You are on page 1of 18

Author’s Accepted Manuscript

Strain hardening and nanocrystallization behaviors


in Hadfield steel subjected to surface severe plastic
deformation

Chen Chen, Bo Lv, Xiaoyong Feng, Fucheng


Zhang, Hossein Beladi
www.elsevier.com/locate/msea

PII: S0921-5093(18)30717-2
DOI: https://doi.org/10.1016/j.msea.2018.05.059
Reference: MSA36497
To appear in: Materials Science & Engineering A
Received date: 11 April 2018
Revised date: 12 May 2018
Accepted date: 18 May 2018
Cite this article as: Chen Chen, Bo Lv, Xiaoyong Feng, Fucheng Zhang and
Hossein Beladi, Strain hardening and nanocrystallization behaviors in Hadfield
steel subjected to surface severe plastic deformation, Materials Science &
Engineering A, https://doi.org/10.1016/j.msea.2018.05.059
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Strain hardening and nanocrystallization behaviors in Hadfield
steel subjected to surface severe plastic deformation
Chen Chena, Bo Lvb,*, Xiaoyong Fenga, Fucheng Zhanga, c, Hossein Beladid
a
State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao
066004, China
b
College of Environmental and Chemical Engineering, Yanshan University, Qinhuangdao 066004, China
c
National Engineering Research Center for Equipment and Technology of Cold Strip Rolling, Yanshan
University, Qinhuangdao 066004, China
d
Institute for Frontier Materials, Deakin University, Geelong, VIC 3216, Australia
*Corresponding author. Tel.: +86 335 8063949; e-mail addresses: lvbo@ysu.edu.cn (B. Lv).
Abstract:
A gradient nanocrystalline layer with a thickness in a range of millimeter magnitude was
successfully produced on the surface of Hadfield steel by a novel severe plastic deformation
technology, high speed pounding. The surface hardness was measured, and the microstructure
evolution during nanocrystallization process was characterized by X-ray diffraction and
transmission electron microscopy. Results showed that the hardness increment and
nanocrystallization in Hadfield steel were obtained at different stages under high speed
pounding. The first stage was strain hardening, where surface hardness of the Hadfield steel
increased gradually during high speed pounding until a steady-state value was obtained. The
hardening degree and rate of the Hadfield steel were determined by deformation stress and
strain rate, respectively. The second stage was microstructure nanocrystallization, at which
twin boundaries interacted with dislocations to form general high angle grain boundaries. In
this stage, the surface hardness of Hadfield steel remained basically the same. Moreover, a
physical model was established to explain the strain hardening and surface nanocrystallization
behaviors in accordance with the microstructure evolution at different stages in Hadfield steel.

Keywords: Severe plastic deformation, Strain hardening, Nanocrystalline, Hadfield steel

1. Introduction
Surface severe plastic deformation (S2PD) procedure can be used as a typical surface
nanocrystallization technology to refine the surface of conventional coarse-grained materials
to obtain gradient microstructure throughout the thickness [1-2]. This procedure has been used
to produce nanocrystalline layer without any voids and defects at the surface of various metals
and alloys, such as iron alloy [3], copper [4, 5], carbon steel [6], stainless steel [7, 8] and
nickel/nickel-based alloys [9, 10]. Lu et al. [11, 12] prepared cylindrical gradient

1
nanocrystallized samples in nickel and copper by surface mechanical attrition, displaying
excellent mechanical properties. However, producing nanocrystalline material in large scale
remains challenging.
The excellent mechanical properties of gradient nanocrystalline materials prepared by S2PD
can be attributed to strain hardening and microstructure refinement. Many studies have
analyzed the microstructure refinement mechanisms through S2PD processing in different
metals and alloys [2, 7, 12-16]. Dislocation slip is known as the main deformation mechanism
for face-centered cubic (fcc) structured alloys with high stacking fault energy (i.e. >35 mJ/m2
[17]), displaying complicated dislocation interaction mechanisms, namely dislocation
generation, rearrangement and annihilation, which contribute to grain refinement. On the
other hand, strain induced martensitic transformation is another refinement mechanism, which
is mainly observed in materials with a low stacking fault energy (i.e. <20 mJ/m2 [18]).
However, no systematic research has reported about the relationship between strain hardening
and microstructure refinement behaviors for Hadfield steel with medium-low stacking fault
energy during plastic deformation with different degrees, during which deformation twins
coordinate with dislocations.
Stacking fault energy is the main factor that affects plastic deformation mechanisms such as
dislocation slip, martensitic transformation and mechanical twining. Hadfield steel has
medium-low stacking fault energy, whose plastic deformation mechanism is different from
iron and copper metals [19]. Hadfield steel possesses good work hardening capacity, high
strength and toughness after hardening, and it has been widely used in industrial applications.
Nanocrystalline layers have been obtained on the surface of railway crossings and
wear-resistant liners made from Hadfield steel subjected to rolling contact fatigue and
friction-wear service [20, 21]. The size of the nanocrystalline on the surface of the Hadfield
steel railway crossings is about 30 nm, and the hardness of the nanocrystalline layers is about
550 HV [20]. Meanwhile, the size and hardness of the nanocrystalline formed on the surface
of the wear-resistant liners are 20 nm and 600 HV, respectively [21]. Nanocrystalline with a
size of 3-10 nm could be obtained on the surface of an austenitic manganese steel after shot
peening treatment [22]. However, the formation mechanism of nanocrystalline and strain
hardening behavior in austenitic manganese steels have not been thoroughly studied.
High speed pounding (HSP) is a novel type of surface optimization technique in metals. It
has been utilized in surface hardening of Hadfield steel crossings to produce surface
nanocrystalline layer [23, 24]. HSP can introduce high-frequency impact stress to the sample
surface, and cause plastic deformation with a large strain rate in deep surface layer. In the
2
present study, HSP treatment with different parameters was conducted on the Hadfield steel,
and nanocrystalline layer with different levels was obtained on the sample surface. Strain
hardening and microstructure refinement behaviors were systematically studied by analyzing
the microstructures, hardness, dislocation density, deformation twin density, and grain size on
the surfaces of Hadfield steel samples subjected to HSP for different times. The relationship
between strain hardening, nanocrystallization and pounding treatment parameters (strain rate,
deformation stress, and pounding time) were also analyzed.
2. Experimental procedures
The material used in the current investigation was Hadfield steel plate with a thickness of
25 mm having the chemical composition of 1.20 C, 12.30 Mn, 0.60 Si, 0.016 S, 0.022 P, and
balanced with Fe (in wt.%). Prior to S2PD, the samples were heat treated at 1050 °C for 60
min followed by water quenching to obtain a uniform austenitic microstructure with an
average grain size of 138 ± 20 m. HSP was carried out as a novel S2PD technology on
samples under different parameter conditions. The HSP set-up and procedure were described
in detail elsewhere [23, 24] and shown schematically in Figure 1. During HSP, the Hadfield
steel samples were subjected to different deformation stresses, namely 1.2, 1.4, and 1.6 GPa,
and marked as S1.2, S1.4, and S1.6, respectively. The strain rate was estimated using finite
element modeling. The maximum strain rates were calculated to be 0.96×103 s−1, 1.47×103 s−1,
and 2.13×103 s−1 on the pounded surfaces for S1.2, S1.4, and S1.6 conditions, respectively.
The strain rate decreased with an increase in depth from the pounded surface. The samples
were also subjected to different pounding times for different deformation stress conditions.
During the pounding process, the surface of sample was cooled by recycled water to maintain
the surface temperature at ambient.

Fig. 1. Schematic illustration of the HSP treatment set-up.


The hardness distribution on the cross-section of the samples was measured with a

3
micro-hardness tester under a load of 200 g for 10 s. Three measurements were done for each
point. The microstructures were examined by X-ray diffraction (XRD) and transmission
electron microscopy (TEM). The XRD patterns of the processed surfaces were collected in a
D/max-2500/PC XRD with a Cu Kα radiation (wavelength (λ)=1.540598 Å) at continuous
scanning mode for 2θ values between 30 and 100°, a step size of 0.02°, and generator settings
of 40 kV and 100 mA.
The micro-strain <ε2>1/2 and average size of sub-grain, D, were calculated using the
classical Williamson–Hall method [25]:
 cos  1 1/2 2sin 
 2 2 ( )
 D 
where λ is the wavelength of the X-ray radiation, β is the full width at the half maximum. The
dislocation density, ρ, was estimated by the following equation:
1/2
  2 3 2 / ( Db)

where b is the Burgers vector.


A JEM-2010 transmission electron microscopy was used to observe the fine
microstructures in the HSP-processed samples. TEM samples were prepared at two
cross-sections, parallel and perpendicular to the plane of the pounded surface for different
conditions. The former was polished from the non-pounded side down to 40μm, whereas the
latter was polished from both sides. Subsequently, the samples were thinned on one side to
perforation using a precision ion polishing system. The overall length of twin boundaries
(TBs) was calculated by inspecting more than 20 TEM images. It was then divided by area of
all TEM images to obtain the twin density. The twin lamella thickness was also measured.
3. Results
Figure 2 shows the hardness distribution curves along the depth of cross-sectional samples
subjected to different HSP processes. The surface of all samples was hardened for different
conditions. The hardness curve had typically an opposite-S shape, which can be divided into
three hardness regions. The region I was located in the vicinity of pounded surface, where the
maximum hardness was observed and appeared to be nearly constant to a certain depth from
the surface, beyond which the hardness gradually decreased with the depth with referring to
region Ⅱ. The hardness at region Ⅲ was similar to the as-received material, which did not
alter with the depth. For the samples processed by the same deformation stress, the surface
(region I) hardness progressively increased with an increase in the pounding time up to 4×104
times beyond which it became nearly similar. This suggests that there was a steady-state work

4
hardening, which depended on the deformation stress, namely ~ 620 HV200gf, ~ 670 HV200gf,
and ~ 780 HV200gf for S1.2, S1.4, and S1.6 samples, respectively.

Fig. 2. Micro-hardness variations as a function of depth from the surface of sample subjected to HSP at
different conditions: (a) S1.2; (b) S1.4; (c) S1.6. (Insets in “a” through “c” correspond to the designated
boxes).

Fig. 3. TEM micrographs of HSP-processed sample surface at 8×104 pounding times under different
deformation stresses: (a) S1.2; (b) S1.4; (c) S1.6.
Surface TEM micrographs of samples subjected to 8×104 pounding times at different
deformation stresses revealed uniform ultrafine grains, mostly equiaxed, having a random
crystallographic orientation, as indicated in their corresponding selected area electron
diffraction (SAED) patterns (Fig. 3). The average grain size was 56, 44, and 23 nm for S1.2,
S1.4, and S1.6 sample conditions, respectively. In other words, the extent of surface
nanocrystallization enhanced with an increase in the pounding deformation stress, which was
consistent with the trend of corresponding strain-hardening at the surface (Fig. 2).
TEM characterization of S1.6 sample surfaces pounded for different times were conducted
to study the microstructure evolution of Hadfield steel during HSP process (Fig. 4). At 1×104
pounding time, the dislocation tangles and dislocation cells were mainly observed in grain
interiors, suggesting that the dislocation slip was the dominant deformation mechanism at an
early stage of S2PD. The dislocation cells were large with thick walls, inside which the
dislocation density was very low (shown by arrows in Fig. 4a). It is worth mentioning that the
high-density dislocation walls were occasionally observed.

5
Fig. 4. TEM images of S1.6 sample HSP-processed for different times: (a) 1 × 104, (b) 2 × 104, (c) 4 × 104.
With increasing the HSP pounding times to 2×104, the plastic strain increased and a large
number of deformation twins appeared on the surface (shown by arrows in Fig. 4b). The
deformation twins with different crystallographic variants were largely intersected, resulting
in rhombic blocks within the microstructure. This suggests that the deformation twinning was
the dominant deformation mechanism in the present pounding condition, even though the
dislocation activities were also observed (Fig. 4b). In this stage, the intersected deformation
twins divided the parent grains into two-dimensional parallelogram structured units, with a
size ranging from dozens to hundreds of nanometers. These units were separated by straight
boundaries, containing numerous dislocation tangles inside, but no dislocation cells were
observed. In this situation, dislocation glide was blocked, and dislocations interacted with TBs
frequently. These intersected twins set a frame for microstructure refinement and the
formation of equiaxed nano-grains upon further plastic deformation.

At 4×104 pounding times, sub-grains in strip shape were mostly formed at the sample
surface (Fig. 4c). The size of the quadrangular sub-grains was below 100 nm and the shape
was very similar to the quadrilaterals formed by intersected deformation twins. The sub-grains
might be formed by the interaction of dislocations and intersected twins. Moreover, fine twins
were observed in some sub-grains, which further divided the microstructure. When the
pounding times were further increased, nanocrystalline structures gradually formed (Fig. 3c).
Figure. 5 shows the microstructures at 5 μm and 800 μm beneath the surface of the S1.6
sample pounded for 8×104 times. Two-sectional TEM analysis of microstructure at 5 μm
beneath the surface revealed equiaxed nano-grains without deformation twins observed (Fig.
5a). The overall microstructure had a relatively random texture. In other words, the
deformation twin density was significantly reduced as nanocrystallization was further
developed. This suggests that the TBs transformed into the high-angle grain boundaries (GBs)

6
during plastic deformation.
At 800 μm beneath the surface, the nanocrystallization was less developed, showing
two-dimensional elongated nano-sized grains with ~50 nm thickness and the length/width
ratio in a range of 1 to 4 (Fig. 5b). Strain rate on the sample decreased with increasing the
distance from the surfaces during S2PD process [10]. It suggests that the microstructure
differences between the surface (5 µm) and sub-surface (800 µm) to some extent were caused
by strain rate. Larger strain rate accelerated the formation of nanocrystals, which was in
accordance with the observations in Figure 3. The elongated diffraction spots in the SAED
indicated a strong texture in the elongated grains, which revealed a small misorientation
among the lamellas. During plastic deformation of the Hadfield steel in HSP process, sample
surface was mainly subjected to pounding stress in the vertical direction. Therefore, it was
deduced that the layer-shaped nanostructures were obtained from the plastic deformation of
ultrafine grains and even nanocrystals. Figure 5c shows twins with a thickness of less than 80
nm in the perpendicular direction of the lamellar grains, which were similar to the
microstructures in Figure 4c. However, their grain size and forced direction were different.
This suggests that a nanocrystalline layer with 800 m in depth could be prepared on the
surface of the Hadfield steel via HSP technology.

Fig. 5. TEM microstructures of S1.6 sample subjected to HSP-process for 8×104 times at different depth
from the surface: (a) 5 μm; (b) 800 μm; (c) a high magnification observation of “b”.
4. Discussion
4.1 Strain hardening behavior

The deformation field created by spherical surface poundings in the HSP process can be
7
approximated by the elastic contact between a sphere and a semi-infinite solid, which is
similar with the process of surface mechanical attrition and shot peening [22, 26]. Based on
Hertzian elastic contact theory [27], we can estimate the depth of plastic deformation region

h 2  1
by the following equation:  3( )1/4 ( b )1/4 , where is the contact stress, R is the
R 3  3

1
spherical radius, is the sphere density, and 1 is the pounding velocity. The most effective
2

factor to the plastic deformation region is the spherical radius, and the lesser factors are
contact stress, sphere density and pounding velocity, whereas the duration has no influence.
This is consistent with the current observation, where the depth of plastic deformation region
shows little change as the HSP duration increases (Fig. 2).
According to the strain hardening curves in Figure 2, the relations among deformation
stress, strain rate, stable hardness, and the corresponding pounding times were determined
(Table 1). The following relationship was found between the stable hardness and the
ΔHv
corresponding deformation stress:  k , where ΔHv is the hardness increment (the
S  σs
difference value between the stable hardness and the original hardness of the Hadfield steel
(~220 Hv200gf)), S is the deformation stress,  s is the yield strength of the Hadfield steel

(~400 MPa), and k is a constant. As shown in Figure 6a, ΔHv presents a good linear
relationship with (S-σs), the slope k is a constant (k ≈ 0.45). The relation between the strain
Rs
rate and surface hardening rate could be expressed as  k  in a given deformation stress,
RH

where Rs is the strain rate, RH is the surface hardening rate and k ' is a constant ( k ' ≈

4.2×107). The constants k and k ' were determined by the deformation condition of the
Hadfield steel. In this work, considering different hardening degrees under each stress, the
reciprocal of pounding times, T, at which the steel reaches the steady-state hardness value, is
chosen as the hardening rate, as shown in Figure 6b.
Table 1. Relationship between deformation stress, strain rate, stable hardness, and the corresponding
pounding times
Deformation stress (GPa) 1.2 1.4 1.6
Steady-state hardness (Hv) 615.0 670.0 780.0
3 -1
Strain rate (×10 s ) 0.96 1.47 2.13

8
Pounding times (×104) 4 3 2

Fig. 6. (a) The steady-state hardness value as a function of deformation stress and (b) the strain rate versus

surface hardening rate.


Based on the above analysis, the deformation stress and strain rate are two important
parameters in the strain hardening process of the Hadfield steel. The plastic deformation stress
determines the steady-state surface hardness value, and the strain rate influences the
hardening rate of the sample. Higher hardening rate is caused by more active microstructure
evolution. Similarly, high strain rate promotes further microstructural nanocrystallization
process.

Fig. 7. The twin density, dislocation density, hardness value, and grain size on the surface of the S1.6
sample as a function of pounding time.
According to the previous reports, high-density dislocations and twins can prohibit the
initiation of the slip systems and dislocation movement during plastic deformation, which
primarily causes the work hardening of metals [28, 29]. For S2PD process of Hadfield steel,
GBs can also be one of the main barriers against the dislocation movement [30, 31], which
can contribute to the hardness (i.e., work-hardening) of Hadfield steel as a large number of
9
GBs is introduced through this process.
The hardness, twin density, dislocation density, and grain size of S1.6 at the surfaces are
illustrated as a function of pounding time in Figure 7. At the initial stage of the HSP, the
surface hardness of the S1.6 sample (black line) increased with an increase in the pounding
time. Interesting, the level of surface hardness became nearly unchanged beyond 2×104
pounding times. The dislocation density of the sample surface responds sensitively to the
pounding times, i.e., increasing rapidly with an increase in the pounding times up to 2×104
times beyond which it decreased promptly (Fig .7). The twin density as a function of
pounding times is also similar to that of dislocation density, though the maxima appears at
4×104 times (Fig. 7). The grain size is nearly constant up to the pounding times of 2×104 and
then dramatically decreases with the pounding times because of the introduction of
sub-grains/grains (Fig. 7).
The extent of hardness increasement (i.e., work-hardening) at the surface during
pounding process strongly depends on the characteristics of deformation features formed in
the microstructure, namely the twin lamella thickness, dislocation density, and sub-grain/grain
size. Accordingly, a formula based on the Hall-Petch relationship is introduced to
quantitatively estimate the effects of these microstructural features on the hardness:
1
1 1 21 1
H  H O  k[( )2  ( ) ]  αGbρd2 (1)
dtwin d

where H O , k and  (0.2~0.6) are constants, dtwin and d are twin lamella thickness and
1
sub-grain/grain size in the second term, respectively [32, 33], whereas the third term αGbρ 2
d

is the empirical formula to express the dislocation density effects on the hardness (i.e., the
classical Taylor formula) [34], G is the shear modulus, b is the Burgers vector and ρd is the
dislocation density. In calculation, G is taken as 79 GPa, b is 0.26 nm (fcc structure
(111)[112]), and α is 0.6 [34]. When the sub-grain/grain size is d, the dislocation density is ρd,
1
and the corresponding twin lamella thickness is dtwin, the formula, H  αGbρ 2
d and
1
1 1 21
( )2  ( ) exhibits approximately linear relationship between the pounding times and
dtwin d

both hardness and work-hardening (Fig. 8).

10
1
1
1 1 21
Fig. 8. The relationship between H  αGbρ 2
d and ( )  ( ) for the S1.6 sample at different
2
dtwin d

pounding times: data taken from surfaces of the undeformed sample, the HSP-processed samples with

ponding times of 1×104, 2×104, 4×104, 6×104, and 8×104

According to Fig. 7, before reaching the stable surface hardness value, no new GBs are
introduced (i.e., no grain refinement). Hence, the surface hardness is determined by
dislocation and twin density. During the plastic deformation of the Hadfield steel, the
deformation twinning occurs beyond a given strain, mostly initiated from stacking faults,
which in turn formed through the dislocation dissociation [6, 35]. At an early stage of
twinning, the twin density is small. The annihilation rate of dislocations was smaller than the
generation rate caused by plastic deformation. The dislocation density and twin density both
increased, resulting in a rising surface hardness of the Hadfield steel. Subsequently, some new
twins with small lamellar spacing continued to form inside the grains, and the high-density
TBs greatly blocked the dislocation movement. Combined with Eq. (1), the hardness also
increased continuously.
As the strain accumulates, the dislocation density gradually reaches the maximum at which
the grains can accommodate, and the twin density also remains high. In this case, the surface
hardness reaches the peak value (Fig. 7). Afterwards, the high-density dislocations inevitably
interact with TBs. The coherent TBs not only can block the dislocation movement effectively,
which is similar to traditional GBs, but also can supply abundant dislocation-stored spaces [30,
36, 37]. As twin density increases, the flow stress of the material enhances gradually,
providing sufficient driving force for the interaction of TBs and dislocations. The formation of
twins, the dislocation interactions, and the interactions between dislocations and twins are
11
processes of dislocation annihilation, which may result in a reduction of dislocation density.
Notably, the low thickness (nanoscale) of the twin lamella is accompanied by increased strain
and twin density. The reduction of twin lamella thickness or the increase in the twin density
further promote the interaction between dislocations and TBs [31]. The TB coherency reduces
with the accumulation of the dislocations and consequently transforming TBs to high-angle
GBs [37]. However, the transformation process consumes the TBs, resulting in a decreased
twin density when the pounding times exceeds 4×104.
GBs and TBs can both block the dislocation movement. Hence, the transformation from
TBs to high-angle GBs is only a change of block form, rather the change in free path for
1
1 1 21
dislocation glide. Thus, the term k[( )2  ( ) ] in Eq. (1) almost does not change when
dtwin d
the TBs transform to GBs at an early stage of nanocrystallization process. However, with the
accumulation of plastic deformation, twins are further introduced in the fine grains (Fig. 5c),
progressively reducing the free path for dislocation glide. Therefore, the contribution of twins
and sub-grains/grains on the surface hardness is enhanced. Though a decrease in the
dislocation density reduces the surface hardness, the range at which the dislocation density
1
influences the hardness is only 0.1–0.6 GPa (  Gbd 2
is calculated using the maximum and
minimum values of dislocation density). When the surface hardness reaches a steady-state
level, the work hardening loss caused by the decreased dislocation density is relatively small.
Thus, from Eq. (1), the surface hardness at this stage does not fluctuate significantly. As such,
the nearly unchanged surface hardness of the S1.6 sample under HSP after 2×104 pounding
times contributes to the dynamic balance of the density of dislocations, TBs, and GBs.
The Hadfield steel samples obtained different steady-state surface hardness values under
different deformation stresses (Fig. 2). Once the steady-state surface hardness (i.e., region Ⅰ)
is obtained, it remains unchanged with an increase in the pounding times, which is only
determined by the deformation stress. However, the deformation stress defines both the
surface strain rate and the plastic deformation degree on the sample surface. Each pounding
treatment can be regarded as a plastic deformation process. The application of higher
deformation stress introduces a greater plastic deformation, and more activities of dislocations
and deformation twins. Therefore, the S1.6 sample obtains the highest steady-state hardness
value with the least pounding times (Table 1). When the surface hardness of the samples
reaches a steady-state value, the yield strength of the Hadfield steel substantially increases to
~1400 MPa for the S1.6 sample. As HSP process is continuously conducted on the samples,
12
the surface plastic deformation degree reduces accordingly. In this situation, the critical stress
for the dislocation motion increases and the dislocation free path slip shortens, decreasing the
contribution of dislocations on work hardening [38, 39]. Meanwhile, the transformation from
twins to sub-grains/grains has little effect on the hardness of the Hadfield steel. As such, the
decreased dislocation activity is the precondition to obtain a steady-state surface hardness
value in the Hadfield steel.
4.2 Nanocrystallization behavior
As discussed in detail in our previous work [37], the interaction between dislocations and
TBs is the dominant grain refinement mechanism in the S2PD-processed Hadfield steel. The
interaction of dislocations and TBs alters the misorientation angle/axis of TBs upon
deformation. In other words, the TBs progressively lose their coherency and convert to
general grain boundary with the deformation [37]. Combined with the current results, the
strain hardening and microstructure refinement behaviors evolving with the pounding times
are modeled, as demonstrated schematically in Figure 9. The nanocrystallization process of
the Hadfield steel can be divided into two stages. The first stage is mostly strain hardening,
where a number of dislocations are generated (including dislocation tangles and walls) inside
the grains at an early stage of plastic deformation. However, no twins form because the
fcc-structured metals have many slip systems, which are easily activated in Hadfield steel [40].
Thus, dislocation slip is the main mechanism of plastic deformation for Hadfield steels. As the
plastic strain accumulates, twins and intersected twins start to populate the microstructures,
and twinning and dislocation slip become the dominant deformation mechanisms. At the first
stage, the deformation energy induces the accumulation of dislocations and twins, without
severe interactions between TBs and dislocations.

Fig. 9. Strain hardening and microstructure refinement behaviors evolving by pounding times

The second stage is mainly grain refinement. During S2PD of the Hadfield steel, the
13
dislocations do not form dislocation cells that break up the grain into sub-gains. Hence, at this
stage, the deformation energy provides driving force for the interactions of dislocations and
TBs. The interactions play a key role in the grain refinement behavior. The twin intersections
are found to be dominant in grain interiors, and they can divide the micro-scaled parent grains
into nano-scaled parallelogram-structured units. Subsequently, as dislocations and TBs
interact, TBs are gradually transformed into low-angle or high-angle boundaries so that
nanocrystalline is produced. Therefore, the dislocation–TB interactions can trigger the
formation of equiaxed nano-grains surrounded by high-angle GBs at the late stage of
nanocrystallization.

5. Conclusions

1. A nanocrystalline surface layer with a thickness of millimeter magnitude was prepared on a


Hadfield steel plate by using HSP technology.
2. Under HSP, the hardness increasement and nanocrystallization in Hadfield steel were
obtained at different stages. The first stage was the strain hardening, where the surface
hardness of the Hadfield steel increased continuously with increasing pounding times until a
stable value was reached. The second stage was microstructure nanocrystallization, at which
the surface hardness of the Hadfield steel remained basically the same, but the grains were
refined until the completion of nanocrystallization.
3. The work hardening degree and work hardening rate of Hadfield steels were determined by
deformation stress and strain rate, respectively. Higher stable hardness value could be
obtained under higher deformation stress, and faster work hardening rate could be obtained at
higher strain rate. The nanocrystallization degree and nanocrystallization rate of Hadfield
steels were directly proportional to the plastic strain accumulation (pounding times) and strain
rate, respectively.
4. At the work hardening stage of the Hadfield steel, a large number of dislocations firstly
formed inside grains, and dislocation slip became the main plastic deformation mechanism.
With the accumulation of plastic deformation and an increase in the dislocation density,
deformation twins were formed. Dislocations and deformation twins coordinated the plastic
deformation of the Hadfield steel, and their densities were increased gradually until a stable
14
surface hardness was obtained. At the nanocrystallization stage, the negative effect of
decreasing density of dislocations and deformation twins was compensated by the positive
effect of increasing amounts of sub-grain/grain boundaries on the surface hardness. Therefore,
the surface hardness remained unchanged.

Acknowledgements
The authors gratefully acknowledge the National Natural Science Youth Fund (No.
51501161).

Data availability
The raw/processed data required to reproduce these findings cannot be shared at this time as
the data also forms part of an ongoing study.

References
[1] N.R. Tao, Z.B. Wang, W.P. Tong, M.L. Sui, J. Lu, K. Lu, An investigation of surface
nanocrystallization mechanism in Fe induced by surface mechanical attrition treatment, Acta
Mater. 50 (2002) 4603-4616.
[2] N.R. Tao, M.L. Sui, J. Lu, K. Lu, Surface nanocrystallization of iron induced by ultrasonic
shot peening, Nanostruct. Mater. 11 (1999) 433-440.
[3] C.S. Wen, W. Li, Y.H. Rong, Nanocrystallization and martensitic transformation in
Fe-23.4Mn-6.5Si-5.1Cr (wt.%) alloy by surface mechanical attrition treatment, Mater. Sci.
Eng. A 481 (2008) 484-488.
[4] T.H. Fang, W.L. Li, N.R. Tao, K. Lu, Revealing Extraordinary Intrinsic Tensile Plasticity
in Gradient Nano-Grained Copper, Science 331 (2011) 1587-1590.
[5] S. Cheng, E. Ma, Y.M. Wang, L.J. Kecskes, K.M. Youssef, C.C. Koch, Tensile properties
of in situ consolidated nanocrystalline Cu, Acta Mater. 53 (2005) 1521-1533.
[6] T. Wang, D.P. Wang, G. Liu, B.M. Gong, N.X. Song, Investigations on the
nanocrystallization of 40Cr using ultrasonic surface rolling processing, Appl. Surf. Sci. 255
(2008) 1824-1829.
[7] X.H. Chen, J. Lu, L. Lu, K. Lu, Tensile properties of a nanocrystalline 316L austenitic
stainless steel, Scripta Mater. 52 (2005) 1039-1044.
[8] H.W. Zhang, Z.K. Hei, G. Liu, J. Lu, K. Lu, Formation of nanostructured surface layer on
AISI 304 stainless steel by means of surface mechanical attrition treatment, Acta Mater. 51
(2003) 1871-1881.
[9] J.C. Villegas, L.L. Shaw, Nanocrystallization process and mechanism in a nickel alloy
subjected to surface severe plastic deformation, Acta Mater. 57 (2009) 5782-5795.
15
[10] N.R. Tao, X.L. Wu, M.L. Sui, J. Lu, K. Lu, Grain refinement at the nanoscale via
mechanical twinning and dislocation interaction in a nickel-based alloy, J. Mater. Res. 19
(2004) 1623-1629.
[11] X.C. Liu, H.W. Zhang, K. Lu, Strain-Induced ultrahard and ultrastable nanolaminated
structure in nickel, Science 342 (2013) 337-340.
[12] W.L. Li, N.R. Tao, K. Lu, Fabrication of a gradient nano-micro-structured surface layer
on bulk copper by means of a surface mechanical grinding treatment, Scripta Mater. 59 (2008)
546-549.
[13] S.Y. Chang, B.D. Ahn, S.K. Hong, S. Kamado, Y. Kojima, D.H. Shin, Tensile
deformation characteristics of a nano-structured 5083 Al alloy, J. Alloys Compd. 386 (2005)
197-201.
[14] K.S. Kumar, H.V. Swygenhoven, S. Suresh, Mechanical behavior of nanocrystalline
metals and alloys, Acta Mater. 51 (2003) 5743-5774.
[15] H. Gleiter, Nanocrystalline Materials, Prog. Mater. Sci. 33 (1989) 223-315.
[16] C. Suryanarayana, Nanocrystalline materials, Int. Mater. Rev. 40 (1995) 41-64.
[17] S. Allain, J.P. Chateau, O. Bouaziz, S. Migot, N. Guelton, Correlations between the
calculated stacking fault energy and the plasticity mechanisms in Fe–Mn–C alloys, Mater. Sci.
Eng. A 387 (2004) 158-162.
[18] Y.K. Lee, C. Choi, Driving force for γ→ ε martensitic transformation and stacking fault
energy of γ in Fe-Mn binary system, Metall. Mater. Trans. A 31(2000) 355-360.
[19] K. Lu, Nanocrystalline metals crystallized from amorphous solids: nanocrystallization,
structure, and properties, Mater. Sci. Eng. R 16 (1996) 161-221.
[20] F.C. Zhang, B. Lv, T.S. Wang, C.L. Zheng, M. Li, M. Zhang, Microstructure in worn
surface of Hadfield steel crossing, Int. J. Mod. Phys. B 23 (2009) 1185-1190.
[21] Y.H Xu, Y.M. Chen, J.L. Xiong, J.H. Zhu, Mechanism of strain-induced
nanocrystallization of Hadfield steel under high energy impact load, Acta Metall. Sin. 37
(2001) 165-170.
[22] W. Yan, L. Fang, K. Sun, Y. Xu, Effect of surface work hardening on wear behavior of
Hadfield steel, Mater. Sci. Eng. A 460 (2007) 542-549.
[23] F.C. Zhang, Z.N. Yang, L.H. Qian, F.C. Liu, B. Lv, M. Zhang, High speed pounding: A
novel technique for the preparation of a thick surface layer with a hardness gradient
distribution on Hadfield steel, Scripta Mater. 64 (2011) 560-563.
[24] X.Y. Feng, F.C. Zhang, Z.N. Yang, M. Zhang, Wear behaviour of nanocrystallised
Hadfield steel, Wear 305 (2013) 299-304.
[25] P. Mukherjee, A. Sarkar, P. Barat, S.K. Bandyopadhyay, S. Pintu, S.K. Chattopadhyay,
Deformation characteristics of rolled zirconium alloys: a study by X-ray diffraction line
profile analysis, Acta Mater. 52 (2004) 5687-5696.
16
[26] L.L. Shaw, Y.T. Zhu. In: J. Groza, J. Shackelford, E. Lavernia, M. Powers, editors. CRC
materials processing handbook. Boca Raton, FL: CRC Press; 2007. 31.
[27] H. Hertz, In: D.E. Jones, G.A. Schott, editors. On the contact of elastic solids. London:
Macmillan; 1896.
[28] I. Gutierrez-Urrutia, D. Raabe, Dislocation and twin substructure evolution during strain
hardening of an Fe-22wt.% Mn-0.6wt.% C TWIP steel observed by electron channeling
contrast imaging, Acta Mater. 59 (2011) 6449-6462.
[29] S. Vercammen, B. Blanpain, B.C.D. Cooman. P. Wollants, Cold rolling behaviour of an
austenitic Fe-30Mn-3Al-3Si TWIP-steel: the importance of deformation twinning, Acta Mater.
52 (2004) 2005-2012.
[30] X.L. Wu, Y.T. Zhu, Y.G. Wei, Q. Wei, Strong Strain Hardening in Nanocrystalline Nickel,
Phys. Rev. Lett. 103 (2009) 205504.
[31] X. Li, Y. Wei, L. Lu, K. Lu, H. Gao, Dislocation nucleation governed softening and
maximum strength in nano-twinned metals, Nature 464 (2010) 877-880.
[32] M. Daoa, L. Lu, Y.F. Shen, S. Suresha, Strength, strain-rate sensitivity and ductility of
copper with nanoscale twins, Acta Mater. 54 (2006) 5421-5432.
[33] O. Bouaziz, N. Guelton, Modelling of TWIP effect on work-hardening, Mater. Sci. Eng.
A 319 (2001) 246-249.
[34] N.K. Kumar, B. Roy, J. Das, Effect of twin spacing, dislocation density and crystallite
size on the strength of nanostructured α-brass, J. Alloys Comp. 618 (2015) 139-145.
[35] I. Karaman, H. Sehitoglu, H.J. Maier, Y. I. Chumlyakov, Competing mechanisms and
modeling of deformation in austenitic stainless steel single crystals with and without nitrogen,
Acta mater. 49 (2001) 3919-3933.
[36] K. Lu, L. Lu, S. Suresh, Strengthening materials by engineering coherent internal
boundaries at the nanoscale, Science 324 (2009) 349-352.
[37] F.C. Zhang, X.Y. Feng, Z.N. Yang, J. Kang, T.S. Wang, Dislocation–twin boundary
interactions induced nanocrystalline via SPD processing in bulk metals, Sci. Rep. 5 (2015)
8981.
[38] C. Chen, B. Lv, F. Wang, F.C. Zhang, Low-cycle fatigue behaviors of pre-hardening
Hadfield steel, Mater. Sci. Eng. A 695 (2017) 144-153.
[39] M.S. Pham, C. Solenthaler, K.G.F. Janssens, S.R. Holdsworth, Dislocation structure
evolution and its effects on cyclic deformation response of AISI 316L stainless steel, Mater.
Sci. Eng. A 528(2011) 3261-3269.
[40] B. Hutchinson, N. Ridley, On dislocation accumulation and work hardening in Hadfield
steel, Scripta Mater. 55 (2006) 299-302.

17

You might also like