Torsion Performance of Pearlitic Steel Wires Effects of Morphology and

You might also like

You are on page 1of 11

Materials Science & Engineering A 743 (2019) 425–435

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Torsion performance of pearlitic steel wires: Effects of morphology and T


crystallinity of cementite

Lichu Zhoua, Feng Fanga, , Liping Wanga, Xianjun Hub, Zonghan Xiec,d, Jianqing Jianga
a
Jiangsu Key Laboratory of Advanced Metallic Materials, Southeast University, Nanjing 211189, China
b
Jiangsu Sha-Steel Group, Zhangjiagang 215625, China
c
School of Mechanical Engineering, University of Adelaide, SA 5005, Australia
d
School of Engineering, Edith Cowan University, WA 6027, Australia

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

Keywords: The torsion behavior of pearlitic steel wires is investigated and the conditions responsible for torsion delami-
Pearlitic steel wires nation are determined. Experimental results reveal that the morphology and crystallinity of cementite regulate
Cold drawing the torsion response of pearlitic steel wires. Specifically, both the torsion strength and torsion fracture strain of
Cementite the wires increase with cold drawing strain up to ε < 0.68. After annealing treatment at 250 °C, the torsion
Torsion delamination
strength increases at the expense of torsion fracture strain irrespective of drawing strains. Moreover, the torsion
Annealing
delamination occurs in the wires prepared at ε > 1.76 and also subjected to annealing treatment between
150 °C and 350 °C. It was found that annealing treatment turns the amorphous cementite to nanocrystalline state,
which prevents dislocations from bowing out. Consequently, localized and severe plastic deformation takes
place, resulting in the torsion delamination. By contrast, the cementite with a mix of amorphous and nano-
crystalline phases is capable of withstanding plastic deformation during torsion tests. Moreover, annealing-
induced spheroidization of cementite is effective in preventing the pearlitic steel wires from torsion delamina-
tion.

1. Introduction torsion tolerance was independent of the tensile strength [30,31,36].


Recently, the residual stress release [37] and diffusion of carbon atoms
Pearlitic steel wires have been widely used in a wide range of en- [15,31,34] have been taken into consideration. Notably, torsion dela-
gineering applications such as suspension bridge cables, automotive mination could take place in pearlitic steel wires prepared under rela-
tyre cords and cutting wires [1–4]. They are produced by cold drawing tively low drawing strains, i.e., without apparent cementite dissolution
process, followed by post-drawing annealing treatment to obtain a good [38,39]. Suzuki [19] also reported that the residual stress in pearlitic
combination of strength and toughness. Mechanical properties of steel wires was insensitive to low temperature annealing. Therefore, the
pearlitic steel wires are regulated by refined lamellar structure [5–9], underlying mechanism that governs the torsion delamination of pear-
heavily deformed cementite [10–15], increased dislocation density in litic steel wires remains unclear.
ferrite [16–18] and residual stress [19,20]. Multiple studies have suggested that torsion delamination might be
High tensile strength enables pearlitic steel wires to withstand se- induced by post-drawing annealing treatment [30,31,34,37]. Amor-
vere tensile loading [1]. However, in practical applications, the wires phous cementite developed in heavily drawn pearlitic steel wires can be
should maintain a certain level of shear resistance [5,7,17,21–23]. transformed into nano-crystallites during the annealing process
Torsion tests are thus used to assess the shear strength and shear frac- [30,40–42]. The occurrence of torsion delamination was observed to be
ture strain of pearlitic steel wires [24–27]. The shear fracture strain of closely related with the presence of nanocrystalline cementite [30].
the wires would decrease substantially under the influence of torsion Cementite in the annealed wires was found to dissolve during torsion
delamination. As such, the torsion delamination has attracted attention tests, and severe local plastic deformation led to the torsion fracture
for many years [28–36]. Initially, torsion delamination was considered [28]. However, to date, the role of cementite substructure has received
to originate from low-temperature annealing or localized work hard- little attention when analyzing the mechanical properties of the wires.
ening [29,33,35]. However, subsequent studies indicated that the The general view is that the influence of cementite upon mechanical


Correspondence to: 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.msea.2018.11.113
Received 6 September 2018; Received in revised form 20 November 2018; Accepted 22 November 2018
Available online 23 November 2018
0921-5093/ © 2018 Elsevier B.V. All rights reserved.
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Table 1 rods were cooled down rapidly to room temperature by water cooling.
The chemical composition of steel rods used in this study (wt%). After pickling and phosphating [45], the rods were cold drawn to
Element C Si Mn Cr S+P Fe 2.0 mm in diameter (ε = 2.06) by multi-pass drawing process. The
temperature during drawing was controlled and kept under 100 °C by
wt% 0.84 0.3 0.20 0.18 < 0.015 Bal. liquid cooling. The post-drawing annealing treatments of the wires
were carried out in oil (150 °C and 250 °C) and salt bath (350 °C, 450 °C
and 550 °C), respectively. The holding time was 60 min for all the an-
properties mainly stems from its distribution and morphology [43,44]. nealing treatments before the wires were cooled in air to room tem-
Besides, there is a lack of systematic studies to unlock the intrinsic re- perature.
lationship between the microstructure and torsion performance of the The length of the test wires was 50 times the diameter. Torsion
pearlitic steel wires. Unpredictable torsion delamination remains to be experiments were conducted using a torsional test machine (Type:
a hidden danger to the mechanical performance and durability of CTT500, Shenzhen Suns Technology Stock Co. Ltd, China). Both the
pearlitic steel wires in actual practice. torsion stress and strain were recorded in real time. The rotational
In present study, the torsion behavior of pearlitic steel wires pre- speed during the tests was set to be 6°/s, which is equivalent to a torsion
pared at different drawing strains and also subjected to annealing strain rate of 6.7 × 10−2 s−1. Five samples were tested under each
treatments were investigated. Effects of cementite's morphology and condition to ensure the reliability and accuracy of test results.
crystallinity upon the torsion performance were examined. To ascertain For microstructural observation, after polishing along the long-
the origin of torsion delamination, key factors such as the direction and itudinal direction, the samples were etched by 4% nitric acid alcohol
distribution of pearlite lamellae, dislocation configuration and ce- solution. The samples were examined using FEI Sirion-400 scanning
mentite substructure were analyzed. Moreover, the conditions that electron microscope (SEM). The orientation distribution of pearlite la-
govern the torsion delamination were discussed. mellae was quantified by analyzing the lamellar direction using ten
SEM micrographs for each type of the wire. TEM samples were prepared
by GATAN 691 Precision Ion Polishing System (PIPS) after the samples
2. Materials and methods were mechanically thinned to a thickness less than 40 µm. The samples
were cooled by liquid nitrogen in PIPS to avoid undesirable tempera-
Starting materials used in this study are steel rods with a diameter of ture rise. At the final stage of TEM sample preparation the ion angle and
~ 5.6 mm. Their chemical composition is shown in Table 1. The steel energy was set to 3° and 2 Kev respectively to minimize microstructure
rods were re-heated in order to obtain fine pearlite microstructure with damage. The microstructure details of the wires were examined by FEI
uniform interlamellar spacing. The heat treatment process can be de- T20-G2 and JEOL 2100F transmission electron microscopes.
scribed as: Firstly, the rods were heated to 920 °C and held for 15 min,
then quenched in liquid salt at 550 °C and kept for 3 min. Finally, the

Fig. 1. SEM micrographs of cold drawn pearlitic steel wires fabricated from different drawing strains. Drawing direction (D. D.) was marked in d) a) ε = 0, b)
ε = 0.68, c) ε = 1.39, d) ε = 2.06.

426
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Fig. 2. SEM micrographs of the pearlitic steel wires prepared from ε = 2.06 and then subjected to annealing at different temperatures for 60 min: a) 150 °C, b) 350 °C,
c) 450 °C, d) 550 °C.

3. Results and cementite lamellae (the thinner ones). As presented in DF-TEM


(Fig. 3b), a small amount of particles with weak contrast can be ob-
3.1. Microstructure characteristics served in the cementite plates. Fig. 3c is the BF-TEM micrograph of the
wires after annealing treatment at 250 °C for 60 min. The wires retain
Fig. 1 shows the microstructure obtained from the longitudinal its lamellar structure after the annealing treatment. As presented in the
section of pearlitic steel wires fabricated with different drawing strains. DF-TEM micrograph(Fig. 3d), a large amount of fine particles appear in
As shown in Fig. 1a, the pearlitic steel rods (i.e., prior to cold drawing) the cementite plate. Fig. 3e is the high-resolution TEM (HRTEM) mi-
exhibited a lamellar structure with random lamellar orientation. Pear- crograph of the as-drawn wires. Ferrite region is separated by 4–6 nm
lite colonies in the wires corresponding to ε = 0.68 were stretched and wide cementite lamellae. The cememtite lamellae contain a mix of
aligned toward the drawing direction (Fig. 1b). The microstructure of amorphous and nanocrystalline phases with extremely small grain size.
the wires prepared with ε = 1.39 is presented in Fig. 1c. The angle It indicates that part of cementite was transformed into amorphous
between the majority of pearlite lamellae and the drawing direction is structure under heavy cold drawing, which is in agreement with pre-
smaller than 45°. After heavy drawing, the resultant microstructure vious reports [3,11,12]. HRTEM micrograph of the pearlitic steel wires
exhibited an axially-aligned lamellar structure. Almost all of pearlite after annealing treatment at 250 °C for 60 min is presented in Fig. 3f.
lamellae in the wires prepared from ε = 2.06 were parallel to the Clear lattice fringes of cementite can be observed between ferrite,
drawing direction (Fig. 1d). Moreover, the interlamellar spacing in the suggesting that the cementite lamellae transformed to primarily nano-
wires was further decreased. crystalline structure during the annealing treatment.
Microstructures of pearlitic steel wires subjected to annealing TEM micrograph of pearlitic steel wires (ε = 2.06) after annealing
treatment are presented in Fig. 2. The pearlitic steel wires were pre- treatment at 550 °C for 60 min is presented in Fig. 4. The annealing
pared from ε = 2.06 before the annealing. Lamellar structure in the treatment resulted in remarkable spheroidization of cementite. Mean-
wires remains unchanged under annealing at 150 °C and 350 °C. As while, heavily deformed ferrite lamellae recrystallized and coalesced
presented in Fig. 2a and b, the two wires still comprise an axially into coarser ferrite grains. Note that the crystal-orientation is random in
aligned lamellar structure. A small amount of disconnected cementite the TEM specimen, which shows that the recovery has occurred during
plates are observed in the 350 °C-annealed wires. During annealing at the annealing, and the dislocation density has decreased considerably
450 °C, the cementite plates turned into round particles or short rods [16,41].
(Fig. 2c). As shown in Fig. 2d, during annealing treatment at 550 °C
cementite in the wires spheroidized into larger particles.
Fig. 3 shows the TEM micrographs of pearlitic steel wires fabricated 3.2. Torsion properties
under ε = 2.06. Fig. 3a is the BF-TEM micrograph of the as-drawn
wires. The lamellar structure in the wires consisits of alternating ferrite The torsion properties of the pearlitic steel wires are shown in
Fig. 5. The torsion strength of both as-drawn wires and 250 °C-annealed

427
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Fig. 3. TEM and HRTEM micrographs of pearlitic steel wires prepared with the drawing strain of ε = 2.06 a) Bright field (BF) TEM micrograph of as-drawn wires, b)
Dark field (DF) micrograph of a), c) BF-TEM micrograph of pearlitic steel wires after annealing treatment at 250 °C for 60 min, d) DF TEM micrograph of c), e)
HRTEM micrograph of as-drawn wires, f) HRTEM micrograph of pearlitic steel wires after annealing treatment at 250 °C for 60 min.

wires are plotted in Fig. 5a. As the drawing strain increases to 2.06, the higher than the as-drawn wires. However, the torsion strength of the
torsion strength of as-drawn wires increases from about 980–1780 MPa. annealed wires prepared under ε = 1.76 and 2.06 have shown larger
The torsion strength of the 250 °C-annealed wires, for the most part, is variations, increasing from about 1500–1850 MPa. Fig. 5b shows the
slightly higher than that of as-drawn wires. For instance, the torsion torsion fracture strain of both as-drawn wires and 250 °C-annealed
strength of annealed wires prepared from ε = 1.39 is about 100 MPa wires. The torsion fracture strain of the as-drawn wires firstly increases

428
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Fig. 6. Torsion properties of pearlitic steel wires (ε = 2.06) subjected to an-


nealing treatment at different temperatures.

the torsion stress-strain curves and torsion fracture surfaces of the wires
Fig. 4. TEM micrograph of pearlitic steel wires fabricated under ε = 2.06 and
then subjected to annealing treatment at 550 °C for 60 min. are presented in Fig. 7. The torsion stress-strain curves of the pearlitic
steel wires (ε = 0.68) are presented in Fig. 7a. As the torsion strain γ
increases to about 0.02, the torsion stress of the two wires increases
from 0.85 (ε = 0) to 1.6 (ε = 0.68), indicating that smaller drawing quickly to the yield point. The small difference in torsion performance
strains can simultaneously improve the torsion strength and torsion between the two wires can be seen after the yield point is reached. The
fracture strain of the pearlitic steel wires. With a further increase of cold 250 °C-annealed wires experiences a slightly higher torsion stress than
drawing strain, however, the torsion fracture strain gradually decreased that of the as-drawn wires up to the torsion fracture. Fig. 7b shows the
from about 1.6 (ε = 0.68) to 1.1 (ε = 2.06). For annealed wires, the torsion stress-strain curves of the pearlitic steel wires made from
evolution of torsion fracture strain with drawing strain was basically ε = 2.06. It can be seen that as with the as-drawn wires the torsion
consistent with the as-drawn wires. The torsion fracture strains of an- stress of the 250 °C-annealed wires increases quickly at the initial stage.
nealed wires prepared from ε = 1.76 and 2.06 exhibited a greater This is followed by a rapid decline in torsion stress at strain γ ≈ 0.025.
variation in value from about 0.4–1.0. The large variations or un- The torsion stress fluctuates with a further increase in torsion strain γ.
expected torsion fracture strains have been reported to be associated The initial decline and subsequent fluctuation of the torsion stress has
with the torsion delamination [28,31,33,35]. been reported to be caused by crack initiation and propagation
Fig. 6 shows the changes of both torsion strength and torsion frac- [28,30,31,33,35]. Finally, the delamination fracture occurred in the
ture strain of pearlitic steel wires (ε = 2.06) in response to annealing wires. The torsion stress-strain curve for the ε = 2.06 wires subjected
treatment. When the annealing temperatures are 150 °C, 250 °C and annealing treatment at 450 °C is presented in Fig. 7b. The curve is
350 °C, both the torsion strength and torsion fracture strain of the wires smooth up to the fracture occurring at torsion strain γ ≈ 0.95.
exhibit large variations, for example, from 1550 MPa (ε = 0.25) to Fig. 7c shows the torsion fracture of as-drawn pearlitic steel wires
1900 MPa to (ε = 0.9). As the annealing temperature increases to (ε = 2.06), without delamination. A smooth fracture surface appears,
450 °C, the torsion properties of the pearlitic steel wires become more and there are no other cracks around the torsion fracture. The torsion
stable. Compared to the as-drawn wires, the torsion fracture strain of fracture of 250 °C annealed pearlitic steel wires (ε = 2.06) consists of
450 °C-annealed wires decreases slightly, i.e., from 1.05 to about 0.95, multiple cracks (Fig. 7d), indicative of a delamination type. The
and the torsion strength decreases to about 1550 MPa. When the an- dominant crack cut through the surface of the fractured pearlitic steel
nealing temperature further increases to 550 °C, the torsion failure wire with smaller cracks kept inside the wire. Previous studies
strain increases at the expense of the torsion strength. [28,30,31,37] have found the sharp drop in torsional stress resulted
To further revealing the difference in torsion performance between from the formation of tiny cracks in the wires. Therefore, a sudden drop
the as-drawn wires and the wires subjected to the annealing treatment, in the torsion stress-strain curve presented in this study (Fig. 7b)

Fig. 5. Torsion properties of pearlitic steel wires a) The torsion strength and b) the torsion fracture strain of the wires as a function of drawing strains.

429
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Fig. 7. Torsion strain-stress curves and torsion fracture observation of pearlitic steel wires a) Torsion strain-stress curve of pearlitic steel wires (ε = 0.68); b) Torsion
strain-stress curve of pearlitic steel wires (ε = 2.06); c) Torsion fracture of as-drawn pearlitic steel wires (ε = 2.06); d) Torsion fracture of 250 °C-annealed pearlitic
steel wires (ε = 2.06).

suggests the development of tiny cracks in the wires at small torsion dislocations has developed uniformly in ferrite lamellae. The cementite
stains. These cracks propagated during the torsion tests, and eventually plates, which comprises a mix of amorphous and nanocrystalline phases
led to the torsion delamination fracture. (Fig. 3), appear not to be affected by the torsion.
The orientation and its distribution of both the as-drawn wires and
those experiencing torsion fracture are presented in Fig. 10 as a func-
3.3. Microstructure deformation tion of cold drawing strains. With increasing drawing strain ε, the
pearlitic lamellae are aligned within 15° around the drawing direction.
The torsion-induced microstructure change in as-drawn (ε = 0.68) The orientation distribution indicates that the pearlitic lamellae have
pearlitic steel wires is presented in Fig. 8. The wires retain its lamellar turned away from the drawing direction (D. D.) after torsion. It is in-
structure after torsion fracture, however, the pearlite lamellae are teresting to note that the orientation of lamellae in the as-drawn wires
found to deviate from the drawing direction (Fig. 8a). It indicates that (ε = 0.68) exhibits a similar distribution to that of the torsion fractured
the pearlite lamellae have twisted during the torsion test. Fig. 8b shows wires as presented in Fig. 10c.
that high density of dislocation bundles have formed, which are dis- Fig. 11 is SEM and TEM micrographs of the pearlitic steel wires
persed in ferrite lamellae separated by cementite plates. (ε = 2.06) after torsion tests. Note that the wires have been annealed at
The torsion-induced microstructure deformation of as-drawn pear- 250 °C for 60 min before the torsion test, which shows that the fracture
litic steel wires (ε = 2.06) is presented in Fig. 9a. The direction of strain γ of the wires is 0.45. However, as a result of inhomogeneous
pearlitic lamellae departs from the drawing direction; however, no plastic deformation during the delamination torsion [28,37], the true
apparent fracture or fragmentation of cementite can be observed in the torsion strain γ of the tested samples may not be 0.45. As shown in the
wires after torsion fracture. Fig. 9b shows that a high density of

430
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Fig. 8. Microstructure of as-drawn pearlitic steel wires (ε = 0.68) after torsion test. a) SEM micrograph, b) TEM micrograph.

SEM micrograph (Fig. 11a), some distorted regions comprise cementite Fig. 12a is the SEM micrograph of the pearlitic steel wires (ε = 2.06)
fragments dispersed among lamellar microstructure. It indicates that obtained after the torsion tests. Note that the pearlitic steel wires were
local cementite plates have been broken into small particles by torsion annealed at 550 °C for 60 min before the torsion tests. A large number of
deformation. As seen in Fig. 11b ferrite lamellae show different contrast randomly oriented cementite particles can be observed in the SEM
between each other. Selected area electron diffraction (SAED) of micrograph. Cementite particles are almost encircled by high density
marked region “1” to “4” is presented in Fig. 11c. These four regions are dislocations in the wires as shown in TEM micrograph (Fig. 12b).
close to each other within the range of 300 nm, the corresponding SAED However, these cementite particles appear not to be influenced by the
patterns, however, are different to each other. Fig. 11d presents the DF- torsion tests.
TEM micrograph of the same area of Fig. 11b by selecting the diffrac-
tion spot marked in Fig. 11c. A small region with high-density dis- 4. Discussion
locations can be observed. These dislocations appear to be blocked by
cementite plates. Previous studies have reported that the mis-orienta- 4.1. The role of cementite in determining the torsional properties
tion of adjacent ferrite lamellae is quite small [17]. Thus, the present
finding suggests that inhomogeneous and local severe deformation of Experiment results presented above suggest that there is a close link
ferrite lamellae occurred during the torsion of the 250 °C annealed between cementite and torsion properties of the pearlitic steel wires.
wires. As a result, the orientations of ferrite lamellae is alerted. Notably, The morphology and substructure of cementite in pearlitic steel wires
the cementite plates in the 200 °C annealed wires are reported to be are affected by drawing deformation and post-drawing annealing. The
dissolved by torsion test [28]. In present study, however, the cementite influence of morphology and crystallinity of cementite upon the torsion
plates can be readily identified in the DF-TEM micrograph (Fig. 11d). properties is discussed below.
Moreover, here the cementite plates are bent or become fragmented.
Such bending and fragmentation of cementite plates is believed to
partly result from stress concentration caused by high density disloca- 4.1.1. Effect of cementite morphology
tions in ferrite. Since pearlitic steel wires have an axially-aligned lamellar structure,
the influence of cementite on the deformation behavior is supposedly

Fig. 9. Microstructure of as-drawn pearlite steel wires (ε = 2.06) after torsion test: a) SEM micrograph, b) TEM micrograph.

431
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Fig. 10. Orientation and distribution of pearlite lamellae of as-drawn pearlitic steel wires and those experiencing torsion fracture: a) ε = 0, b) ε = 0.21, c) ε = 0.68,
d) ε = 1.39, e) ε = 1.76, f) ε = 2.06.

anisotropic [19,46]. Cementite plates are understood to act as barriers nanometers. There is also evidence that part of cementite is turned into
that prevent dislocations from moving through ferrite [18,47]. The amorphous state in heavily drawn pearlitic steel wires
deformation direction along cementite plates is called the ‘soft’ or- [3,11,12,22,40,55].
ientation for plastic deformation [48]. By contrast, when the applied Nanoscale amorphous interlayer between Cu/Nb nano-grains has
stress is perpendicular to the cementite plates, the cementite plates been demonstrated to be capable of absorbing dislocations and dif-
trends to fracture and the entanglement of high density dislocations fusing the strain concentration [56]. The failure strain of the Cu/Nb
would develop in ferrite near the fractured cementite [46,48]. Fur- alloy is thus increased by more than 10 times with the presence of such
thermore, micro-voids and micro-cracks are detected near the fractured amorphous interlayer. In this study, the cementite plates containing
cementite; as a result, the ductility of the wires deteriorates [49]. amorphous phase are believed to have played an important role during
In this study, both the torsion strength and torsion fracture strain of the torsion test. As presented in Fig. 9, the cementite plates consists of a
the pearlitic steel wires are found to improve simultaneously with in- mix of amorphous and nanocrystalline phases are conducive to plastic
creasing the drawing deformation up to ε < 0.68 (Fig. 5). The increase deformation during torsion. When subjected to a torsion strain of ~ 1,
of the torsion strength can be attributed to pearlite lamellae refinement the cementite plates did not fracture or bend. By contrast, cementite
[50] and the increase of dislocation density [51]. During the torsion plates consisting primarily of nanocrystalline phase, as seen for ex-
tests, the shear stress is distributed along the circumferential direction. ample in the 250 °C-annealed pearlitic steel wires (ε = 2.06) (Fig. 11),
The analysis of pearlitic lamellae orientation (Fig. 10) has indicated would block the dislocation motion during the torsion, causing the
that, during the torsion tests, the pearlite lamellae twisted to cater to breakage of cementite plates and local severe deformation. A stress
the torsion stress. It's interesting to note that the orientation distribu- decline at a torsion strain of only 0.025 (Fig. 7) means that cracks has
tion of pearlite lamellae in the ε = 0.68 wires remains unchanged after been generated. Moreover, both the microstructure analysis and torsion
the torsion tests (Fig. 10c). This indicates that the directionality of stress-strain curves suggest that the cementite plates behave like a hard
pearlite lamellae caused by cold drawing deformation might contribute and brittle phase during the torsion. Consequently, the 250 °C annealed
to the increase of the torsion fracture strain at ε < 0.68. pearlitic wires (ε = 2.06) exhibit delamination fracture during the
After the annealing treatment at 450 °C or 550 °C, the cementite torsion test.
plates in pearlitic steel wires have transformed into short rods or The torsion strength of the pearlitic steel wires also increases as a
spherical particles (Fig. 2a and Fig. 4). Under such situation, high result of low-temperature annealing. It has been reported that the dis-
density dislocations in ferrite can bow out, as presented in Fig. 12b. locations are pinned by newly formed cementite crystals [40,57].
Thus, the torsion fracture strain of the pearlitic steel wires is improved Therefore, the low-temperature annealing strengthening as seen in the
by the annealing treatment at 450 °C and 550 °C (Fig. 6). torsion tests is believed to relate with the newly formed cementite
nanocrystalline phase.

4.1.2. Effect of cementite substructure


A key question remains; that is, why the torsion performance of 4.2. Torsion delamination of pearlitic steel wires
pearlitic steel wires is sensitive to annealing treatment at low tem-
perature, even though such treatment incurred almost no change in Torsion delamination is the direct manifestation of torsion-caused
microstructure. The influence of cementite upon the mechanical prop- deterioration in material. The pearlitic steel wires with torsion dela-
erties of steel is generally believed to be closely associated with the mination show a “random” or unpredictable nature in the torsion per-
morphology of cementite [42,52–54]. However, the cementite plates in formance and failure [28,37]. In particular, the cracks might form when
heavily drawn pearlitic steel wires have been refined to only a few the delamination-prone pearlitic steel wire experience even very small

432
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Fig. 11. Microstructure analysis of 250 °C-annealed pearlitic steel wires (ε = 2.06) after torsion tests a) SEM micrograph, b) Bright field (BF) TEM micrograph, c)
SADE of the distortion region marked in a), and d) Dark field (DF) micrograph of b).

Fig. 12. Microstructure analysis of 550 °C-annealed pearlitic steel wires (ε = 2.06) after the torsion tests: a) SEM micrograph, b) TEM micrograph.

torsion strain, which can influence the stress loading of the wires steel wires, especially in relation to safety-vital applications.
considerably (Fig. 7). Such easy crack generation and random fracture Many different or even contradictory standpoints have made it
nature have serious implications for the use of heavily drawn pearlitic difficult to understand the conditions responsible for the torsion

433
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

Table 2
The influence of cementite upon the torsion performance of pearlitic steel wires.
Cementite microstructure Deformation behavior during torsion Torsion performance

Non-aligned Dislocations multiplied along the lamellar direction Uniform torsion


Aligned lamellae (cold-drawing) Cementite containing amorphous phase favoring plastic deformation Uniform torsion
Aligned lamellae (drawing + annealing) Dislocations impeded by cementite plates consisting of recrystallized particles Local severe deformation /Delamination
Spherical cementite (Spheriodization) Dislocations bowing out the cementite particles Uniform torsion

delamination of pearlitic steel wires. Firstly, the dissolution of ce- fabricated with drawing strain ε > 1.76.
mentite is found to be insignificant in the pearlitic steel wires prepared 2. During the torsion test, dislocation multiplication occurred in ferrite
from a drawing strain ε < 20 [38,39]. Suzuki et al. have also reported along the lamellar direction in the wires with pearlite lamellae not
that the residual stress in the pearlitic steel wires were hardly affected parallel to the drawing direction. In heavily drawn pearlitic steel
by low temperature annealing at 150 °C [19]. It suggests that the two wires with aligned pearlite lamellae, cementite contains a mix of
factors, the distribution of carbon and residual stress, might not be key amorphous and nanocrystalline phases and the dislocation motion
inducements for torsion delamination. Based on the experimental re- in ferrite was almost unhindered. By contrast, following the an-
sults and discussions above, the nature of cementite in the pearlitic steel nealing treatment, the cementite plates comprise primarily nano-
wires is believed to play an important role in the torsion performance. crystalline grains, which acted as barriers against dislocation motion
The influence of cementite upon the torsion behavior can be summar- during the torsion test.
ized as Table 2. 3. The spheroidization of cementite can prevent the pearlitic steel
The cracks generated at small torsion strains are supposedly re- wires from torsion delamination. The conditions for the torsion
sponsible for delamination fracture of the annealed pearlitic steel wires delamination can thus be described as follows: a) the formation of
[24,28,30,37]. Notably, however, both previous reports and the present axially-aligned lamellar structure, and b) the transformation of ce-
study have found that not all of annealed pearlitic steel wires exhibited mentite substructure from amorphous to nanocrystalline state,
delamination fracture during the torsion tests. Firstly, the annealing without incurring spheroidization.
temperature is believed to be one of the key factors. As stated above,
the occurrence of cementite spheroidization is observed in the pearlitic Acknowledgement
steel wires after annealing treatment at 450 °C and 550 °C. The torsion
ductility of pearlitic steel wires is improved as expected. Moreover, the This work was supported by the Science and Technology
annealing treatments at low temperatures such as 140 °C [58] or 120 °C Advancement Program of Jiangsu Province (grant no. BA2017112), the
[37] were unable to cause torsion delamination of the pearlitic steel Six Talent Peaks Project in Jiangsu Province (2015-XCL-004) and the
wires. Secondly, the drawing strain of the pearlitic steel wires is another Industry-University Strategic Research Fund of Jiangsu Province
key factor. In the present study, only when the drawing strain ε of the (BY2016076-08). The study was also partly supported by the
wires subjected to annealing treatment is greater than 1.76, the dela- Fundamental Research Fund for the Central Universities and the
mination fracture would occur. Besides, the occurrence of torsion de- Scientific Research Foundation of the Graduate School of Southeast
lamination have not been reported in the wires fabricated from drawing University (YBJJ1674). Dr. Z. Xie acknowledges the support of the
strain ε < 1.33 [25,30,31,37,59,60]. It might be due to the alignment Australian Research Council Discovery Projects. Authors also thank C.
imperfection of pearlite lamellae in those wires, since the orientation W. Jin for assistance with TEM analysis.
and its distribution of pearlitic lamellae would influence the torsion
performance. Finally, metallurgical defects (abnormal local micro- Data availability statement
structure, pro-eutectoid cementite or inclusions) and surface cracks
might also need to be considered [2]. These defects can increase the All data included in this study are available upon request by con-
possibility of delamination crack generation. For example, torsion de- tacting the corresponding author.
lamination has been reported in as-drawn wires or in the wires un-
dergoing spheroidization with short time [29,31,34,35,61]. The dela- References
mination was possibly caused by the metallurgical or surface defects
which have served as crack precursors in the pearlitic steel wires. [1] D. Raabe, P.-P. Choi, Y. Li, A. Kostka, X. Sauvage, F. Lecouturier, K. Hono,
R. Kirchheim, R. Pippan, D. Embury, MRS Bull. 35 (12) (2010) 982–991.
[2] M. Zelin, Acta Mater. 50 (17) (2002) 4431–4447.
5. Conclusions [3] S. Goto, R. Kirchheim, T. Al-Kassab, C. Borchers, Trans. Nonferrous Met. Soc. China
17 (6) (2007) 1129–1138.
The influence of both the drawing strain and post-drawing an- [4] S. Djaziri, Y. Li, G. Nematollahi, B. Grabowski, S. Goto, C. Kirchlechner, A. Kostka,
S. Doyle, J. Neugebauer, D. Raabe, Adv. Mater. 28 (35) (2016) 7753–7757.
nealing treatment on the torsion behavior of pearlitic steel wires is [5] K. Han, D. Edmonds, G. Smith, Metall. Mater. Trans. A 32 (6) (2001) 1313–1324.
investigated. The relationship between the torsion performance and the [6] K. Han, T. Mottishaw, G. Smith, D. Edmonds, Mater. Sci. Technol. 10 (11) (1994)
microstructure (in particular cementite features) is examined. The 955–963.
[7] K. Han, G. Smith, D. Edmonds, Metall. Mater. Trans. A 26 (7) (1995) 1617–1631.
conditions responsible for the torsion delamination of pearlitic steel
[8] J. Embury, R. Fisher, Acta Metall. 14 (2) (1966) 147–159.
wires are discussed. The conclusions can be drawn as follows: [9] G. Langford, Metall. Mater. Trans. B 1 (2) (1970) 465–477.
[10] J.G. Sevillano, Mater. Sci. Eng. 21 (1975) 221–225.
[11] F. Fang, Y. Zhao, P. Liu, L. Zhou, X.-j. Hu, X. Zhou, Z.-h. Xie, Mater. Sci. Eng.: A 608
1. Both the torsion strength and torsion fracture strain of pearlitic steel
(2014) 11–15.
wires are simultaneously improved by the cold drawing up to [12] C. Borchers, T. Al-Kassab, S. Goto, R. Kirchheim, Mater. Sci. Eng.: A 502 (1) (2009)
ε < 0.65, that is, from 980 MPa to 1300 MPa and 0.95–1.65, re- 131–138.
spectively. With the drawing strain ε further increasing to 2.06, the [13] J. Languillaume, G. Kapelski, B. Baudelet, Acta Mater. 45 (3) (1997) 1201–1212.
[14] M. Hong, K. Hono, W. Reynolds, T. Tarui, Metall. Mater. Trans. A 30 (3) (1999)
torsion strength increases to 1780 MPa but the torsion fracture 717–727.
strain decrease to 1.1. After annealing treatment at 250 °C, the [15] N. Maruyama, T. Tarui, H. Tashiro, Scr. Mater. 46 (8) (2002) 599–603.
torsion strength of the wires increases at the expense of torsion [16] Y. Chen, G. Csiszár, J. Cizek, X. Shi, C. Borchers, Y. Li, F. Liu, R. Kirchheim, Metall.
Mater. Trans. A 47 (2) (2016) 726–738.
fracture strain. Moreover, torsion delamination occurs in the wires [17] X. Zhang, A. Godfrey, N. Hansen, X. Huang, Acta Mater. 61 (13) (2013) 4898–4909.

434
L. Zhou et al. Materials Science & Engineering A 743 (2019) 425–435

[18] X. Zhang, N. Hansen, A. Godfrey, X. Huang, Acta Mater. 114 (2016) 176–183. [41] C. Borchers, Y. Chen, M. Deutges, S. Goto, R. Kirchheim, Philos. Mag. Lett. 90 (8)
[19] T. Suzuki, Y. Tomota, M. Isaka, A. Moriai, N. Minakawa, Y. Morii, ISIJ Int. 44 (8) (2010) 581–588.
(2004) 1426–1430. [42] Y. Li, P. Choi, S. Goto, C. Borchers, D. Raabe, R. Kirchheim, Acta Mater. 60 (9)
[20] M. Elices, J. Mater. Sci. 39 (12) (2004) 3889–3899. (2012) 4005–4016.
[21] X. Zhang, N. Hansen, Y. Gao, X. Huang, Acta Mater. 60 (16) (2012) 5933–5943. [43] T. Furuhara, T. Mizoguchi, T. Maki, ISIJ Int. 45 (3) (2005) 392–398.
[22] C. Borchers, R. Kirchheim, Prog. Mater. Sci. 82 (2016) 405–444. [44] Y. Funakawa, T. Shiozaki, K. Tomita, T. Yamamoto, E. Maeda, ISIJ Int. 44 (11)
[23] Y. Li, D. Raabe, M. Herbig, P.-P. Choi, S. Goto, A. Kostka, H. Yarita, C. Borchers, (2004) 1945–1951.
R. Kirchheim, Phys. Rev. Lett. 113 (10) (2014) 106104. [45] F. Fang, J.-h. Jiang, S.-Y. Tan, A.-b. Ma, J.-q. Jiang, Surf. Coat. Technol. 204 (15)
[24] B. Goes, A. Martı́n-Meizoso, J. Gil-Sevillano, I. Lefever, E. Aernoudt, Eng. Fract. (2010) 2381–2385.
Mech. (1998) 255–272. [46] Z. Lichu, H. Xianjun, M. Chi, Z. Xuefeng, F.F. Jianqing JIANG, Acta Metall. Sin. 51
[25] C. Cordier-Robert, B. Forfert, B. Bolle, J.-J. Fundenberger, A. Tidu, J. Mater. Sci. 43 (8) (2015) 897–903.
(4) (2008) 1241–1248. [47] T. Takahashi, M. Nagumo, Trans. Jpn. Inst. Met. 11 (2) (1970) 113–119.
[26] F. Fang, L. Zhou, X. Hu, X. Zhou, Y. Tu, Z. Xie, J. Jiang, Mater. Des. 79 (2015) [48] X. Zhang, A. Godfrey, W. Liu, Q. Liu, Mater. Sci. Technol. 27 (2) (2011) 562–567.
60–67. [49] W.J. Nam, C.M. Bae, Mater. Sci. Eng. A 203 (1) (1995) 278–285.
[27] T. Tarui, S. Nishida, A. Yoshie, H. Ohba, Y. Asano, I. Ochiai, T. Takahashi, Nippon [50] X. Zhang, A. Godfrey, X. Huang, N. Hansen, Q. Liu, Acta Mater. 59 (9) (2011)
Steel Tech. Report. 80 (0) (1999). 3422–3430.
[28] M. Tanaka, H. Saito, M. Yasumaru, K. Higashida, Scr. Mater. 112 (2016) 32–36. [51] Y. Chen, G. Csiszár, J. Cizek, S. Westerkamp, C. Borchers, T. Ungár, S. Goto, F. Liu,
[29] C.M. Bae, W.L. Nam, C.S. Lee, Scr. Mater. 35 (5) (1996) 641–646. R. Kirchheim, Metall. Mater. Trans. A 44 (8) (2013) 3882–3889.
[30] L. Zhou, F. Fang, L. Wang, H. Chen, Z. Xie, J. Jiang, Mater. Sci. Eng.: A 713 (2018) [52] F. Fang, X.-j. Hu, S.-h. Chen, Z.-h. Xie, J.-q. Jiang, Mater. Sci. Eng.: A 547 (2012)
52–60. 51–54.
[31] S. Joung, U. Kang, S. Hong, Y. Kim, W. Nam, Mater. Sci. Eng.: A 586 (2013) [53] Y. Li, A. Kostka, P. Choi, S. Goto, D. Ponge, R. Kirchheim, D. Raabe, Acta Mater. 84
171–177. (2015) 110–123.
[32] H.M. Baek, S.K. Hwang, H.S. Joo, Y.-T. Im, I.-H. Son, C.M. Bae, Mater. Des. (1980- [54] W. Guo, Y. Meng, X. Zhang, V. Bedekar, H. Bei, S. Hyde, Q. Guo, G.B. Thompson,
2015) 62 (2014) 137–148. R. Shivpuri, J.-m. Zuo, Acta Mater. 152 (2018) 107–118.
[33] D. Park, E. Kang, W. Nam, J. Mater. Process. Technol. 187 (2007) 178–181. [55] H.G. Read, W. Reynolds, K. Hono, T. Tarui, Scr. Mater. 37 (8) (1997) 1221–1230.
[34] J. Lee, J. Lee, Y. Lee, K. Park, W. Nam, J. Mater. Process. Technol. 209 (12) (2009) [56] A. Khalajhedayati, Z. Pan, T.J. Rupert, Nat. Commun. 7 (2016).
5300–5304. [57] N. Min, W. Li, X. Jin, X. Wang, T. Yang, C. Zhang, Acta Metall. Sin.-Chin. Ed. 42 (10)
[35] K. Shimizu, N. Kawabe, ISIJ Int. 41 (2) (2001) 183–191. (2006) 1009.
[36] Y. Yang, J. Bae, C. Park, Mater. Sci. Eng.: A 508 (1) (2009) 148–155. [58] S.K. Lee, D.C. Ko, B.M. Kim, Mater. Des. 30 (8) (2009) 2919–2927.
[37] X. Hu, L. Wang, F. Fang, Z. Ma, Z.-h. Xie, J. Jiang, J. Mater. Sci. 48 (16) (2013) [59] N. Guo, B. Song, B.-S. Wang, Q. Liu, Acta Metall. Sin. (Engl. Lett.) 28 (6) (2015)
5528–5535. 707–714.
[38] K. Hono, M. Ohnuma, M. Murayama, S. Nishida, A. Yoshie, T. Takahashi, Scr. [60] T.-Z. Zhao, S.-H. Zhang, G.-L. Zhang, H.-W. Song, M. Cheng, Mater. Des. 59 (2014)
Mater. 44 (6) (2001) 977–983. 397–405.
[39] Y. Li, P. Choi, C. Borchers, S. Westerkamp, S. Goto, D. Raabe, R. Kirchheim, Acta [61] A. Durgaprasad, S. Giri, S. Lenka, S.K. Sarkar, A. Biswas, S. Kundu, S. Mishra,
Mater. 59 (10) (2011) 3965–3977. S. Chandra, R. Doherty, I. Samajdar, Metall. Mater. Trans. A (2018) 1–11.
[40] L. Zhou, F. Fang, X. Zhou, Y. Tu, Z. Xie, J. Jiang, Scr. Mater. 120 (2016) 5–8.

435

You might also like