater. 45 (2001) 1469–1472,
Trans. A 30 (1999) 717–727, https://doi.org/10.1007/s11661-999-1003-y. https://doi.org/10.1016/S1359-6462(01)01185-X.
[5] Y.J. Li, P. Choi, C. Borchers, S. Westerkamp, S. Goto, D. Raabe, R. Kirchheim, [24] M. Herbig, P. Choi, D. Raabe, Combining structural and chemical information at the
Atomic-scale mechanisms of deformation-induced cementite decomposition in nanometer scale by correlative transmission electron microscopy and atom probe
pearlite, Acta Mater. 59 (10) (2011) 3965–3977, https://doi.org/10.1016/j. tomography, Ultramicroscopy 153 (2015) 32–39, https://doi.org/10.1016/j.
actamat.2011.03.022. ultramic.2015.02.003.
[6] X.D. Zhang, A. Godfrey, N. Hansen, X.X. Huang, Hierarchical structures in cold- [25] K. Hono, M. Ohnuma, M. Murayama, S. Nishida, A. Yoshie, T. Takahashi, Cementite
drawn pearlitic steel wire, Acta Mater. 61 (13) (2013) 4898–4909, https://doi.org/ decomposition in heavily drawn pearlite steel wire, Scr. Mater. 44 (6) (2001)
10.1016/j.actamat.2013.04.057. 977–983, https://doi.org/10.1016/S1359-6462(00)00690-4.
[7] P. Watté, J.V. Humbeeck, E. Aernoudt, I. Lefever, Strain aging in heavily drawn [26] Q. Liu, S. Zhao, Cu precipitation on dislocation and interface in quench-aged steel,
eutectoid steel wires, Scr. Mater. 34 (1) (1996) 89–95, https://doi.org/10.1016/ MRS Commun. 2 (2012) 127–132, https://doi.org/10.1557/mrc.2012.21.
1359-6462(95)00479-3. [27] W. Sha, L. Chang, G.D.W. Smith, C. Liu, E.J. Mittemeijer, Some aspects of atom-
[8] J. Languillaume, G. Kapelski, B. Baudelet, Cementite dissolution in heavily cold probe analysis of Fe-C and Fe-N systems, Surf. Sci. 266 (1992) 416–423, https://doi.
drawn pearlitic steel wires, Acta Mater. 45 (1997) 1201–1212, https://doi.org/10. org/10.1016/0039-6028(92)91055-G.
1016/S1359-6454(96)00216-9. [28] C. Zhu, A. Cerezo, G.D.W. Smith, Carbide characterization in low-temperature
[9] J. Languillaume, G. Kapelski, B. Baudelet, Evolution of the tensile strength in tempered steels, Ultramicroscopy 109 (2009) 545–552, https://doi.org/10.1016/j.
heavily cold drawn and annealed pearlitic steel wires, Mater. Lett. 33 (1997) ultramic.2008.12.007.
241–245, https://doi.org/10.1016/S0167-577X(97)00109-2. [29] Y.Z. Chen, G. Csiszár, J. Cizek, S. Westerkamp, C. Borchers, T. Ungár, S. Goto,
[10] V.T.L. Buono, B.M. Gonzalez, M.S. Andrade, Kinetics of strain aging in drawn F. Liu, R. Kirchheim, Defects in carbon-rich ferrite of cold-drawn pearlitic steel
pearlitic steels, Metall. Mater. Trans. A 29 (5) (1998) 1415–1423, https://doi.org/ wires, Metall. Mater. Trans. A 44 (2013) 3882–3889, https://doi.org/10.1007/
10.1007/s11661-998-0356-y. s11661-013-1723-x.
[11] J. Toribio, Relationship between microstructure and strength in eutectoid steels, [30] Y.Z. Chen, G. Csiszár, J. Cizek, C. Borchers, T. Ungár, S. Goto, R. Kirchheim, On the
Mater. Sci. Eng. A 387–389 (2004) 227–230, https://doi.org/10.1016/j.msea.2004. formation of vacancies in α-ferrite of a heavily cold-drawn pearlitic steel wire, Scr.
01.084. Mater. 64 (2011) 390–393, https://doi.org/10.1016/j.scriptamat.2010.10.039.
[12] T. Suzuki, Y. Tomota, M. Isaka, A. Moriai, N. Minakawa, Y. Morii, Strength aniso- [31] N. Maruyama, T. Tarui, H. Tashiro, Atom probe study on the ductility of drawn
tropy and residual stress in drawn pearlite steel wire, Trans. Iron Steel Inst. Jpn. 44 pearlitic steels, Scr. Mater. 46 (2002) 599–603, https://doi.org/10.1016/S1359-
(2004) 1426–1430, https://doi.org/10.2355/isijinternational.44.1426. 6462(02)00037-4.
[13] D.B. Park, J.W. Lee, Y.S. Lee, K.T. Park, W.J. Nam, Effects of the annealing tem- [32] Y.J. Li, P. Choi, C. Borchers, Y.Z. Chen, S. Goto, D. Raabe, R. Kirchheim, Atom probe
perature and time on the microstructural evolution and corresponding the me- tomography characterization of heavily cold drawn pearlitic steel wire,
chanical properties of cold-drawn steel wires, Met. Mater. Int. 14 (1) (2008) 59–64, Ultramicroscopy 111 (2011) 628–632, https://doi.org/10.1016/j.ultramic.2010.
https://doi.org/10.3365/met.mat.2008.02.059. 11.010.
[14] X.D. Zhang, A. Godfrey, X.X. Huang, N. Hansen, Q. Liu, Microstructure and [33] M. Herbig, D. Raabe, Y.J. Li, P. Choi, S. Zaefferer, S. Goto, Atomic-scale quantifi-
strengthening mechanisms in cold-drawn pearlitic steel wire, Acta Mater. 59 (9) cation of grain boundary segregation in nanocrystalline material, Phys. Rev. Lett.
(2011) 3422–3430, https://doi.org/10.1016/j.actamat.2011.02.017. 112 (2014) 126103, , https://doi.org/10.1103/PhysRevLett.112.126103.
[15] J. Takahashi, M. Kosaka, K. Kawakami, T. Tarui, Change in carbon state by low- [34] A. Lamontagne, V. Massardier, X. Kléber, X. Sauvage, D. Mari, Comparative study
temperature aging in heavily drawn pearlitic steel wires, Acta Mater. 60 (1) (2012) and quantification of cementite decomposition in heavily drawn pearlitic steel
387–395, https://doi.org/10.1016/j.actamat.2011.09.014. wires, Mater. Sci. Eng. A 644 (2015) 105–113, https://doi.org/10.1016/j.msea.
[16] Y.J. Li, P. Choi, S. Goto, C. Borchers, D. Raabe, R. Kirchheim, Evolution of strength 2015.07.048.
and microstructure during annealing of heavily cold-drawn 6.3 GPa hypereutectoid [35] V. Massardier, N. Lavaire, M. Soler, J. Merlin, Comparison of the evaluation of the
pearlitic steel wire, Acta Mater. 60 (9) (2012) 4005–4016, https://doi.org/10. carbon content in solid solution in extra-mild steels by thermoelectric power and by
1016/j.actamat.2012.03.006. internal friction, Scr. Mater. 50 (2004) 1435–1439, https://doi.org/10.1016/j.
[17] Y.J. Li, D. Raabe, M. Herbig, P.P. Choi, S. Goto, A. Kostka, H. Yarita, C. Borchers, scriptamat.2004.03.010.
R. Kirchheim, Segregation stabilizes nanocrystalline bulk steel with near theoretical [36] L.C. Zhou, F. Fang, L.F. Wang, H.Q. Chen, Z.H. Xie, J.Q. Jiang, Torsion delamina-
strength, Phys. Rev. Lett. 113 (10) (2014) 106104, , https://doi.org/10.1103/ tion and recrystallized cementite of heavy drawing pearlitic wires after low tem-
PhysRevLett.113.106104. perature annealing, Mater. Sci. Eng. A 713 (2018) 52–60, https://doi.org/10.1016/
[18] F. Fang, L.C. Zhou, X.J. Hu, X.F. Zhou, Y.Y. Tu, Z.H. Xie, J.Q. Jiang, Microstructure j.msea.2017.12.055.
and mechanical properties of cold-drawn pearlitic wires affect by inherited texture, [37] L.C. Zhou, F. Fang, L.P. Wang, X.J. Hu, Z.H. Xie, J.Q. Jiang, Torsion performance of
Mater. Des. 79 (2015) 60–67, https://doi.org/10.1016/j.matdes.2015.04.036. pearlitic steel wires: effects of morphology and crystallinity of cementite, Mater.
[19] A. Lamontagne, V. Massardier, X. Sauvage, X. Kléber, D. Mari, Evolution of carbon Sci. Eng. A 743 (2019) 425–435, https://doi.org/10.1016/j.msea.2018.11.113.
distribution and mechanical properties during the static strain ageing of heavily [38] T. Tarui, N. Maruyama, J. Takahashi, S. Nishida, H. Tashiro, Microstructure control
drawn pearlitic steel wires, Mater. Sci. Eng. A 667 (2016) 115–124, https://doi.org/ and strengthening of high carbon steel wires, SHINNITTETSU GIHO (2004) 51–56
10.1016/j.msea.2016.04.091. https://www.nipponsteel.com/en/tech/report/nsc/pdf/n9112.pdf.
[20] L.C. Zhou, F. Fang, X.F. Zhou, Y.Y. Tu, Z.H. Xie, J.Q. Jiang, Cementite nano-crys-
tallization in cold drawn pearlitic wires instigated by low temperature annealing,

114

You might also like