You are on page 1of 55

Materials Science & Engineering A 661 (2016) 115–125

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Strain-controlled low cycle fatigue properties of a rare-earth


containing ME20 magnesium alloy
F.A. Mirza a,n, K. Wang a, S.D. Bhole a, J. Friedman a, D.L. Chen a, D.R. Ni b, B.L. Xiao b,
Z.Y. Ma b,n
a
Department of Mechanical and Industrial Engineering, Ryerson University, 350 Victoria Street, Toronto, Ontario M5B 2K3, Canada
b
Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, 72 Wenhua Road, Shenyang 110016, China

art ic l e i nf o a b s t r a c t

Article history: The present study was aimed to evaluate the strain-controlled cyclic deformation characteristics and low
Received 10 January 2016 cycle fatigue (LCF) life of a low ( ! 0.3 wt%) Ce-containing ME20-H112 magnesium alloy. The alloy con-
Received in revised form tained equiaxed grains with ellipsoidal particles containing Mg and Ce (Mg12Ce), and exhibited a rela-
3 March 2016
tively weak basal texture. Unlike the high rare earth (RE)-containing magnesium alloy, the ME20M-H112
Accepted 3 March 2016
Available online 5 March 2016
alloy exhibited asymmetrical hysteresis loops somewhat similar to the RE-free extruded Mg alloys due to
the presence of twinning-detwinning activities during cyclic deformation. While cyclic stabilization was
Keywords: barely achieved even at the lower strain amplitudes, cyclic softening was the predominant characteristics
Magnesium alloy at most strain amplitudes. The ME20M-H112 alloy showed basically an equivalent fatigue life to that of
Rare-earth elements
the RE-free extruded Mg alloys, which could be described by the Coffin-Manson law and Basquin's
Cyclic deformation
equation. Fatigue crack was observed to initiate from the near-surface imperfections, and in contrast to
Twinning-detwinning
Texture weakening the typical fatigue striations, the present alloy showed some shallow dimples along with some fractions
of quasi-cleavage features in the crack propagation area.
& 2016 Elsevier B.V. All rights reserved.

1. Introduction There are currently intensive studies in the development of


Mg alloys with high strength, good corrosion resistance and su-
Due to the increasing global energy demand and great cogni- perior formability for structural applications [6,12,14,16–28].
zance of human-caused pollution such as CO2 emissions in recent Nevertheless, several restraints are faced in the application of Mg
years, lightweighting has turned into a crucial approach in the au- alloys, including limited ductility, tension-compression yield
tomotive and aerospace industries [1–10]. There has been major asymmetry, pronounced directional anisotropy arising from the
interest and great deal of research activity for finding alternatives to presence of strong crystallographic texture related to their hex-
reduce the fuel consumption of passenger vehicles [10,11]. Being agonal close-packed (HCP) structure due to a limited number of
the ultra-lightweight structural metallic material, magnesium (Mg) slip systems that could activated during extrusion or rolling
alloys have been increasingly applied in the auto industry for ve- processes [29–31]. For the vehicle components subjected to dy-
hicles weight reduction [12]. The application of Mg alloys as a namic cyclic loading, this mechanical anisotropy along with the
structural material in automotive and aerospace industry would tension-compression yield asymmetry could lead to irreversi-
require the evaluation of fatigue and cyclic deformation character- bility of cyclic deformation, which may have a serious influence
on the material performance [32]. Both the tension-compression
istics, since structural components would unavoidably experience
yield asymmetry and formability at room-temperature could be
dynamic loading in service, which leads to the occurrence of fatigue
effectively improved by the addition of alloying elements, espe-
failure [13–15]. Thus, to guarantee the structural integrity and
cially rare-earth (RE) elements due to their affinity to induce
durability of such engineering components it is essential to un-
texture randomization during hot deformation processes (e.g.,
derstand the fatigue and cyclic deformation behavior of Mg alloys.
extrusion or rolling), which results in the decrease of texture
As there is a growing application of Mg alloys in powertrain, chassis
intensities and the activation of basal slip [29,33–38]. Though
and body areas, developing wrought Mg components with im-
these alterations in the tension-compression yield asymmetry
proved mechanical properties became a prominent necessity [16]. and mechanical anisotropy due to RE elements additions are re-
cently reported, the potential advantage of such wrought RE-Mg
n
Corresponding authors. alloys as structural components under dynamic cyclic loading
E-mail addresses: f4mirza@ryerson.ca (F.A. Mirza), zyma@imr.ac.cn (Z.Y. Ma). conditions has not yet been well appreciated.

http://dx.doi.org/10.1016/j.msea.2016.03.024
0921-5093/& 2016 Elsevier B.V. All rights reserved.
116 F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125

Recently several studies on the fatigue of RE-Mg alloys have a grit number of 600 to remove the machining marks and to
been reported in the literature [29,39–47]. For instance, Yang et al. achieve a smooth surface.
[29] conducted very high cycle fatigue tests of a high RE-con- Tensile tests were performed in accordance with the ASTM: E8
taining extruded Mg-12Gd-3Y-0.5Zr alloy, and observed much standard by means of a computerized United tensile testing ma-
relieved tension-compression yield asymmetry and enhanced fa- chine with a gauge length of 25 mm at a strain rate of 1 # 10 $ 3 s $ 1
tigue failure resistance in comparison with RE-free AZ31 alloy. The at room temperature. Strain-controlled “pull-push” type fatigue
fatigue strength of extruded Mg-10Gd-1Nd and Mg-10Gd alloys in tests (in accordance with the ASTM: E606 standard) were con-
the form of S-N curves was also evaluated via stress-controlled ducted in air at room temperature with a 25 mm extensometer
high cycle fatigue tests [47]. A few strain-controlled low cycle fa- using a computerized Instron 8801 fatigue testing system that was
tigue tests on the RE-Mg alloys have been performed as well [48– controlled by a Fast Track Low Cycle Fatigue (LCF) program. The
55]. Although the high performance of the Mg–10Gd-3Y-0.5Zr cyclic deformation test conditions consisted of a zero mean strain
(GW103K) alloy was achieved [48–50], it was quite expensive due (i.e., a strain ratio of Rε ¼ $ 1, completely reversed strain cycle) and a
to the addition of a fairly high amount (totally ! 13 wt%) of RE constant strain rate of 1 # 10 $ 2 s $ 1 with triangular loading wave-
elements. Since the cost is one of the major considerations in the form, as noted in the ASTM: E606 standard for continuous cyclic
automotive applications, recently developed rolled ME20M-H112 tests and generally for strain-rate sensitive materials, since the tri-
alloy with a low amount of RE element ( ! 0.3 wt% Ce) would be a angular waveform results in a constant strain rate in the course of
promising candidate. To the authors’ knowledge, no studies have cycling. Low cycle fatigue tests were performed at total strain am-
been conducted on the cyclic deformation behavior of ME20M- plitudes of 0.2%, 0.4%, 0.6%, 0.8%, 1.0%, and 1.2% (at least two samples
H112 Mg alloy under strain control in the open literature. It is were tested at each level of the strain amplitudes). At lower strain
unclear to what extent the ME20M-H112 alloy would exhibit the amplitudes (e.g., 0.1% and 0.2%), strain-controlled tests were sus-
tension-compression asymmetry, whether cyclic hardening or tained for 10,000 cycles before being converted to load control, with
softening would occur, and what are the cyclic stress response and a sine cyclic waveform at a frequency of 50 Hz. Once tests were
fatigue life. Therefore, the objective of the present study was completed, SEM was used to examine the fracture surfaces of fati-
aimed at identifying cyclic deformation behavior of a low RE- gued samples, aiming to identify the various features involving fa-
containing rolled ME20M-H112 alloy, and evaluating the fatigue tigue initiation and propagation mechanisms. In addition, a special
life under varying strain amplitudes. interest was given to the near fracture surface areas of the fatigued
samples, which were cut, mounted, ground, polished and etched to
examine the eventual appearance of residual twins.
2. Material and experimental procedure

The material under investigation in the present study is a 3. Results and discussion
newly developed Mg alloy ME20 sheet, which was processed by
hot rolling and H112 tempered, whose nominal composition is 3.1. Initial microstructure and texture
provided in Table 1. Microstructural evolution was examined in
the samples by using an optical microscope (OM) equipped with Fig. 1 shows a typical optical micrograph and a backscattered
Clemex quantitative image analysis software, and scanning elec- electron (SEM) image of a rolled ME20M-H112 alloy, where the
tron microscope (SEM) JSM-6380LV equipped with Oxford energy arrow indicates the rolling direction (RD). The microstructure of
dispersive X-ray spectroscopy (EDS) system. Standard metallo- the alloy was composed of equiaxed grains with an average grain
graphic sample preparation techniques were used to grind and size ! 12 mm due to the dynamic recrystallization occurred in the
polish sample surfaces, and etching was done with an acetic picral rolling process [48]. The grain size was fairly small in comparison
solution containing 4.2-g picric acid, 10-ml acetic acid, 10-ml H2O, with the common extruded Mg alloys, such as AZ31 and AM30
and 70-ml ethanol. Textures were obtained using a PANalytical [14,16–19]. This was due to the role of added RE elements [48–50]
X-ray diffractometer (XRD) with Cu Kα radiation at 45 kV and and the grains can be prohibited from coarsening during the hot
40 mA in a back reflection mode by measuring partial pole figures deformation process as cerium (Ce) can form some dispersed and
(i.e., ranging between Ψ ¼0° and 75°). Texture data were after- thermally stable particles [57]. It was reported that a uniform re-
wards analyzed based on MTEX software [56]. It should be noted crystallized grain structure was observed with a grain size de-
that the defocusing stemming from the rotation of the XRD sample creasing from ! 13 to ! 10 mm as Ce content increased from 0% to
holder was corrected using the experimental data obtained from 1% for rolled and rolled/annealed ME alloys [58,59]. A similar type
Mg powder diffraction. The sample was positioned in the machine of microstructure of as-cast ME alloys has been reported in refs.
with the rolling direction parallel to the x-direction and defocusing [60–62]. The microstructure of the ME20 sample also revealed a
arising from the rotation of XRD sample holder was corrected few dispersed, fine, ellipsoidal particles as seen from Fig. 1(b)
using experimentally determined data obtained from the diffrac- (indicated using arrow). An EDS line scan was performed as shown
tion of Mg powders received from Magnesium Elektron. Sub-sized in Fig. 2(a) and (b), which confirmed the presence of RE particle
tensile and fatigue samples were machined with the loading axis (Ce) along with Mn. The very fine solid particles were identified as
parallel to the rolling direction (RD) in accordance with ASTM: E8 pure Mn, where Li et al. [61] also reported the existence of similar
standard. The samples had a gauge length of 25 mm (or a parallel ellipsoidal particles containing Mg and Ce (Mg12Ce). Similar find-
length of 32 mm), thickness of ! 7 mm, and a width of !6 mm. ings have been reported for ME20 alloys in Refs. [62–64].
The gauge section of tensile and fatigue samples was ground Fig.3 shows the crystallographic textures (basal (0001),
progressively along the loading direction with emery papers up to prismatic (1010¯ ), and pyramidal (1011 ¯ ) pole figures) of rolled

Table 1
Chemical composition of the rolled ME20M-H112 Mg alloy.

Element Al Zn Mn Ce Si Fe Cu Ni Be Mg

Content (wt%) o 0.2 o0.3 1.69–1.81 0.29–0.33 o 0.10 o 0.05 o0.05 o 0.007 o 0.01 Balance
F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125 117

RD
Fig. 1. Microstructure of ME20M-H112 alloy, (a) OM image and (b) SEM (backscattered electron) image, where RD stands for the rolling direction.

1600
Mg
1400

Intensity, Counts
1200
1000
800
600
400
200 Ce Mn
0
0 10 20 30 40 50 60 70 80 90
Distance , μm
Fig. 2. SEM backscattered electron images indicating an EDS line scan position and the corresponding EDS line scan results of a ME20M-H112 alloy.

ME20M-H112 alloy evaluated using MTEX software, where RD 3.2. Tensile properties
indicates the rolling direction, TD stands for transverse direction
and ND indicates the normal direction. A relatively weak con- Typical tensile stress-strain curve of the rolled ME20M-H112
centration (with a maximum intensity of 3.8 multiples of random alloy at a strain rate of 1 # 10 $ 3 s $ 1 is shown in Fig. 4 and the
distribution (MRD)) of basal (0001) pole mainly displays some tensile properties obtained are listed in Table 2. The present ME20
splitting towards the RD and TD, along with prismatic (1010¯ ) and alloy exhibited a flow curve with a gradual transition from elastic to
pyramidal (1011¯ ) poles towards the RD, in comparison with the plastic deformation like most of the RE-free extruded Mg alloys, i.e.,
extruded AM30 [65,66] and rolled AZ31 [67]. This indicates that AZ31, AM30 [16,17]. However, as seen from Fig. 4 and Table 2, a
the majority of hcp unit cells in the undeformed ME20 have their lower tensile yield strength (YS) was attained in the present ME20
c-axes oriented perpendicular to the rolling direction. The pre- alloy in comparison with the RE-free extruded Mg alloys [16,17], but
sence of such a weaker texture, in comparison with the extruded had an equivalent or slightly higher ductility. While one benefit of
AM30 [65,66] and rolled AZ31 [67], was another benefit of the RE the RE elements added into Mg alloys was to improve the tensile
properties, e.g., microalloying with RE elements exhibited a higher
elements added into Mg alloys. As also reported by Stanford and
tensile yield strength than the extruded AM30 alloy [48], however,
Barnett [68], a similar type of weaker texture was observed for an
it seemed that the tensile properties of the rolled ME20 alloy were
extruded ME10 alloy prior to tensile deformation and revealed a
not improved, in comparison with the reported results in other RE-
strong orientation peak which was termed as the RE texture
Mg alloys [29,48–50,67]. The fine grained ME20 should exhibit a
component [69]. Huppmann et al. [59] have also discovered two
higher yield strength than coarser RE-free extruded Mg alloys
additional RE texture components in an extruded ME21 Mg alloy
[16,48], which is probably related to the weaker textures with a
that were parallel to the extrusion direction: 1122¯ and 2021¯ .
broad basal pole scatter from RD (Fig. 3). Such textures are generally
The RE extrusion texture was associated with the presence of Ce in very easy to deform, for example, by basal slip at lower stresses
dynamic recrystallization (DRX) nucleation at shear bands of de- compared with sheet textures similar to standard sheet basal tex-
formed grains [69], which leads to randomize the texture and thus tures, where c-axis compression leads to a very high yield strength
enhance ductility [59]. Similar weaker textures in Mg-based alloys [60]. Similar results have also been reported for finer grains of ME20
with the addition of RE elements were reported in Refs. [29,32,48– sheets by Li et al. [60], where the tensile yield strength was less
50,61,62,63]. than 200 MPa as well.
118 F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125

RD

TD
ND

Basal plane Prismatic plane Pyramidal plane


(0001) (10 1 0 ) (10 11)

¯ ) plane, and pyramidal (1011


Fig. 3. Pole figures of basal (0001) plane, prismatic (1010 ¯ ) plane of the rolled ME20M alloy, where RD indicates the rolling direction, TD stands
for the transverse direction and ND indicates the normal direction.

current ME20 alloy was observed to exhibit unsymmetrical hys-


300 teresis loops, which were somewhat similar to those of RE-free
Engineering stress, MPa

extruded Mg alloys [16–18]. However, the extent of the asymmetry


250 or skewness of the hysteresis loops is less than that of RE-free
extruded Mg alloys [16–18]. This was, however, in contrast to the
hysteresis loops in the RE extruded Mg alloys [48–50], where the
200 hysteresis loops were nearly symmetrical, which were somewhat
similar to those of face-centered cubic (fcc) metals as a con-
sequence of the dislocation slip-dominated deformation in most
150 materials [70]. This unsymmetrical tensile and compressive
yielding phenomenon was distinguished as the Bauschinger-like
100 effect [16,17], and was associated with twinning in the compres-
sive phase and detwinning in the tensile phase [16,17]. It was re-
ported that low yield strength and strain hardening plateau in
50 compression is a direct characteristic of materials deformed by
twinning (e.g., in the descending phase of the first hysteresis loop
in Fig. 5), and the deformation by detwinning is responsible for the
0 low yield strength and strain hardening rate phenomena in the
0 5 10 15 20 25 subsequent ascending phase [71,72]. Twinning-induced plastic
strain could be completely recovered during detwinning, however,
Engineering strain, % due to its lower critical resolved shear stress (CRSS) detwinning
Fig. 4. Tensile stress-strain curve of the rolled ME20M-H112 alloy tested at a strain finished at the early stage of tensile loading [73]. Therefore, acti-
rate of 1 # 10 $ 3 s $ 1. vation of slip system is needed to accommodate the strain applied
in the subsequent stages of tensile loading, and this would trigger
a significant increase in tensile stress curve [74], as shown in Fig. 5
3.3. Hysteresis loops (the ascending phase of the hysteresis loop). Furthermore, unlike
the fcc metals where the slope of hysteresis loops after the strain
Typical stress-strain hysteresis loops of the first, second, and reversal either at the maximum or minimum stress was basically
mid-life cycles at a total strain amplitude of 1.2% and strain ratio of equal to the value of Young’s modulus, the pseudoelastic or non-
Rε ¼ $ 1 for the as-rolled ME20M-H112 alloy is shown in Fig. 5. The linear elastic behavior in both descending and ascending phases in
F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125 119

Table 2
Tensile properties of the rolled ME20M-H112 alloy, in comparison with those reported in the literature [16,17,48,50].

Tensile properties Yield strength, MPa Ultimate tensile strength, MPa Elongation, % Young's modulus, GPa Strain hardening exponent (n) K, MPa

Rolled ME20M-H112 140 234 20.0 42 0.19 390


Extruded AZ31 [16] 201 264 15.2 45 0.13 –
Extruded AM30 [17] 189 242 13.4 44 0.15 –
Extruded GW103 K [48,50] 232 318 8.3 – 0.15 –

200 200 1.2%


Total strain amplitudes
175 1.0%
150

Stress amplitude, MPa


0.8%
150 0.6%
100 0.4%
Stress, MPa

125 0.2%
50 0.1%
100
0
75
-50
50
-100
1st cycle 25
-150 2nd cycle 0
Mid-life cycle 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
-200
-2 -1 0 1 2 Number of cycles, N
Total strain, % Fig. 6. Stress amplitude vs. the number of cycles of the rolled ME20M-H112 alloy
tested at different total strain amplitudes at a strain ratio of Rε ¼ $ 1.
Fig. 5. Typical stress-strain hysteresis loops of different cycles of the rolled
ME20M-H112 alloy tested at a given total strain amplitude of 1.2% and strain ratio
of Rε ¼ $ 1.
0.8 Total strain amplitudes 1.2%
Plastic strain amplitude, %

the current ME20 alloy was also visible. It seemed reasonable to 1.0%
0.7
consider that the pseudoelastic behavior would be an inherent 0.8%
characteristic of Mg alloys, as it originated from reversible move- 0.6%
ment of dislocations, twinning, and stress induced phase trans-
0.6
0.4%
formation [16–18,75].
0.5 0.2%
0.1%
3.4. Cyclic stress and strain responses
0.4
The evolution of stress amplitudes and plastic amplitudes with
respect to the number of cycles at different strain amplitudes from
0.3
0.2% to 1.2% is shown in Figs. 6 and 7, respectively, under a semi-
logarithmic scale along the X-axis. It is observed that both stress
0.2
amplitudes and plastic strain amplitudes augmented, whereas
fatigue life of the rolled ME20M-H112 alloy decreased with in- 0.1
creasing total strain amplitudes. Unlike the extruded GW103K al-
loy [48–50] where microalloying with RE elements led to the 0
cyclic stabilization at almost all the strain amplitudes, it seemed 1E+0 1E+1 1E+2 1E+3 1E+4 1E+5
that cyclic stabilization was barely achieved in the rolled ME20M-
H112 alloy as seen from Fig. 6, except for 0.1% strain amplitude. At
Number of cycles, N
strain amplitudes of 1.2% and 1.0%, only slight change, i.e., an initial Fig. 7. Plastic strain amplitude vs. the number of cycles of the rolled ME20M-H112
slight cyclic hardening within the first three cycles and then minor alloy tested at different total strain amplitudes at a strain ratio of Rε ¼ $ 1.
cyclic softening up to failure, could be seen from Fig. 6. More or
less for all other strain amplitudes (0.2%, 0.4%, 0.6%, and 0.8%), In the low cycle fatigue tests, plastic strain amplitude has been
cyclic softening continues up to failure. In general, cyclic response
considered as a physical quantity that results in several damaging
(strain hardening [76] and strain softening [77]) is governed by the
processes and influences the internal microstructure which is
cyclic stability of the microstructural features, dislocation glide
closely related to the strain resistance and eventually the fatigue
and multiplication, twinning, and twin-dislocation interactions
[78]. The cyclic strain softening behavior for ME20 may be ex- life [16,17,48]. The variation of the plastic strain amplitude (Δεp/2)
plained by the twin softening mechanism; softening is caused by during cyclic deformation is shown in Fig. 7, which corresponded
the reorientation of lattice in the twinned area to provide an easier well to the change of the stress amplitude during cyclic de-
orientation for slip to happen [77,79]. formation as shown in Fig. 6 at different total strain amplitudes. It
120 F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125

1.4 -1
AM30 [1] Total
1.2 NZ30K [9] -1.5 Plastic
Log(∆εt/2) = -0.3276(2Nf) - 1.0335
AM30 [14] R² = 0.8757
Elastic
1.0 GW103K [29] -2

Log (∆ε/2)
AZ31 [56]
∆εt /2, %

0.8 AZ31 [75] -2.5 Log(∆εe/2) = -0.1512(2Nf) - 1.8008


AZ61-ED [80] R² = 0.7698
0.6 ME20-present study -3

0.4 -3.5 Log(∆εp/2) = -0.7636(2Nf) - 0.1269


R² = 0.8691

0.2 -4
1 2 3 4 5 6
0.0 Log(2Νf)
1E+1 1E+3 1E+5 1E+7
Fig. 9. Total, plastic, and elastic strain amplitude-fatigue life response of the rolled
Number of cycles to failure, Nf ME20M-H112 alloy.

Fig. 8. Total strain amplitude versus the number of cycles to failure for the rolled
ME20M-H112 alloy, in comparison with the data reported in the literature for independent of the elastic strain amplitude, which could be ex-
various wrought Mg alloys [1,9,14,29,56,75,80]. pressed by the following equation [17],

Δσ ⎛ Δεp ⎞n ′
= K ′⎜ ⎟ ,
2 ⎝ 2 ⎠ (2)
is seen from Fig. 7 that as the total strain amplitudes increased the
plastic strain amplitudes also increased however fatigue life de- where Δσ
is the mid-life stress amplitude,
Δεp
is the mid-life plastic
2 2
creased. Cyclic stabilization only occurred at the strain amplitude
strain amplitude, n′ is the cyclic strain-hardening exponent and K’
of 0.1%. For the rest of the strain amplitudes, it showed a trend of
is the cyclic strength coefficient. Fig.9 illustrates the elastic, plastic,
increasing with the number of cycles, which is another proof of
and total strain amplitudes as a function of the number of re-
cyclic softening, that occurred in all the tested samples except the
versals to failure (2Nf). In order to make sure that cyclic stabili-
one under strain amplitude of 0.1%.
zation, also called cyclic saturation, has already occurred, the
stress and strain values of the mid-life cycles were used. Hence,
3.5. Fatigue life and fatigue parameters
the fatigue life parameters obtained by means of (Eqs. (1) and 2)
were presented in Table 3, in comparison with the data reported in
Fig.8 displays the total strain amplitude (Δεt /2) as a function of
the number of cycles to failure (Nf, i.e., fatigue life) for the rolled the literature for various extruded Mg alloys [16,17,48,50]. It is
ME20M-H112 alloy, in comparison with the data reported in the seen that while the obtained fatigue parameters were well within
literature for both RE-free and RE-containing wrought Mg alloys the range in other Mg alloys reported in the literature
[1,9,14,29,56,75,80]. The run-out data points were specified by [16,17,48,50], the cyclic strain hardening exponent n’ of the
arrows pointing horizontally at or over 107 cycles. The present ME20M-H112 alloy was lower than that of RE-free Mg alloys
rolled ME20M-H112 alloy showed a trend of increasing fatigue life [16,17].
with decreasing strain amplitude and overall, the alloy basically Fig. 10 illustrates a superposition of the monotonic tensile
showed an improved fatigue life than the RE-free extruded Mg stress-strain curve at a strain rate of 1 # 10 $ 2 s $ 1 and cyclic stress-
alloys [18–22,69]. However, at lower stain amplitude of 0.2% the strain curve (CSSC) at mid-life cycles with the Y-axis stress in-
ME20 alloy showed the least fatigue life. This phenomenon is dicated as the normal stress amplitude ( Δσ ) for the ME20M-H112
2
closely related with material strength and ductility. The present Mg alloy tested at a strain ratio of Rε ¼ $ 1. It is seen that cyclic
rolled ME20-H112 alloys showed a higher ductility and lower stress-strain curve was positioned below the monotonic one of the
strength compared with other alloys [18–22,69] and this resulted ME20M-H112 alloy, suggesting the presence of cyclic softening
in better fatigue resistance at larger stain amplitudes. The total and being consistent with the response of stress and plastic strain
strain amplitude could be expressed as two parts of elastic strain amplitudes as shown in Figs. 6 and 7. The obtained cyclic yield
amplitude and plastic strain amplitude according to the Basquin strength (s‘y) and cyclic strain hardening exponent (n0 ) of the
equation and Coffin-Manson relation [1,6–9,14,17,19,23], i.e., present ME20M-H112 alloy were also slightly lower than the
corresponding monotonic yield strength and strain hardening
Δεt Δε Δεp σ f′(2Nf )b
= e + = + εf′(2Nf )c , exponent, as seen from Tables 2 and 3. This could be the reason
2 2 2 E (1)
why this alloy exhibited a certain extent of cyclic softening even at
where E is the Young's modulus (for the present alloy the average higher strain amplitude of 1.2%. In addition, as the monotonic
value obtained during fatigue testing was ! 42 GPa), Nf is the fa- hardening exponent n was previously obtained to be about 0.19
tigue life or the number of cycles to failure, σ′f is the fatigue (Table 2), whereas the cyclic strain hardening exponent (n0 ) was
strength coefficient, b is the fatigue strength exponent, ε′f is the about 0.11 which was lower than n. The fact that the normal CSSC
fatigue ductility coefficient, and c is the fatigue ductility exponent. was located below the monotonic stress-strain curve (Fig.10) di-
In addition, cyclic deformation behavior is normally considered to rectly reflected a slight cyclic softening at total strain amplitudes
be related to the portion of the plastic strain amplitude and is from 0.4% to 1.2%, as seen from Fig.10. This is in contrast to the
F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125 121

Table 3
Low cycle fatigue parameters obtained for the rolled ME20M-H112 alloy, in comparison with those reported in the literature [16,17,48,50].

Low cycle fatigue parameters Rolled ME20M-H112 Extruded AZ31 [16] Extruded AM30 [17] Extruded GW103 K [48,50]

Cyclic yield strength, s‘y,MPa 115 – – 215


Cyclic strain hardening exponent, n′ 0.11 0.34 0.33 0.19
Cyclic strength coefficient, K’, MPa 240 1976 1610 734
Fatigue strength coefficient, s‘f, MPa 665 616 678 518
Fatigue strength exponent, b $ 0.15 $ 0.15 $ 0.16 $ 0.11
Fatigue ductility coefficient, ε‘f 0.75 1.78 2.97 0.05
Fatigue ductility exponent, c $ 0.76 $ 0.40 $ 0.44 $ 0.44

seen from the low magnification images (Fig. 12) that fatigue crack
200 initiated basically from the specimen surface. It was observed
clearly that fatigue crack propagation area was larger at a lower
strain amplitude of 0.2% than at a higher strain amplitude of 1.0%
∆σ/2 at mid-life, MPa

150 (Fig. 11(a) and (b)). In contrast to typical fatigue striations


[14,16,17,19,34,48–51], the present alloy showed mainly some
shallow dimples along with some fractions of quasi-cleavage fea-
tures in the crack propagation area as shown in Fig.13(a) and (b) at
100 a higher magnification.

3.7. Comparison between different Mg alloys

50 Since the present rolled ME20M-H112 alloy exhibited a certain


Monotonic stress-strain curve extent of asymmetrical stress-strain behavior in tension and
compression (Fig. 5), which are of major concern in relation to the
Cyclic stress-strain curve
influence of RE element addition as well as the twinning-det-
0 winning behavior, it is thus necessary to identify the difference
0 0.2 0.4 0.6 0.8 1 1.2 1.4 from RE-Mg and RE-free Mg alloys based on the stress-strain
∆εt /2 at mid-life, % hysteresis loops. Fig. 14 illustrates the stress-strain hysteresis loops
of the mid-life cycle for a RE-free AM30 Mg alloy and a RE-con-
Fig. 10. Cyclic stress-strain curve for the rolled ME20M-H112 alloy, where the taining GW103K Mg alloy, along with that of the present RE-con-
corresponding monotonic stress-strain curve is also potted for comparison. taining rolled ME20M-H112 alloy. It is seen from Fig. 14 that the
hysteresis loop of ME20M-H112 alloy showed a strong Bauschin-
situation of a RE-free extruded AZ31 Mg alloy which exhibited ger-like effect associated with a skewed asymmetrical shape si-
strong cyclic hardening, where the values of 0.13 and 0.34 for n milar to the AM30 alloy, which was predominantly associated with
and n0 , respectively, were reported [16]. Therefore, it could be the twinning-detwinning process [14,16–19,71]. The similarity in
concluded that the present ME20M-H112 alloy was susceptible to hysteresis loops between the ME20M-H112 alloy and AM30 alloy
cyclic softening under cyclic loading conditions rather than the is directly related to the presence of crystallographic texture which
monotonic ones. dominates the orientation of slip or twinning planes and direc-
tions relative to the externally applied stress [81]. As mentioned
3.6. Fractography earlier in Section 3.1, the present ME20M-H112 alloy showed re-
latively weak textures (3.8 MRD) (Fig. 3), which lay in-between the
Fig.11 showed the overall view of the fracture surfaces of the textures of both comparison alloys (e.g., AM30 and GW103K, ! 7,
rolled samples at a total strain amplitude of (a) 0.2% and (b) 1.0%, and 2.1 MRD, respectively). In this context, Fig. 14 actually helped
respectively, containing fatigue crack initiation (as indicated by emphasize the conclusions that the stress-strain behavior of the
dashed yellow-box), propagation, final fast fracture regions. It is present ME20M-H112 alloy exhibit a characteristic lying in-

Fig. 11. SEM images of overall fracture surfaces of the rolled ME20M-H112 specimens fatigued at a total strain amplitude of (a) 0.2% and (b) 1.0%.
122 F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125

Fig. 12. SEM micrographs of the fracture surface near crack initiation of the rolled ME20M-H112 specimens fatigued at a total strain amplitude of (a) 0.2% and (b) 1.0%.

Fig.13. SEM micrographs of the fatigue crack propagation region of the rolled specimens fatigued at total strain amplitude of (a) 0.2% and (b) 1.0%.

400 (Fig.14), which was also in-between both comparison alloys (e.g.,
Mid-life cycle AM30 and GW103K, ! 0.55 and ! 1, respectively). This indicates
300 that yield asymmetry is not very prominent as compared with
other wrought Mg alloy, and this is probably due to the light ad-
200 dition of RE element addition. While there was almost no yield
asymmetry for GW103K during cyclic loading as reported in ref.
Stress, MPa

[48], the RE content is over 10 wt% for this alloy as compared with
100
merely 0.29–0.33 wt% for ME20. Fig.15 shows the distribution of
deformation features in the area near the fracture surface in the
0
fatigued samples of the rolled ME20M-H112 alloy, along with
AM30 alloy and GW103K alloy. It is seen that the grain size of the
-100
ME20M-H112 alloy was significantly smaller than that of AM30
alloy and even GW103K alloy (Fig.15). While the grain size was
-200 ME20
reported to have a significant effect on the tendency of twinning
GW103K [80], it is in contrast to the usual hypothesis in the case of ME20M-
-300
AM30 H112 alloy. The area near the fracture surface of ME20M-H112
alloy exhibited lots of small twins. Thus, one of the obvious factors
-400
responsible for the asymmetry of hysteresis loops (Figs. 5 and 14)
-1.5 -1 -0.5 0 0.5 1 1.5 is the presence of abundant deformation twins, even though grain
Strain, % refinement was achieved due to the RE-element and process
control. The combined role of the deformation twins and ellip-
Fig. 14. Typical stress-strain hysteresis loops of the mid-life cycle at a given total
soidal particles containing Mg and Ce (Mg12Ce) as reported by Li
strain amplitude of 1.2% and strain ratio of Rε ¼ $ 1 for the rolled ME20, GW103 K,
and AM30 alloys, respectively. et al. [61] would be responsible for the asymmetric hysteresis loop
(Figs. 5 and 14). The above results and analyses indicate that even
between both comparison alloys. However, this is in contrast with the presence of RE-particles and the grain refinement in the pre-
the typically cast Mg alloys, which have no such yield asymmetry sent ME20M-H112 alloy, the decisive role in suppressing the oc-
because of their weak or no texture [68]. currence of deformation twins was absent, which led to the ten-
In addition, the ratio of the compressive-to-tensile yield stress sion-compression asymmetry almost similar to AM30 alloy
was ! 0.67 for ME20 alloy in the mid-life cycle of hysteresis loop (Fig. 5 and Fig. 14).
F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125 123

Fig. 15. Optical micrographs in the area near the fracture surface at a total strain amplitude of 1.2%, showing the distribution of deformation features in the fatigued samples
of (a) rolled ME20M-H112 alloy, (b) GW103K alloy, and (c) AM30 alloy, respectively.

4. Conclusions shallow dimples along with some fractions of quasi-cleavage


features in the crack propagation area.
Strain-controlled low cycle fatigue tests were conducted on a
low rare-earth containing rolled ME20M-H112 alloy under varying
strain amplitudes. The following conclusions could be drawn from Acknowledgements
this investigation:
The authors would like to thank the Natural Sciences and En-
1. The microstructure of the ME20M-H112 alloys consisted of gineering Research Council of Canada (NSERC) and AUTO21 Net-
equiaxed grains with an average grain size ! 12 mm arising from work of Centres of Excellence for providing financial support. D.R.
the dynamic recrystallization occurred in the rolling process. Ni would like to give thanks to National Natural Science Founda-
Unlike the RE-free extruded magnesium alloys, the present alloy tion of China for the financial support (Grant no. 51371179). D.L.
exhibited a relatively weak basal texture with a maximum in- Chen is grateful for the financial support by the Premier's Research
tensity of 3.8 MRD. Excellence Award (PREA), NSERC-Discovery Accelerator Supple-
2. Unlike the high RE-containing GW103K alloy, the rolled ment (DAS) Award, Canada Foundation for Innovation (CFI), and
ME20M-H112 alloy exhibited asymmetrical hysteresis loops si- Ryerson Research Chair (RRC) program. The authors would like to
milar to the RE-free extruded Mg alloys. This was pre- thank Dr. R. Tandon and Dr. B. Davies (Magnesium Elektron) for
dominantly due to the twinning-detwinning activities during supplying magnesium powders for the defocusing calibration. The
cyclic deformation. authors would also like to thank Messrs. A. Machin, Q. Li, J.
3. The cyclic stabilization was barely achieved even at the lower Amankrah and R. Churaman for easy access to the laboratory
strain amplitudes. At higher strain amplitude, an initial cyclic facilities of Ryerson University and their assistance in the
hardening occurred within a few cycles, followed by cyclic experiments.
softening.
4. The fatigue life of the ME20M-H112 alloy was equivalent to that
of other RE-free extruded Mg alloys within the experimental References
scatter, which could be well described by the Coffin-Manson
law and Basquin’s equation. [1] M. McNutt, The beyond-two-degree inferno, Science 349 (6243) (2015) 7.
5. SEM examinations revealed that fatigue cracks initiated from [2] T.M. Pollock, Weight loss with magnesium alloys, Science 328 (2010) 986–987.
[3] J.F. Nie, Y.M. Zhu, J.Z. Liu, X.Y. Fang, Periodic segregation of solute atoms in fully
the specimen surface or near-surface defects. In contrast to the coherent twin boundaries, Science 340 (2013) 957–960.
typical fatigue striations, the present alloy showed some [4] S. Chu, A. Majumdar, Opportunities and challenges for a sustainable energy
124 F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125

future, Nature 488 (2012) 294–303. Mg‐rare earth alloys, Metall. Mater. Trans. A 43 (4) (2012) 1347–1362.
[5] F. Creutzig, P. Jochem, O.Y. Edelenbosch, L. Mattauch, D.P. van Vuuren, [39] J. Dong, W.C. Liu, X. Song, P. Zhang, W.J. Ding, A.M. Korsunsky, Influence of
D. McCollum, J. Minx, Transport: a roadblock to climate change mitigation? heat treatment on fatigue behavior of high-strength Mg-10Gd-3Y alloy, Mater.
Science 350 (2015) 911–912. Sci. Eng. A 725 (2010) 6053–6063.
[6] M. McNutt, Climate change impacts, Science 341 (2013) 435. [40] Z.M. Li, Q.G. Wang, A.A. Luo, P.H. Fu, L.M. Peng, Y.G. Wang, High cycle fatigue of
[7] C. Ash, E. Culotta, J. Fahrenkamp-Uppenbrink, D. Malakoff, J. Smith, A. Sugden, cast Mg-3Nd-0.2Zn magnesium alloys, Metall. Mater. Trans. A 44A (2013)
S. Vignieri, Once and future climate change, Science 341 (2013) 473. 5202–5215.
[8] S.J. Davis, K. Caldeira, H.D. Matthews, Future CO2 emissions and climate [41] Y. Yang, Y.B. Liu, S.Y. Qin, Y. Fang, High cycle fatigue properties of die-cast
change from existing energy infrastructure, Science 329 (2010) 1330–1333. magnesium alloy AZ91D with addition of different concentrations of cerium, J.
[9] J.D. Shakun, P.U. Clark, F. He, S.A. Marcott, A.C. Mix, Z.Y. Liu, B. Otto-Bliesner, Rare Earths 24 (2006) 591–595.
A. Schmittner, E. Bard, Global warming preceded by increasing carbon dioxide [42] Y. Yang, Y.B. Liu, High cycle fatigue characterization of two die-cast magne-
concentrations during the last deglaciation, Nature 484 (2012) 49–54. sium alloys, Mater. Charact. 59 (2008) 567–570.
[10] T.A. Schaedler, A.J. Jacobsen, W.B. Carter, Toward lighter, stiffer material, Sci- [43] H. Bayani, E. Saebnoori, Effect of rare earth elements addition on thermal fa-
ence 341 (2013) 1181–1182. tigue behaviors of AZ31 magnesium alloy, J. Rare Earths 27 (2009) 255–258.
[11] M. Anderson, M. Auffhammer, Pounds that kill: the external costs of vehicle [44] L. Nascimento, S.B. Yi, J. Bohlen, L. Fuskova, D. Letzig, K.U. Kainer, High cycle
weight, Rev. Econ. Stud. 81 (2014) 535–571. fatigue behaviour of magnesium alloys, Procedia Eng. 2 (2010) 743–750.
[12] A.A. Luo, Magnesium: current and potential automotive applications, JOM 54 [45] Y.L. Xu, K. Zhang, X.G. Li, J. Lei, Y.S. Yang, T.J. Luo, High cycle fatigue properties
(2002) 42–48. of die-cast magnesium alloy AZ91D-1%MM, Trans, Nonfer. Met. Soc. China 18
[13] V.V. Ogarevic, R.I. Stephens, Fatigue of magnesium alloys, Annu. Rev. Mater. (2008) 306–311.
Sci. 20 (1990) 141–177. [46] W.-C. Liu, J. Dong, P. Zhang, L. Jin, T. Peng, C.Q. Zhai, W.-J. Ding, Fatigue be-
[14] X.Z. Lin, D.L. Chen, Strain controlled cyclic deformation behavior of an ex- havior of hot extruded Mg-10Gd-3Y magnesium alloy, J. Mater. Res. 25 (4)
truded magnesium alloy, Mater. Sci. Eng. A 496 (2008) 106–113. (2010) 773–783.
[15] F.A. Mirza, D.L. Chen, Fatigue of lightweight magnesium alloys, in: S. Zhang, D. [47] P. Maier, G. Tober, C.L. Mendis, S. Müller, N. Hort, Influence of Nd in extruded
L. Zhao (Eds.), Aerospace Materials Handbook, CRC Press, Taylor & Francis, New Mg-10Gd base alloys on fatigue strength, Adv. Mater. Res. 783 (2014) 419–424.
York, 2013, pp. 647–698. [48] F.A. Mirza, D.L. Chen, D.J. Li, X.Q. Zeng, Low cycle fatigue of a rare-earth con-
[16] S. Begum, D.L. Chen, S. Xu, A.A. Luo, Low cycle fatigue properties of an ex- taining extruded magnesium alloy, Mater. Sci. Eng. A 575 (2013) 65–73.
truded AZ31 magnesium alloy, Int. J. Fatigue 31 (2009) 726–735. [49] F.A. Mirza, D.L. Chen, D.J. Li, X.Q. Zeng, Effect of strain ratio on cyclic de-
[17] S. Begum, D.L. Chen, S. Xu, A.A. Luo, Strain-controlled low-cycle fatigue formation behavior of a rare-earth containing extruded magnesium alloy,
properties of a newly developed extruded magnesium alloy, Metall. Mater. Mater. Sci. Eng. A 5 (2013) 250–259.
Trans. A 39 (2008) 3014–3026. [50] F.A. Mirza, D.L. Chen, D.J. Li, X.Q. Zeng, Cyclic deformation behavior of a rare-
[18] S. Begum, D.L. Chen, S. Xu, A.A. Luo, Effect of strain ratio and strain rate on low earth containing extruded magnesium alloy: effect of heat treatment, Metall.
cycle fatigue behavior of AZ31 wrought magnesium alloy, Mater. Sci. Eng. A Mater. Trans. A 46 (3) (2015) 1168–1187.
517 (2009) 334–343. [51] F.A. Mirza, D.L. Chen, D.J. Li, X.Q. Zeng, Cyclic deformation of rare-earth con-
[19] C.L. Fan, D.L. Chen, A.A. Luo, Dependence of the distribution of deformation taining magnesium alloys, Adv. Mater. Res. 891 (2014) 391–396.
twins on strain amplitudes in an extruded magnesium alloy after cyclic de- [52] F.H. Wang, J. Dong, Y.Y. Jiang, W.J. Ding, Cyclic deformation and fatigue of
formation, Mater. Sci. Eng. A 519 (2009) 38–45. extruded Mg-Gd-Y magnesium alloy, Mater. Sci. Eng. A 561 (2013) 403–410.
[20] X.B. Liu, R.S. Chen, E.H. Han, Effects of ageing treatment on microstructures [53] F.H. Wang, J. Dong, M.L. Feng, J. Sun, W.J. Ding, Y.Y. Jiang, A study of fatigue
and properties of Mg-Gd-Y-Zr alloys with and without Zn additions, J. Alloys damage development in extruded Mg-Gd-Y magnesium alloy, Mater. Sci. Eng.
Compd. 465 (2008) 232–238. A 589 (2014) 209–216.
[21] K.U. Kainer, Magnesium - Alloys and Technology, Wiley-VCH, Cambridge, [54] F. Mokdad, D.L. Chen, Strain-controlled low cycle fatigue properties of a rare-
2003. earth containing ZEK100 magnesium alloy, Mater. Des. 67 (2015) 436–447.
[22] N. Tahreen, D.L. Chen, M. Nouri, D.Y. Li, Influence of aluminum content on [55] F. Mokdad, D.L. Chen, Cyclic deformation and anelastic behavior of ZEK100
twinning and texture development of cast mg-al-Zn alloy during compression, magnesium alloy: effect of strain ratio, Mater. Sci. Eng. A 640 (2015) 243–258.
J. Alloys Compd. 623 (2015) 15–23. [56] F. Bachmann, R. Hielscher, H. Schaeben, Texture analysis with MTEX - free and
[23] N. Tahreen, D.F. Zhang, F.S. Pan, X.Q. Jiang, C. Li, D.Y. Li, D.L. Chen, Character- open source software toolbox, Solid State Phen. 160 (2010) 63–68.
ization of hot deformation behavior of an extruded Mg‐Zn‐Mn‐Y alloy con- [57] K. Yu, S.T. Rui, J.M. Song, W.X. Li, L. Guo, Effects of grain refinement on me-
taining LPSO phase, J. Alloys Compd. 644 (2015) 814–823. chanical properties and microstructures of AZ31 alloy, Trans, Nonfer. Met. Soc.
[24] N. Tahreen, D.L. Chen, M. Nouri, D.Y. Li, Effects of aluminum content and strain China 18 (2008) s39–s43.
rate on strain hardening behavior of cast magnesium alloys during compres- [58] M. Masoumi, M. Hoseini, M. Pekguleryuz, The influence of Ce on the micro-
sion, Mater. Sci. Eng. A 594 (2014) 235–245. structure and rolling texture of Mg-1%Mn alloy, Mater. Sci. Eng. A 528 (2011)
[25] N. Tahreen, D.F. Zhang, F.S. Pan, X.Q. Jiang, C. Li, D.Y. Li, D.L. Chen, Influence of 3122–3129.
yttrium content on phase formation and strain hardening behavior of Mg-Zn- [59] M. Huppmanna, S. Gall, S. Müller, W. Reimers, Changes of the texture and the
Mn magnesium alloy, J. Alloys Compd. 615 (2014) 424–432. mechanical properties of the extruded Mg alloy ME21 as a function of the
[26] W.Q. Xu, N. Birbilis, G. Sha, Y. Wang, J.E. Daniels, Y. Xiao, M. Ferry, A high- process parameters, Mater. Sci. Eng. A 528 (2010) 342–354.
specific-strength and corrosion-resistant magnesium alloy, Nature Mater 14 [60] X. Li, T. Al-Samman, G. Gottstein, Microstructure development and texture
(2015) 1229–1235. evolution of ME20 sheets processed by accumulative roll bonding, Mater. Lett.
[27] G.S. Frankel, Magnesium alloys: ready for the road, Nature Mater 14 (2015) 65 (2011) 1907–1910.
1189–1190. [61] X. Li, T. Al-Samman, S. Mu, G. Gottstein, Texture and microstructure devel-
[28] L.Y. Chen, J.Q. Xu, H. Choi, M. Pozuelo, X.L. Ma, S. Bhowmick, J.M. Yang, opment during hot deformation of ME20 magnesium alloy: experiments and
S. Mathaudhu, X.C. Li, Processing and properties of magnesium containing a simulations, Mater. Sci. Eng. A 528 (2011) 7915–7925.
dense uniform dispersion of nanoparticles, Nature 528 (2015) 539–543. [62] X. Li, T. Al-Samman, G. Gottstein, Mechanical properties and anisotropy of
[29] F. Yang, F. Lv, X.M. Yang, S.X. Li, Z.F. Zhang, Q.D. Wang, Enhanced very high ME20 magnesium sheet produced by unidirectional and cross rolling, Mater.
cycle fatigue performance of extruded Mg-12Gd-3Y-0.5Zr magnesium alloy, Des. 32 (2011) 4385–4393.
Mater. Sci. Eng. A 528 (2011) 2231–2238. [63] Q.S. Yang, B. Jiang, X. Li, H.W. Dong, W.J. Liu, F.S. Pan, Microstructure and
[30] S.M. Yin, H.J. Yang, S.X. Li, S.D. Wu, F. Yang, Cyclic deformation behavior of as- mechanical behavior of the Mg‐Mn‐Ce magnesium alloy sheets, J. Magnesium
extruded Mg-3%Al-1%Zn, Scr. Mater. 58 (2008) 751–754. Alloys 2 (2014) 8–12.
[31] E.A. Ball, P.B. Prangnell, Tensile-compressive yield asymmetries in high [64] X. Li, W. Qi, Effect of initial texture on texture and microstructure evolution of
strength wrought magnesium alloys, Scr. Metall. Mater. 31 (1994) 111–116. ME20 Mg alloy subjected to hot rolling, Mater. Sci. Eng. A 560 (2013) 321–331.
[32] J. Bohlen, M.R. Nurnberg, J.W. Senn, D. Letzig, S.R. Agnew, The texture and [65] D. Sarker, D.L. Chen, Detwinning and strain hardening of an extruded mag-
anisotropy of magnesium-zinc-rare earth alloy sheets, Acta Mater. 55 (2007) nesium alloy during compression, Scr. Mater. 67 (2012) 165–168.
2101–2112. [66] D. Sarker, D.L. Chen, Texture transformation in an extruded magnesium alloy
[33] J.B. Jordon, J.B. Gibson, M.F. Horstemeyer, H. El-Kadiri, J.C. Baird, A.A. Luo, Ef- under pressure, Mater. Sci. Eng. A 582 (2013) 63–67.
fect of twinning, slip, and inclusions on the fatigue anisotropy of extrusion- [67] F. Lv, F. Yang, Q.Q. Duan, Y.S. Yang, S.D. Wu, S.X. Li, Z.F. Zhang, Fatigue prop-
textured AZ61 magnesium alloy, Mater. Sci. Eng. A 528 (2011) 6860–6871. erties of rolled magnesium alloy (AZ31) sheet: influence of specimen or-
[34] F.A. Mirza, D.L. Chen, Fatigue of rare-earth containing magnesium alloys: a ientation, Int. J. Fatigue 33 (2011) 672–682.
review, Fatigue Fract. Eng. Mater. 37 (2014) 831–853. [68] N. Stanford, M. Barnett, Effect of composition on the texture and deformation
[35] T.A. Samman, X. Li, Sheet texture modification in magnesium-based alloys by behaviour of wrought Mg alloys, Scr. Mater. 58 (2008) 179–182.
selective rare earth alloying, Mater. Sci. Eng. A 528 (2011) 3809–3822. [69] J. Hirsch, T. Al-Samman, Superior light metals by texture engineering: opti-
[36] J. Bohlen, S.B. Yi, D. Letzig, K.U. Kainer, Effect of rare earth elements on the mized aluminum and magnesium alloys for automotive applications, Acta
microstructure and texture development in magnesium-manganese alloys Mater. 61 (2013) 818–843.
during extrusion, Mater. Sci. Eng. A 527 (2010) 7092–7098. [70] S. Suresh, Fatigue of Materials, second ed, Cambridge University Press, Cam-
[37] R. Cottam, J. Robson, G. Lorimer, B. Davis, Dynamic recrystallization of Mg and bridge, 1998.
Mg-Y alloys: crystallographic texture development, Mater. Sci. Eng. A 485 [71] L. Wu, A. Jain, D.W. Brown, G.M. Stoica, S.R. Agnew, B. Clausen, D.E. Fielden, P.
(2008) 375–382. K. Liaw, Twinning-detwinning behavior during the strain-controlled low-cycle
[38] J.P. Hadorn, K. Hantzsche, S.B. Yi, J. Bohlen, D. Letzig, J.A. Wollmershauser, S. fatigue testing of a wrought magnesium alloy, Acta Mater. 56 (2008) 688–695.
R. Agnew, Role of solute in the texture modification during hot deformation of [72] D. Sarker, J. Friedman, D.L. Chen, De-twinning and texture change in an
F.A. Mirza et al. / Materials Science & Engineering A 661 (2016) 115–125 125

extruded AM30 magnesium alloy during compression along normal direction, Quart. 44 (2) (2005) 155–166.
J. Mater. Sci. Tech. 31 (2015) 264–268. [78] F.A. Mirza, D.L. Chen, D.J. Li, X.Q. Zeng, Low cycle fatigue of extruded Mg-3Nd-
[73] S.G. Hong, S.H. Park, C.S. Lee, Enhancing the fatigue property of rolled AZ31 0.2Zn-0.5Zr (wt%) magnesium alloy, Mater. Des. 64 (2014) 63–73.
magnesium alloy by controlling {10–12} twinning-detwinning characteristics, [79] S.R. Kalidindi, A.A. Salem, R.D. Doherty, Role of deformation twinning on
J. Mater. Res. 25 (2010) 784–792. strain hardening in cubic and hexagonal polycrystalline metals, Adv. Eng.
[74] S.H. Park, S.-G. Hong, B.H. Lee, W.K. Bang, C.S. Lee, Low-cycle fatigue char- Mater. 5 (2003) 229–232.
acteristics of rolled Mg-3Al-1Zn alloy, Int. J. Fatigue 32 (2010) 1835–1842. [80] J. Koike, T. Kobayashi, T. Mukai, H. Watanabe, M. Suzuki, K. Maruyama,
[75] C.H. Cáceres, T. Sumitomo, M. Veidt, Pseudoelastic behaviour of cast magne- K. Higashi, The activity of non-basal slip systems and dynamic recovery at
sium AZ91 alloy under cyclic loading-unloading, Acta Mater. 51 (20) (2003) room temperature in fine-grained AZ31B magnesium alloys, Acta Mater. 51
6211–6218. (2003) 2055–2065.
[76] M.R. Barnett, Twinning and ductility of magnesium alloys: (part I) “tension” [81] J. Jain, W.J. Poole, C.W. Sinclair, M.A. Gharghouri, Reducing the tension-com-
twins, Mater. Sci. Eng. A 464 (2007) 1–7.
pression yield asymmetry in a Mg-8Al-0.5Zn alloy via precipitation, Scr. Mater.
[77] S. Xu, V.Y. Gertsman, J. Li, J.P. Thompson, M. Sahoo, Role of mechanical twin-
62 (2010) 301–304.
ning in tensile compressive yield asymmetry of die cast Mg alloys, Can. Metall.
JOM, Vol. 68, No. 5, 2016
DOI: 10.1007/s11837-015-1796-7
Ó 2016 The Minerals, Metals & Materials Society

Microstructure and Fatigue Properties of Ultrasonic Spot


Welded Joints of Aluminum 5754 Alloy

F.A. MIRZA ,1,2 A. MACWAN,1 S.D. BHOLE,1 and D.L. CHEN1,3

1.—Department of Mechanical and Industrial Engineering, Ryerson University, 350 Victoria Street,
Toronto, ON M5B 2K3, Canada. 2.—e-mail: f4mirza@ryerson.ca. 3.—e-mail: dchen@ryerson.ca

The purpose of this investigation was to evaluate the microstructural change,


lap shear tensile load, and fatigue resistance of ultrasonic spot welded joints of
aluminum 5754 alloy for automotive applications. A unique ‘‘necklace’’-type
structure with very fine equiaxed grains was observed to form along the weld
line due to the mechanical interlocking coupled with the occurrence of dy-
namic recrystallization. The maximum lap shear tensile strength of 85 MPa
and the fatigue limit of about 0.5 kN (at 1 9 107 cycles) were achieved. The
tensile fracture occurred at the Al/Al interface in the case of lower energy
inputs, and at the edge of nugget zone in the case of higher energy inputs. The
maximum cyclic stress for the transition of fatigue fracture mode from the
transverse through-thickness crack growth to the interfacial failure increased
with increasing energy input. Fatigue crack propagation was mainly charac-
terized by the formation of fatigue striations, which usually appeared per-
pendicular to the fatigue crack propagation.

to expand the use of Al alloys in automotive body


INTRODUCTION
manufacturing, lower-cost joining methods are
Light-weighting in ground vehicles is deemed as important.5 One of the predominant processes of
one of the most effective strategies to improve fuel welding Al alloys is resistance spot welding
economy and reduce anthropogenic climate-chang- (RSW),5,6 however it has a limited application to
ing, environment-damaging, costly and human Al welding due to high energy consumption.6 More-
death-causing* emissions.1,2 To manufacture over, the major metallurgical problem of joining
lighter-weight vehicles, advanced high-strength dissimilar metals is the formation of brittle inter-
steels, aluminum (Al) alloys, magnesium alloys, metallic compounds (IMCs), which can seriously
and polymers are being used in the aerospace and degrade a joint’s mechanical properties.3 Nowadays,
automotive sectors. The development and applica- two new solid state welding processes, i.e., friction
tion of Al alloys have been significantly increasing stir spot welding (FSSW)7–9 and ultrasonic spot
in the transportation industry because of their high welding (USW),3,10,11 are being considered as poten-
strength-to-weight ratio, good machinability, envi- tial replacement technologies for RSW.5,7 On the
ronmental friendliness and recyclability.3–5 In order other hand, USW uses localized high-frequency
vibratory energy and moderate clamping forces
during joining and demonstrated the ability to
achieve superior strengths than FSSW when com-
*According to Science News entitled ‘‘Air pollution kills 7 million pared on the basis of the alloy thickness and
people a year’’ on March 25, 2014 at http://news.sciencemag.org/ processing condition (e.g., same nugget
signal–noise/2014/03/air-pollution-kills-7-million-people-year: ‘‘Air area).3–5,12–14 From an energy consumption point
pollution isn’t just harming Earth; it’s hurting us, too. Startling of view, Al welding using the USW process con-
new numbers released by the World Health Organization today
sumes only about 0.3 kWh per 1000 joints, com-
reveal that one in eight deaths are a result of exposure to air
pollution. The data reveal a strong link between the tiny particles
pared to 20 kWh with RSW, and 2 kWh with
that we breathe into our lungs and the illnesses they can lead to, FSSW.5 It has also been reported that ultrasonic
including stroke, heart attack, lung cancer, and chronic obstruc- spot welded (USWed) joints have higher failure
tive pulmonary disease.’’ energy than that of FSSWed joints.9,15

(Published online January 11, 2016) 1465


1466 Mirza, Macwan, Bhole, and Chen

In higher-power USW of thicker gauge sheet joining mechanisms and microstructure formation
compatible with automotive applications, to date, in a high power ultrasonic spot-welded 6111 Al
only a few studies have reported on joining alu- sheet and sound welds were obtained with a shear
minum.9,13,16–19 It was reported that USW of opti- failure load up to 3.5 kN. Strong welds could be
mized Al 1 mm- to 2-mm-gauge sheet joints are achieved only after the welding energy exceeded a
produced at weld energies in the range of 0.7– threshold, then a failure mode of button pullout
1.5 kJ.9 At a lower level of weld energy, the occurred.10 Typically, the welding energy required
microbond density was too low, failure occurred during USW is closely related to the clamping
through the interface.10 Higher weld energies can pressure and welding time, as well as the vibration
effectively promote joining in the entire weld, thus frequency. However, with the exception of the
improving weld strength. Jahn et al.13 evaluated notable work of Jahn et al.4,13 and a few other
the effects of welding energy on the microstructure authors,10,20 the integrity and durability issues of
and weld strength of ultrasonic spot welds of similar Al welds especially under cyclic loading
AA6111-T4 Al alloy using a single transducer, must be considered in the design of many struc-
unidirectional wedge-reed welding machine. It was tures. To the authors’ knowledge, there is no
observed that the weld strength increased with systematic studies on fatigue behavior for such
increasing welding energy up to 500 J, and the USW similar thicker-sheet Al welds in relation to
maximum failure load obtained was between 2.6 kN the change in the failure mode that has been
and 3.1 kN. Bakavos and Prangnell16 reported the reported in the open literature. This study was,

Fig. 1. EBSD orientation maps of an USWed Al5754 similar joint at a welding energy of 2500 J: (a) BM and (b) near the nugget zone.
Microstructure and Fatigue Properties of Ultrasonic Spot Welded Joints 1467
of Aluminum 5754 Alloy

Fig. 2. Grain size distribution of an USWed Al5754 similar joint at a welding energy of 2500 J: (a) BM and (b) near the nugget zone.

Fig. 3. Grain boundary mapping of an USWed Al5754 similar joint at a welding energy of 2500 J: (a) BM and (b) near the nugget zone.

therefore, aimed to gain a better understanding of USW system. An energy input ranging from 500 J
the lap shear tensile load and fatigue properties of to 3000 J, a constant power setting of 2000 W with
USWed-similar Al5754-O Al alloy joints, with a an impedance setting of 8 and a pressure of
particular emphasis on the weld microstructure 0.414 MPa on the machine were used for welding
evolution. the samples. Five samples were welded in each
welding condition. One of them was used for
microstructural examination, specifically electron
MATERIAL AND EXPERIMENTAL
back-scattered diffraction (EBSD), two were used
PROCEDURE
for the lap shear tensile tests and the remaining two
In the present study, commercial 1.5-mm-thick were used for fatigue tests. The metallographic
sheet of Al alloy Al5754-O (3.42% Mg, 0.63% Mn, samples for the EBSD examination were cut from
0.23% Sc, 0.22% Zr, and balance Al) was selected for the weld cross-section, then ground and polished
USW. The specimens, 60 mm long and 15 mm wide, using diamond paste and MasterPrep. EBSD obser-
were sheared and 120 emery papers were used to vations were performed on the as-polished material
ground the surfaces. The samples were cleaned and using JEOL JSM-6380LV-HKL equipped with a
dried before welding. The welding was performed Nordlys detector for EBSD analysis. An acceleration
with a dual wedge-reed Sonobond-MH2016 HP- voltage of 20 keV and a working distance of 15 mm
1468 Mirza, Macwan, Bhole, and Chen

Fig. 4. Misorientation angle distribution of an USWed Al5754 similar joint at a welding energy of 2500 J: (a) BM and (b) near the nugget zone.

100
system at different maximum loads. To avoid poten-
90 tial buckling of the test specimens, tension–tension
Lap shear strength, MPa

80 cyclic loading at a stress ratio of R (Pmin/Pmax) = 0.2


70 was applied at a frequency of 50 Hz and sinusoidal
60
waveform. Two samples were tested at each cyclic
load level. To prevent the rotation and bending
50 moment of the specimen during the tensile and
40 fatigue tests, two spacers with a thickness of
30 1.5 mm, width of 15 mm and length of 35 mm were
Interfacial Nugget attached at both ends of the specimen. The fatigue
20 failure zone
failure fracture surfaces were examined via a scanning
10 electron microscope (SEM) to identify the initiation
0 sites and propagation mechanism of the fatigue
0 500 1000 1500 2000 2500 3000 3500 crack.
Weld energy, J
Fig. 5. Lap shear tensile strength of an USWed similar Al5754 alloy
RESULTS AND DISCUSSION
at different energy inputs. Microstructure Evolution
Figure 1a and b shows the EBSD orientation
image mapping of base metal (BM) and the center of
the weld nugget (WZ) of USWed-similar Al5754
were used for the examination of sample alloy sample (2500 J energy input), respectively. It
microstructural development. The sample was tilted can be seen from Fig. 1a that the BM exhibits
70° for the collection of data. Analysis of the EBSD almost equiaxed grain structure. The average grain
results was performed using MTEX software. To size was approximately 8 lm (as shown in Fig. 2-
evaluate the mechanical strength of the joints and a). It can be seen from Fig. 1b that the material at
establish the optimum welding conditions, tensile the center of the weld nugget exhibits slightly more
shear tests of the welds were conducted to measure equiaxed and finer grain structure than the base
the lap shear failure load using a fully computerized metal due to dynamic recrystallization.12,21–24 The
United testing machine with a constant crosshead grain size distribution clearly indicates finer grain
speed of 1 mm min 1 in air at room temperature. At size than base metal, i.e., approximately 7 lm as
least two samples were tested at each energy level; shown in Fig. 2b. A similar type of fine grain
however, due to the experimental scatter at certain formation has been reported at the weld zone in
energy levels, i.e., 1500 J and 2000 J, five samples ultrasonic welding16 and friction stir welding25,26
were tested. Load control fatigue tests were per- for Al alloys. Dislocations would be introduced due
formed in accordance with ASTM E466 on a fully to severe plastic deformation during welding, fol-
computerized Instron 8801 servo-hydraulic testing lowed by quick recrystallization at higher frictional
Microstructure and Fatigue Properties of Ultrasonic Spot Welded Joints 1469
of Aluminum 5754 Alloy

Fig. 6. Typical tensile failure locations (indicated by yellow arrows) of an USWed similar Al5754 alloy at different energy inputs: (a) 500 J, (b)
1000 J, (c) 1500 J, (d) 2000 J, (e) 2500 J, and (f) 3000 J.

temperature leading to the formation of fine grains newly formed grains deform as they grow, and the
in and near the weld zone. It has been reported that build-up of stored strain energy in the new grains
optimum weld strengths are obtained when exten- may reduce the driving force for their continued
sive plastic deformation takes place throughout the growth but increase the frequency of initiation of
entire sheet thickness between the sonotrode new grains. A ‘‘necklace’’-type structure of newly
tips.16,27 It has also been reported that discontinu- formed grains may result when prior grain bound-
ous dynamic recrystallization (DDRX) occurs in aries are the initiation sites for DDRX. This is
metals of moderate to low stacking fault energy reflected in almost equiaxed subgrains with nearly
during hot deformation.25 In DDRX, new, disloca- dislocation-free interiors, constant subgrain size
tion-free grains form at sites such as prior grain and subgrain boundary misorientation of grains
boundaries, deformation band interfaces or bound- near the weld zone (Fig. 3b) as compared with the
aries of newly recrystallized grains (Fig. 1b). The base metal (Fig. 3a). A condition for initiation of
driving force for the initiation of recrystallization is DDRX may be given by,25
the strain-induced build-up of stored strain energy
in the form of fine cells and a high free dislocation q3m 2cb
i ð1Þ
density.25 During deformation, a low stacking fault e_ KLMGb5
;
energy retards the climb and cross-slip processes,
where qm is the mobile dislocation density, e_ is the
thereby enabling the formation of stable grain
strain rate, cb is the grain boundary energy, K is a
nuclei at various sites in the deformation
constant fraction of the dislocation line energy that
microstructure at the center of the nugget zone as
is stored in the newly formed grains, L is mean slip
pointed out by the arrow (as shown in Fig. 1b). The
1470 Mirza, Macwan, Bhole, and Chen

Fig. 7. Typical SEM images of tensile fracture surface of an USWed Al5754 similar joint at a welding energy of 2500 J: (a) and (b) overall view
(with an arrow indicating the crack initiation), (c) and (d) at lower magnification, (e) and (f) at higher magnification.

distance of dislocations in these grains, M is the values of qm .25 Generally, if the restricted number of
boundary mobility, G is the shear modulus and b is slip systems or solute drag inhibit the slip contri-
Burger’s vector. This inequality suggests a strong butions and boundary migration, lattice rotation
dependence of the conditions for DDRX on the may develop progressively at grain boundaries as
mobile dislocation density: a low stacking fault dislocations accumulate during straining. This may
energy tends to suppress recovery, leading to high lead to dynamic recrystallization by lattice rotation
Microstructure and Fatigue Properties of Ultrasonic Spot Welded Joints 1471
of Aluminum 5754 Alloy

3.5 Lap Shear Tensile Strength and Failure Mode


(a) 1000J
3.0 2000J The lap shear failure strength of USWed similar
Maximum load, kN 2.5
Al5754 alloy as a function of welding energy is
shown in Fig. 5. The average maximum lap shear
2.0 strength was found to be 85 MPa at 2000 J energy
input. As shown in Fig. 5, the lap shear tensile
1.5
strength increased with increasing energy inputs
1.0 due to high temperatures and strain rate, which
leads to increased diffusion between the Al alloy
0.5
sheets. However, at high energy inputs, the lap
0.0 shear strength was decreased. At lower energy
1E+0 1E+2 1E+4 1E+6 inputs, temperature was not high enough to soften
Number of cycles to failure, Nf the sample, thereby Al has not been fully deformed
2.4 to fill up the valleys of the knurl patterns of the
(b) sonotrodes and exhibits very low strengths, i.e., 20–
40 MPa for 500 J and 1000 J. It was observed that,
2 Interfacial failure mode in the cases of lower energy inputs, the lap shear
tensile fracture occurred from the Al/Al interface, as
max

shown in Fig. 6 (typical tensile failure locations of


1.6
USWed-similar Al5754 alloy at different energy
Log

Transverse through-thickness
crack growth mode inputs). While on the other hand, at higher energy
1.2 inputs, the weld specimen was subjected to higher
temperatures under a larger vibration amplitude
for a longer time. As a result, more diffusion
0.8 occurred between the two Al alloy sheets, which
0 2 4 6 8 10 deteriorated the strength. This corresponded well to
Log (2Nf)
the failure that occurred at the edge of the nugget
Fig. 8. (a) S-N curves of USWed similar Al5754 alloy joints zone (Fig. 6f). Thus, it can be concluded that this
tested at RT, R = 0.2, and 50 Hz, and (b) the maximum tensile was a consequence of the competition between the
shear stress versus the number of reversals to failure (2Nf) in
the double-log scale for welded joints at a welding energy of increasing diffusion bonding arising from higher
2000 J. temperatures at the higher energy inputs and the
deterioration effect of the Al/Al interface.11 Similar
results have been reported by Bakavos and Prang-
nell16 and Zhang et al.10 For dissimilar Al alloy to
Mg alloy sheet welded joints, Patel et al.11 also
as new grains form at prior boundaries, and a
reported that the lap shear strength was increased
necklace structure similar to that often observed in
with increasing energy input and it was decreased
DDRX (Fig. 1b). In other words, the necklace-type
at very high energy inputs.
structure of the newly formed grains can be inter-
preted as the mechanical interlocking as reported
Lap Shear Tensile Fractography
by Zhang et al.10 since the welding mechanisms of
the Al alloy involve solid-state deformation and Typical SEM images of tensile fracture surfaces of
mechanical interlocking, as well as the formation USWed similar Al 5754 alloy at 2500 J energy
and progressive spreading of microwelds.16,28 This inputs are shown in Fig. 7. As seen from the overall
is associated with a considerable rise in tempera- view of both sides of the joint (Fig. 7a and b, a bright
ture in the weld zone, which softens the material zone was observed at the outside periphery of the
and allows the sonotrode tips to sink into the sheet weld. As identified by Zhang et al.,10 this zone is an
surfaces. In the process, the weld interface is oxide layer formed during welding due to the direct
displaced into complex wave-like flow pat- contact of air with the material at an elevated
terns10,16,27,28 (Fig. 1b). Figure 3 shows the grain temperature. It can also be seen from Fig. 7a and b
boundary mapping of USWed-similar Al5754 alloy that there is a dark zone just inside the oxide zone
of base metal and near the nugget zone. The which represents the bonded zone (as indicated by
frequencies of misorientaion of grains are mostly in label ‘‘A’’). It can be seen from the low-magnification
the range of 35°–50° (as indicated by the black and images (Fig. 7a and b) that the crack initiated from
yellow lines) for both cases. The wave-like patterns one end of the bonding zone as indicated by the
or ‘‘necklace’’-type grain structure is also obvious arrow, where a big crack can be seen from the
from Fig. 3b and the misorientation angles in that border area (Fig. 7b). The initiation site (bonded
section are in the range of 55°–60°. The corre- zone) had obvious dimple-rupture failure features
sponding misorientation angle distribution of as observed at higher magnifications (Fig. 7c and d).
USWed-similar Al5754 alloy is also shown in At higher magnification (Fig. 7e and f), the elon-
Fig. 4. gated dimples indicate the shear fracture via void
1472 Mirza, Macwan, Bhole, and Chen

Fig. 9. Typical fatigue failure locations (indicated by yellow arrows) at different Pmax values of an USWed similar Al5754 alloy at 2000 J energy
input: (a) 3 kN (interfacial failure mode), (b) 2.5 kN (interfacial failure mode), (c) 2 kN (interfacial failure mode, initiation of crack), (d) 1.5 kN
(interfacial failure mode, crack starting to propagating perpendicular to the loading direction), (e) 1 kN (transverse through-thickness crack growth
mode), and (f) 0.5 kN (transverse through-thickness crack growth mode).

formation/nucleation, growth, and coalescence.10 occurring at the edge of the nugget zone on the
There are also several secondary cracks which base metal at lower Pmax loads as reported by Patel
occurred during the tensile shear. This suggested et al.,3 which is referred to here as the transverse
that this occurred where the base material locally through-thickness crack growth mode. Typically,
reached its fracture strength and the fracture the transverse through-thickness crack growth
initiated in that area. Apart from some small, mode occurred (perpendicular to the loading direc-
isolated dimples and tearing ridges, the predomi- tion) at the edge of the sonotrode in contact with
nant feature in the center of the joint was the the sample due to the higher welding energies
scrubbing lines parallel to the loading direction (leading to greater temperatures and a greater
(Fig. 7e), which might be generated due to abrasion level of bending deformation in the weld zone).3 As
or galling of asperities of the microscale roughness a result, a small micro-level crack tip at the notch
of the ground sheet surfaces.29 of the two sheets was produced and this crack tip
experienced a higher stress concentration effect
which allowed the cracks to grow in the transverse
Fatigue Behavior and Failure Mode
direction of the Al sample. In the 2000-J USWed
Fatigue tests of the similar USWed Al5754 Al sample, the results of the fatigue tests were
alloy joints were carried out at room temperature consistent with the tensile results, which showed
at a stress ratio of R = 0.2, and a frequency of the highest fatigue life at the applied Pmax load of 2
50 Hz. The obtained S–N curves are plotted in kN or less. It can be noticed from Fig. 8a that the
Fig. 8a. Like the tensile testing (Fig. 5), it can be fatigue limit was 0.5 kN for all welding conditions.
seen that at the higher Pmax loads of 2 kN, 2.5 kN, Figure 8b presents the failure mode in conjunction
and 3 kN, the 2000-J samples showed higher with the logarithmic S–N plots in the form of the
fatigue resistance compared to the 1000-J welded maximum tensile shear cyclic stress versus the
samples. However, lower fatigue life was obtained number of reversals to failure (2Nf) in terms of a
for the for the 2000-J welded samples at lower Basquin-type relationship for the 2000-J USWed
Pmax loads especially at 0.5 kN and 1 kN as samples. It should be noted that, since there were
compared to the 1000-J welded samples. This no run-out data of non-failed samples at or over
might due to the premature fatigue fracture 1 9 107 cycles, all the data points were included in
Microstructure and Fatigue Properties of Ultrasonic Spot Welded Joints 1473
of Aluminum 5754 Alloy

Fig. 10. Typical SEM images of fatigue fracture surface of an USWed Al5754 similar joint at a welding energy of 2000 J tested at a Pmax of 2.0
kN: (a) overall view, (b) crack initiation area, (c) magnified image of crack initiation area, and (d and e) fatigue striations in the crack propagation
zone at higher magnifications.

the curve fitting. For the 2000-J welded samples, Now, shear stress acting on the area was such that
interfacial failure mode was observed at the higher the remaining cross-section could not withstand
cyclic load levels where the slope was flatter, while the shear overload, and the severely deformed
transverse through-thickness crack growth mode nugget edge in due course experienced transverse
was observed at the lower cyclic load levels where through-thickness crack growth failure mode
the slope was steeper. Figure 9 shows a typical (Fig. 9e and f).
fatigue failure location at the different Pmax load On the other hand, in the 1000-J samples, most of
values of USWed -imilar Al5754 alloy joint at the failure occurred at the interface (except for the
2000 J energy input. It can be seen that, when the 0.5-kN applied fatigue load where the failure was
samples were tested at cyclic loads of 3 and 2.5 kN, through-thickness). The 1000-J welded samples
interfacial failure occurred. When samples were experienced lower sonotrode penetration (i.e., no
tested at 2 kN and 1.5 kN, interfacial failure also micro-level cracks at the edge of the nugget zone).
occurred; however, initiation of cracks was started The change of the failure mode from the transverse
at the circumference around the nugget (Fig. 9c through-thickness crack growth mode to the inter-
and d). At 1 kN applied load, the crack was facial failure occurred at logrmax = 1.49 or rmax =
propagated perpendicular to the loading direction. 31 MPa in the 2000-J USWed joints (Fig. 8b).
1474 Mirza, Macwan, Bhole, and Chen

Likewise, such a transition of failure mode from the which usually appeared perpendicular to the
deformed edge failure to the interfacial failure fatigue crack propagation direction.
occurred at logrmax = 1.32 or rmax = 21 MPa in the
1000-J samples (not shown in Fig. 8b). Therefore, it ACKNOWLEDGEMENTS
can be concluded that the maximum cyclic stress for
the transition of fatigue fracture mode from the The authors would like to thank the Natural
transverse through-thickness crack growth to the Sciences and Engineering Research Council of Ca-
interfacial failure increased with increasing energy nada (NSERC) and AUTO21 Network of Centres of
input. Excellence for providing financial support. The au-
thors thank Dr. A.A. Luo, Ohio State University
Fatigue Fractography (formerly General Motors Research and Develop-
ment Center) for the supply of test materials. One of
For a more detailed study of the fracture frac- the authors (D.L. Chen) is grateful for the financial
tures, SEM fracture surface images of the failed support by the Premier’s Research Excellence
2000-J welded samples tested at a minimum cyclic Award (PREA), NSERC-Discovery Accelerator
load of 2 kN are shown in Fig. 10. It can be seen Supplement (DAS) Award, Canada Foundation for
from Fig. 10a that the fatigue crack developed from Innovation (CFI), and Ryerson Research Chair
the region ‘‘b’’ and propagated at right angles to the (RRC) program. The authors would also like to
loading direction along the Al sheet with increasing thank Messrs. A. Machin, Q. Li, J. Amankrah and
number of cycles, and later experienced transverse R. Churaman for easy access to the laboratory
through-thickness crack growth mode (Fig. 9e). facilities of Ryerson University and their assistance
Figure 10b shows the location of the crack initiation in the experiments.
and Fig. 10c shows the magnified image of region
‘‘c’’ in Fig. 10b. Fatigue crack propagation was REFERENCES
mainly characterized by the formation of fatigue 1. S. Chu and A. Majumdar, Nature 488, 294 (2012).
striations (Fig. 10d and e), which usually appeared 2. M. McNutt, Science 341, 435 (2013).
3. V.K. Patel, S.D. Bhole, and D.L. Chen, Metall. Mater. Trans.
perpendicular to the fatigue crack propagation A 45, 2055 (2014).
direction. Fatigue striations normally occur by a 4. G. Wagner, F. Balle, and D. Eifler, JOM 64, 401 (2012).
repeated plastic blunting–sharpening process in the 5. P.B. Prangnell, F. Haddadi, and Y.C. Chen, Mater. Sci.
face-centered cubic materials, stemming from the Technol. 27, 617 (2011).
glide of dislocations on the slip plane along the slip 6. L. Xiao, L. Liu, Y. Zhou, and S. Esmaeili, Metall. Mater.
Trans. A 41–6, 1511 (2010).
direction within the plastic zone ahead of the 7. D. Bakavos and P.B. Prangnell, Sci. Technol. Weld. Join. 14,
fatigue crack tip.14 443 (2009).
8. T. Liyanage, J. Kilbourne, A.P. Gerlich, and T.H. North, Sci.
CONCLUSION Technol. Weld. Join. 14, 500 (2009).
9. P.B. Prangnell and D. Bakavos, Mat. Sci. Forum. 638–642,
The following conclusions can be drawn from this 1237 (2010).
10. C.Y. Zhang, D.L. Chen, and A.A. Luo, Weld. J. 93, 131s
investigation: (2014).
(1) A special ‘‘necklace’’-type structure containing 11. V.K. Patel, S.D. Bhole, and D.L. Chen, Sci. Technol. Weld.
Join. 17, 202 (2012).
newly formed very fine grains was observed 12. L.E. Murr, G. Liu, and J.C. McClure, J. Mater. Sci. Lett. 16,
along the weld line after USW due to the 1801 (1997).
occurrence of dynamic crystallization, along 13. R. Jahn, R. Cooper, and D. Wilkosz, Metall. Mater. Trans. A
with the presence of mechanical interlocking. 38, 570 (2007).
14. S.H. Chowdhury, D.L. Chen, S.D. Bhole, X. Cao, and P.
(2) The lap shear tensile strength increased with Wanjara, Mater. Sci. Eng. A 556, 500 (2012).
increasing energy inputs and reached the 15. S. Elangovan, S. Semeer, and K. Prakasan, J. Mater. Pro-
maximum of 85 MPa at an energy input of cess. Technol. 209, 1143 (2009).
2000 J and then decreased. The lap shear 16. D. Bakavos and P.B. Prangnell, Mater. Sci. Eng. A 527, 6320
tensile fracture occurred from the Al/Al inter- (2010).
17. A. Macwan, X.Q. Jiang, and D.L. Chen, JOM 67, 1468
face in the cases of lower energy inputs, (2015).
whereas at higher energy inputs the failure 18. A. Macwan and D.L. Chen, Mater. Des. 84, 261 (2015).
of the welded specimen occurred at the edge of 19. J. Schneider and R. Radzilowski, JOM 66, 2123 (2014).
the nugget zone. 20. S.M. Allameh, C. Mercer, D. Popoola, and W.O. Soboyejo,
J. Eng. Mater. Technol. 127, 65 (2005).
(3) In both welding conditions, the fatigue limit 21. N. Pathak, K. Bandyopadhyay, M. Sarangi, and S.P.
was found to be 0.5 kN. The maximum cyclic Kumar, JMEPEG 22, 131 (2013).
stress at which transition of the fatigue frac- 22. A. Panteli, Y.C. Chen, D. Strong, X. Zhang, and P.B.
ture mode from the transverse through thick- Prangnell, JOM 64, 414 (2012).
ness crack growth to the interfacial failure 23. F. Khodabakhshi, M. Haghshenas, S. Sahraeinejad, J.
Chen, B. Shalchi, J. Li, and A.P. Gerlich, Mater. Charact.
increased with increasing energy input. Fati- 98, 73 (2014).
gue crack propagation was mainly character- 24. L. Ying, L.E. Murr, and J.C. McClure, Mater. Sci. Eng. A
ized by the formation of fatigue striations, 271, 213 (1999).
Microstructure and Fatigue Properties of Ultrasonic Spot Welded Joints 1475
of Aluminum 5754 Alloy
25. Z. Zhang, B.L. Xiao, and Z.Y. Ma, Mater. Charact. 106, 255 27. Y.C. Chen, D. Bakavos, A. Gholinia, and P.B. Prangnel, Acta
(2015). Mater. 60, 2816 (2012).
26. H.E. Hu, L. Zhen, B.Y. Zhang, L. Yang, and J.Z. Chen, 28. F. Haddadi, D. Strong, and P.B. Pragnell, JOM 64, 407 (2012).
Mater. Charact. 59, 1185 (2008). 29. H.P.C. Daniels, Ultrasonics 3, 190 (1965).
Materials 2015, 8, 5138-5153; doi:10.3390/ma8085138
OPEN ACCESS

materials
ISSN 1996-1944
www.mdpi.com/journal/materials
Article

A Unified Model for the Prediction of Yield Strength in


Particulate-Reinforced Metal Matrix Nanocomposites
F. A. Mirza and D. L. Chen *

Department of Mechanical and Industrial Engineering, Ryerson University, 350 Victoria Street,
Toronto, ON M5B 2K3, Canada; E-Mail: f4mirza@ryerson.ca

* Author to whom correspondence should be addressed; E-Mail: dchen@ryerson.ca;


Tel.: +1-416-979-5000 (ext. 6487); Fax: +1-416-979-5265.

Academic Editor: Geminiano Mancusi

Received: 9 July 2015 / Accepted: 5 August 2015 / Published: 10 August 2015

Abstract: Lightweighting in the transportation industry is today recognized as one of the


most important strategies to improve fuel efficiency and reduce anthropogenic climate-changing,
environment-damaging, and human death-causing emissions. However, the structural
applications of lightweight alloys are often limited by some inherent deficiencies such as
low stiffness, high wear rate and inferior strength. These properties could be effectively
enhanced by the addition of stronger and stiffer reinforcements, especially nano-sized
particles, into metal matrix to form composites. In most cases three common strengthening
mechanisms (load-bearing effect, mismatch of coefficients of thermal expansion, and
Orowan strengthening) have been considered to predict the yield strength of metal matrix
nanocomposites (MMNCs). This study was aimed at developing a unified model by taking
into account the matrix grain size and porosity (which is unavoidable in the materials
processing such as casting and powder metallurgy) in the prediction of the yield strength of
MMNCs. The Zener pinning effect of grain boundaries by the nano-sized particles has also
been integrated. The model was validated using the experimental data of magnesium- and
titanium-based nanocomposites containing different types of nano-sized particles (namely,
Al2O3, Y2O3, and carbon nanotubes). The predicted results were observed to be in good
agreement with the experimental data reported in the literature.

Keywords: metal matrix nanocomposites; Orowan strengthening effect; Hall-Petch


relationship; Zener pinning effect
Materials 2015, 8 5139

1. Introduction

Lightweighting in ground vehicles is deemed as one of the most effective strategies to improve fuel
economy and reduce anthropogenic climate-changing, environment-damaging, costly, and human
death-causing emissions [1–7] due to the tremendous environmental concerns. To manufacture lightweight
vehicles, advanced high-strength steels, aluminum alloys, magnesium (Mg) alloys, and polymers are
being used in the automotive and aerospace sectors, but substantial weight reductions could be further
achieved by employing ultra-lightweight Mg alloys due to their low density, high strength-to-weight ratio,
and superior damping capacity [1,4,8,9]. However, the applications of Mg alloys are often restricted by
some inherent deficiencies such as low stiffness, high wear rate, and inferior creep resistance or
mechanical strength at elevated temperatures [10,11]. Reinforcement with a discontinuous phase
especially nano-sized particles has been considered to be the most favored choice by researchers in recent
years to improve physical, mechanical, and damping properties of Mg beyond the limits dominated by
traditional alloying [11–19].
In the past few years, several studies have been done to develop constitutive relationships that can be
used to predict the mechanical properties of metal matrix nanocomposites (MMNCs) as a function of
the reinforcement, matrix, and processing conditions [10,15,20–27]. Ramakrishnan [20] predicted the
yield strength of micro-sized particulate-reinforced metal matrix composites (MMCs), using a composite
sphere model for the intra-granular type of MMCs and incorporating two improvement parameters
associated with the dislocation strengthening of the matrix and the load-bearing effect of the reinforcement.
Recently, Zhang and Chen [10,21] predicted the yield strength of MMNCs via considering the Orowan
strengthening mechanism, enhanced dislocation density due to the mismatch of coefficients of thermal
expansion (CTE) between the reinforcement and matrix, and load-bearing effect which has been used
by many researchers, e.g., [15,23–27], to predict the yield strength of particulate-reinforced MMNCs.
However, in all of the above models no effect of porosity was taken into account, which could lead to
an overestimate of the yield strength of composites. In the composite processing (casting, powder
metallurgy, electrodeposition, plasma and cold spray, etc.), the complete elimination of porosity is either
difficult or impossible [28–37]. Porosity is among the dominant factors causing failure of discontinuous
reinforced metal matrix composites (DRMMCs) in the tensile tests [17]. The commonly encountered
spherical pores or gas porosity are observed to create stress concentrations and thus lead to failure.
The ductility and toughness of most metal based composites are influenced by the presence of voids and
the balance between reinforcing particles sharing the load [14]. The large amount of porosity associated
with oxides is often surrounded by either individual or clusters of reinforcement particles, which
significantly decreases the mechanical properties of DRMMCs [32]. Also, it was reported that the
presence of nanoparticles led to a higher level of porosity and more irregular pores with bigger pore size
in the nanocomposites [34]. Thus, it is necessary to consider the effect of porosity on the mechanical
properties of nanocomposites.
Furthermore, very limited investigations have been reported to account for the influence of matrix
grain size in the MMNCs, while the relation between the yield or flow stress and grain size in
polycrystalline alloys was well established by Hall and Petch [38,39]. Recrystallization studies in
DRMMCs have shown that the nucleation potency increases with increasing reinforcement size and
volume fraction, called particle stimulated nucleation (PSN) [40,41]. It would be expected that the
Materials 2015, 8 5140

nucleation density is high and the recrystallized grain size is smaller than the interparticle spacing, since
normal grain growth continues after complete recrystallization until the grain boundaries are pinned by
the reinforcement particles. As reported by Hassold et al. [42], the inert, second-phase particles can
inhibit grain growth and lead to a pinned microstructure where grain growth ceases (as shown in Figure 1).
Most particles are located on the grain boundaries, which are strongly pinned by the particles. The grain
boundaries are highly non-random in location and the particles are more closely spaced as the volume
fraction increases. Also, the pinned grain size decreases with increasing volume fraction of particles
(Figure 1b). Zener [40–45] first observed this effect, commonly called Zener pinning (or Zener drag) [41,45],
and predicted the dependence of the pinned grain size on the volume fraction and radius of the particles.
The presence of Zener pinning in the MMNCs will further decrease the matrix grain size, depending on
the second-phase particles size and volume fraction. The objective of this study was, therefore,
to develop a unified model of integrating all the above strengthening effects to achieve a more accurate
prediction of the yield strength of MMNCs, which was then validated using the experimental data of
magnesium-based nanocomposites available in the literature.

Figure 1. Pinned microstructures of nanocomposites for two volume fractions of particles


of (a) Vp = 0.01 and (b) Vp = 0.05 [42].

2. Model Development

To facilitate a better understanding of an analytical model for predicting the yield strength of
MMNCs, a brief review on the underlying factors affecting the yield strength of composites is given
as follows.

2.1. Load Transfer to the Reinforcement Particles

Due to the nano-size of the reinforcement particles and the sound synthesizing methods, there is a
strong cohesion at the atomic level between the matrix and nano-sized particles, i.e., the nano-sized
particles are directly bonded to the matrix [10,12,20–22,46]. Then the nano-sized particles would carry
a certain portion of load depending on the volume fraction, giving rise to some degree of load transfer
between matrix and reinforcement [10,21,22]. Provided that the porosity is present in the composites,
the porosity would affect the load-bearing term. Thus, the improvement factor associated with the
load-bearing effect of reinforcement in the presence of porosity can be derived as [46],
Materials 2015, 8 5141

Vp  P
1
fl (1)
2
where fl is an improvement factor associated with load-bearing effect, P is the volume fraction of porosity,
and Vp is the volume fraction of the reinforcement nano-particles.

2.2. Dislocation Density

In MMNCs, the increased interfacial area between the reinforcement and matrix contributes to the
enhanced mechanical properties due to the presence of nano-sized particles [10,21]. Also, thermal
mismatch dislocations in the matrix around the nano-sized particles would be generated to relieve the
thermal stresses occurred at the interface during cooling from the processing temperature. The thermal
stresses around the nanoparticles would be large enough to induce plastic deformation in the matrix near
the interface region [13,47,48]. The improvement factor related to the dislocation density in the matrix,
fd can be expressed as follows [10,21,46],
fd kGm b ȡ / ı ym (2)

12'Į'TVp
bd p (1  Vp )
ȡ (2a)

where ıym is the yield strength of the monolithic matrix without any porosity, Gm is the shear modulus
of the matrix, b is the Burgers vector of the matrix, k is a constant (approximately equal to 1.25, based
on the theoretical estimates in reference [10]), ȡ is the enhanced dislocation density due to the difference
in the coefficients of thermal expansion between the reinforcement phase and the matrix (ǻĮ), dp is the
particle size, and ǻT is the difference between the processing and test temperatures.

2.3. Orowan Strengthening

In the precipitation-hardened alloys the internally precipitated nano-sized particles are proven to be
either sheared (for smaller and softer particles) or by-passed by dislocations (i.e., Orowan strengthening
mechanism of dislocation bowing) for larger and harder particles [37]. In the case of MMNCs where the
hard and external nano-sized particles are added, it would be reasonable to assume that the Orowan
strengthening mechanism would occur [10,21]. The improvement factor related to the Orowan
strengthening of nanoparticles added, fOrowan can be expressed as follows [10,21,46],
0.13Gmb r
fOrowan ln (3)
Ȝı ym b

ª º
§ 1 ·3 »
Ȝ | d p ««¨
1

¨ 2V ¸¸  1»
«¬© p ¹ »¼
(3a)

where r is the particle radius (r = dp/2), and Ȝ is the interparticle spacing.


Materials 2015, 8 5142

2.4. Porosity

As mentioned above, some inherent characteristics such as porosity were inevitably present in
MMNCs which would have a significant effect on the yield strength of MMNCs [28–37]. For example,
the occurrence of porosity as discontinuities in cast MMC interrupted the balance between the
reinforcing particles carrying the load, generated a stress concentration and facilitated the crack initiation
and propagation, and thus reduced its mechanical properties [32]. Also, most fatigue cracks initiated
from the discontinuities in materials, mainly in the highly stressed regions of components. With
increasing volume fraction of reinforcement particles, the likelihood of forming the processing-induced
voids became higher, leading to a degradation of the yield strength [12]. Accordingly, it is necessary to
consider the effect of porosity in more realistically predicting the yield strength of MMNCs. Based on
our previous publication [46], fporosity is the deterioration factor associated with the presence of porosity
in MMNCs, which could be expressed as [29],
f porosity 1  e nP (4)

where n is an empirical constant depending on the porosity characteristics such as pore size, geometry,
and orientation [29]. If the average pore shape was assumed to be close to cylinder orientated between
45° and 90° with respect to the loading axis, n could be estimated as [29],

 0.318  1.22
dp
n 0.405 l (4a)
dp l

where l is the length of the particle, respectively.


Based on the above mechanisms, the concept of multiplication as reported by Zhang and Chen [10,21],
now referred to as the “Zhang and Chen method”, “Zhang and Chen approach”, “Zhang and Chen
model”, or “Zhang and Chen (ZC) summation method”, e.g., in [26,49–53], was used to account for both
additive and synergistic effects of the strengthening and weakening factors, and a further modified
analytical model for predicting the yield strength of MMNCs with consideration of the effect of porosity
was proposed as follows [46],
ı yc ı ym (1  f l )(1  f d )(1  f O ro w an )(1  f p o ro sity ) (5)
where ıyc is the yield strength of MMNCs. It should be noted that since the yield strength of MMNCs is
considered in this model, no strain hardening effect during the subsequent plastic deformation beyond
yielding is taken into account.

2.5. Effects of Grain Size on the Yield Strength of MMNCs

The influence of grain size on the mechanical properties is complex since the grain boundaries may
either act as obstacles to dislocation slip (strengthening effect) or provide a positive contribution to the
deformation of the material (softening effect). However, the following well-known Hall-Petch relation
between the yield stress and grain size was proposed [38,39,54],
ı YS V o  kd  1/ 2 (6)
where d is the grain size, k and ıo are constants. This equation was revisited Li et al. in terms of the
collective motion of interacting dislocations and rearranged as follows [55],
Materials 2015, 8 5143

ıYS ıo  KG b d (6a)

where K is constant, G is the shear modulus, and b is the burgers vector. The typical value of K for pure
fcc metals is 0.05 to 0.5 [54,56].
As mentioned earlier, the second-phase particles can impede grain growth and lead to a pinned
microstructure where grain growth ceases [40]. The dependence of the pinned grain size on the volume
fraction and radius of the particles could be expressed as follows [40,42–44],
4rp
R (7)
3Vp

where R is the matrix grain radius. By combining Equations (6a) and (7), a modified Hall-Petch equation
with consideration of Zener pinning is obtained as follows,
§ 4d ·
ı o  KG b ¨ p ¸¸
¨ 3Vp
ı YS
© ¹
(8)

and the improvement factor can be expressed as,


f H all-P etch -Z en er ' ı H all-P etch -Z en er / ı ym (9)
where ' ı H all-P etch-Zener is,

'ı Hall-Petch-Zener
3bVp
KGm (9a)
4d p

Based on this additional consideration of the grain growth retardation by dispersed-particle-pinning


of grain boundaries, the analytical model for predicting the yield strength of MMNCs becomes,
ı yc ı ym (1  f l )(1  f d )(1  f O ro w an )(1  f H all-P etch -Z en er )(1  f P o ro sity ) (10)
Now substituting Equations (1)–(4a) and (9) into Equation (10) and considering ' T T p ro c e ss  T te st
and ' Į Į m  Į p (the difference in the coefficients of thermal expansion between the reinforcement phase
and the matrix), the following equation for the yield strength of MMNCs can be derived,
§ AB ·§ C ·  nP
(1  0.5Vp  P) ¨ ı ym  A  B  ¸¨ 1 ¸¸ e
¨ ı ym ¸¨
ı yc
© ¹© ı ym ¹
(11)

where A, B, and C can be expressed as follows,


12(Tprocess  Ttest )(Įm  Įp )Vp
bdp (1  Vp )
A 1.25Gmb (11a)

0.13Gmb d
ª º 2b
B ln p
« § 1 ·3 »
1

dp «¨  1»
¨ 2V ¸¸
(11b)

«¬© p ¹
»¼
Materials 2015, 8 5144

3bVp
C KGm (11c)
4d p

3. Results and Model Validation

First of all, the individual contribution of all improvement factors (using Equations (1)–(3) [10,21,46],
and (9)) with respect to the size and volume fraction of nanoparticles in Mg/Al2O3 nanocomposites is
identified and plotted in Figures 2 and 3. Considering Vp + Vm + P = 1 and substituting Equation (4a)
into Equation (4), the following equation for fporosity can be derived,
§ ·

¨ 0.405 l  0.318 p 1.22 ¸ 1Vp Vm
f porosity 1  e
d
© dp l ¹ (12)

where Vm is the volume fraction of matrix. Figure 2 shows the variation of the improvement factors with
the nanoparticle size with specific volume fractions of 0.03 and 0.01 for nanoparticles and porosity in
Mg/Al2O3 nanocomposites as an example. It is seen that the contributions of fl and fporosity are fairly small;
the contributions of fd, fOrowan, and fHall-Petch-Zener are relatively large and in the sequence of
fd > fOrowan > fHall-Petch-Zener which increase monotonically and strongly with decreasing size of
nanoparticles. It is also observed that fOrowan and fHall-Petch-Zener increase more rapidly when the
nanoparticle size becomes very small (e.g., <50 nm). As seen from Figure 3 the change of the
improvement factors with the volume fraction of nanoparticles illustrates similar findings. That is, fl and
fporosity do not exhibit a significant influence, while fd, fOrowan, and fHall-Petch-Zener monotonically increase
with increasing volume fraction of nanoparticles in the same order of fd > fOrowan > fHall-Petch-Zener.

1
fl Mg/Al2O3
Vp= 0.03
fd
0.8
Improvement factor

fOrowan

0.6

0.4

0.2

0
0 100 200 300 400 500 600
Size of nanoparticles, nm

Figure 2. The improvement factors as a function of nanoparticle size with a volume fraction
of 0.03 and 0.01 for nanoparticles and porosity, respectively, in Mg/Al2O3 nanocomposites.
Materials 2015, 8 5145

1.5
fl Mg/Al2O3
fd dp=50 nm

1.2 fOrowan

Improvement factor
fporosity

0.9

0.6

0.3

0
0 0.02 0.04 0.06 0.08 0.1
Volume fraction of nanoparticles

Figure 3. The improvement factors as a function of volume fraction of nanoparticles with a


particle size of 50 nm and a volume fraction of 0.01 for porosity in Mg/Al2O3 nanocomposites.

To further indicate the changes of the yield strength of MMNCs with the volume fraction of
nanoparticles (Vp), the same set of data for the Al2O3-reinforced Mg nanocomposites tested at room
temperature in reference [46] is selected here, which was originally from Gupta and co-workers [11,12]
together with the relevant values from Brassell et al. [29]: ıym = 97 MPa, Gm = 16.5 GPa, b = 0.32 nm,
Įm= 28.4 × 10í6 (°C)í1, Įp = 7.4 × 10í6 (°C)í1, Tprocess = 300 °C, Ttest = 20 °C, M = 3.06, k = 70 MPa¥ȝm,
dp = 50 nm, K = 0.05 [54,56], and n = 1.94 (according to Equation (4a) for equiaxed particles where l is
equal to dp). In view of the change of porosity amount present in different composite fabrication
processes, e.g., 0.07–1.04 vol% [12], 0.25–1.15 vol% [14], 3–12 vol% [29], 0–1.18 vol% [34], up to
12.25 vol% [32], about 6 vol% [57], up to 13 vol% [33], and up to 12.45 vol% [58], three typical porosity
values of 1 vol%, 3 vol%, and 5 vol% were selected in the present model calculation.
To verify the validity of the present model with consideration of grain size and porosity, comparisons
were made between the model prediction and experimental data of three types of Mg-based
nanocomposites containing Al2O3, carbon nanotube (CNT), and Y2O3, respectively. Prior to doing so,
a comparison between the present model prediction and those in the literature [10,15,20,59–63] is shown
in Figure 4. It is seen that the present model shows a similar trend to those of other models, and lies
indeed in-between the other models.
Materials 2015, 8 5146

300

250

Yield Strength, MPa 200

150

100
Present model (P=1%) Present model (P=3%)
Present model (P=5%) Zhang and Chen's model [10]
Zhong's model [15] Ramkrishnan's model [20]
50 Geranmayeh's model [59] Tohgo's model [60]
Goh's model [61] Asgharzadeh's model [62]
Yazdi's model [63]
0
0 0.005 0.01 0.015
Volume fraction of nanoparticles

Figure 4. Comparison of the present model with different models reported in


literature [10,15,20,59–63].

First, the yield strength predicted for carbon nanotube (CNT)-reinforced Mg nanocomposites is shown
in Figure 5, together with the experimental data reported in refs. [14,16–18]. In the calculation of yield
strength from Equation (11) the following data for the CNT-reinforced Mg nanocomposites tested at
room temperature were used [11,12,29]: ıym = 97 MPa, Gm= 16.5 GPa, b = 0.32 nm, Įm = 28.4 × 10í6 (°C)í1,
Įp = í1.52 × 10í6 (°C)í1 (thermal contraction of CNT occurred, leading to a negative Į value [64]),
Tprocess = 350 °C, Ttest = 20 °C, K = 0.05 [54,56], n = 1.94 (for equiaxed particles), dp = 30 nm, P = 1%,
3% and 5%. As shown in Figure 5, due to the considerations of porosity weakening effect and grain
growth retardation by dispersed-particle-pinning of grain boundaries, the predicted yield strength for
P = 1%, 3% and 5% in the present model lies almost in-between those of Zhang and Chen [10,21] and
Ramakrishnan’s [20] model.
However, the predicted yield strength in a lower porosity level shows a better agreement with Zhang
and Chen’s model [10,21]. The first and second experimental data were slightly lower and higher than
that predicted from all methods including the present model, since all nanoparticles were assumed to be
spherical like Zhang and Chen model [10,21]. The scatter of the experimental data might also be related
to the waviness (or curviness) and agglomeration of CNTs [17,18,64–66], since the uniform dispersion
of such tiny nano-sized particles in nanocomposites still poses a significant challenge [65]. Next, four
data points showed a close agreement with the present model. It follows that incorporating the effects of
porosity and dispersed-particle-pinning of grain growth in the analytical models could improve
the predictability.
Materials 2015, 8 5147

250

200

Yield strength, MPa


150

100
Present Model (P=1%) Present Model (P=3%)
Present model (P=5%) Zhang and Chen's model [10]
50 Ramkrishnan's model [20] Experimental data-1 [14]
Experimental data-2 [16] Experimental data-3 [17]
Experimental data-4 [18]
0
0 0.001 0.002 0.003
Volume fraction of nanoparticles

Figure 5. Comparison of the present model (solid curves) with the experimental data
reported in [14,16–18], along with Zhang and Chen’s model [10] and Ramakrishnan’s
model [20], for CNT-reinforced Mg nanocomposites.

To further examine the trend of the yield strength of MMNCs, another prediction of yield strength
for Al2O3-reinforced Mg nanocomposites is shown in Figure 6, together with the experimental data
reported in References [11,12,66–69]. The same set of data for the Al2O3-reinforced Mg nanocomposites
tested at room temperature in reference [46] is first selected here, which was originally from Gupta and
co-workers [11,12] together with the relevant values from Brassell et al. [29]: ıym = 97 MPa, Gm = 16.5 GPa,
b = 0.32 nm, Įm= 28.4 × 10í6 (°C)í1, Įp = 7.4 × 10í6 (°C)í1, Tprocess = 300 °C, Ttest = 20 °C, Į = 0.3, Į1 = 0.35,
ȕ = 1.25, ij = 6 for Mg, k = 70 MPa¥ȝm, dp= 50 nm, K = 0.05 [54,56], n = 1.94 (according to
Equation (4a) for equiaxed particles), and P = 1%, 3% and 5%. Again, the yield strength predicted via
the present model is in fairly good agreement with the experimental data reported in
references [11,12,66–69]. Furthermore, with a higher amount of nanoparticles the dependence of the
yield strength on the volume fraction of porosity became stronger. The scatter of the experimental data
shown in Figure 6 would be associated with the processing techniques and conditions/parameters during
the fabrication of Mg-Al2O3 nanocomposites [11,12,66,67].
A further comparison involves Y2O3 particulate-reinforced Ti nanocomposite, as shown in Figure 7,
in which the following parameters were used in the calculation [70–73]: ıym = 450 MPa, Gm = 44.8 GPa,
b = 0.29 nm, Įm= 11.9 × 10í6 (°C)í1, Įp = 9.3 × 10í6 (°C)í1, Tprocess = 827 °C for particle size dp = 10, 12,
20, and 22 nm and Tprocess = 900 °C for particle size dp = 10, 40, and 170 nm, Ttest = 20 °C, K = 0.05 [54,56],
n = 1.94 (for equiaxed particles), P = 1%.
On the basis of the values of the weight fraction given in [70], the following converted values of
volume fraction/percent Vp = 0.25%, 0.27%, 0.38%, 0.41%, 0.54%, and 0.59% were utilized. The
straight line with a slope of m = 1 was drawn to show the value deviation. It is seen from Figure 7 that
Materials 2015, 8 5148

the present model with the consideration of both dispersed-particle-pinning of grain boundaries and
porosity predicts the yield strength of the Ti-Y2O3 nanocomposites fairly nicely, where a combined effect
of varying volume fraction of nanoparticles, thermo-mechanical treatment, and microstructures has been
taken into consideration.

250

200
Yield Strength, MPa

150

100
Present Model (P=1%)
Present Model (P=3%)
Present model (P=5%)
Zhang and Chen's model [10]
50 Ramkrishnan's model [20]
Experimental data-1 [11]
Experimental data-2 [12]
Experimental data-3 [66]
Experimental data-4 [67]
0
0 0.005 0.01 0.015
Volume fraction of nanoparticles

Figure 6. Comparison of the present model (solid curves) with the experimental data
reported in references [11,12,66–69], in conjunction with Zhang and Chen’s model [10] and
Ramakrishnan’s model [20], for Al2O3-reinforced Mg nanocomposites tested at 20 °C.

900
Predicted yield strength, MPa

800

700

600

500

400
400 500 600 700 800 900
Experimental yield strength, MPa

Figure 7. A comparison of the prediction via the present model with the experimental data
of Y2O3-reinforced Ti nanocomposites tested at 20 °C.
Materials 2015, 8 5149

4. Conclusions

While it was confirmed that the Orowan strengthening plays a significant role in MMNCs, the above
comparisons between the present model prediction and the experimental data reported in the literature
corroborate that both the positive and fairly strong effect of grain size refinement (or Hall-Petch
equation) and the negative effect of porosity should also be taken into consideration in predicting the
yield strength of MMNCs, although the effect of a small amount of porosity is rather small. The Zener
pinning effect of grain boundaries by the nano-sized particles has been integrated into the present
modelling as well. The proposed model has been validated using the experimental data of a number of
MMNCs containing different types of nano-sized particles, and it shows fairly good agreement.
This suggests that the positive effects of the Orowan strengthening mechanism, enhanced dislocation
strengthening mechanism and load-bearing effect of the reinforcement, and grain size effect, as well as
the negative weakening effect of porosity all need to be considered.

Acknowledgments

The authors would like to thank the financial support provided by the Natural Sciences and
Engineering Research Council of Canada (NSERC), Premier’s Research Excellence Award (PREA),
NSERC-DAS Award, AUTO21 Network of Centers of Excellence, and Ryerson Research Chair
program for providing financial support. The helpful discussion with Sanjeev Bhole is gratefully
acknowledged. This paper is presented at the 20th International Conference on Composite Materials
(ICCM20), Copenhagen, Denmark, July 19–24, 2015.

Author Contributions

F.A.M. and D.L.C. conceived, analyzed the data, and wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

1. Pollock, T.M. Weight loss with magnesium alloys. Science 2010, 328, 986–987.
2. Chu, S.; Majumdar, A. Opportunities and challenges for a sustainable energy future. Nature 2012,
488, 294–303.
3. Zheng, X.Y.; Lee, H.; Weisgraber, T.H.; Shusteff, M.; DeOtte, J.; Duoss, E.B.; Kuntz, J.D.;
Biener, M.M.; Ge, Q.; Jackson, J.A.; et al. Ultralight, ultrastiff mechanical metamaterials. Science
2014, 344, 1373–1377.
4. Nie, J.F.; Zhu, Y.M.; Liu, J.Z.; Fang, X.Y. Periodic segregation of solute atoms in fully coherent
twin boundaries. Science 2013, 340, 957–960.
5. Underwood, E. Models predict longer, deeper U.S. droughts: Future western “megadroughts” could
be worse than ever. Science 2015, 347, 707.
6. McNutt, M. Climate change impacts. Science 2013, 341, 435.
Materials 2015, 8 5150

7. Schmale, J.; Shindell, D.; von Schneidemesser, E.; Chabay, I.; Lawrence, M. Air pollution:
Clean up our skies. Nature 2014, 515, 335–337.
8. Luo, A.A. Magnesium: Current and potential automotive applications. JOM 2002, 54, 42–48.
9. Kainer, K.U. Magnesium—Alloys and Technology; Wiley-VCH Verlag GmbH & Co. KGaA:
Weinheim, Germany, 2003.
10. Zhang, Z.; Chen, D.L. Consideration of Orowan strengthening effect in particulate-reinforced metal
matrix nanocomposites: A model for predicting their yield strength. Scr. Mater. 2006, 54, 1321–1326.
11. Hassan, S.F.; Gupta, M. Development of high performance magnesium nanocomposites using
solidification processing route. Mater. Sci. Tech. 2004, 20, 1383–1388.
12. Wong, W.L.E.; Karthik, S.; Gupta, M. Development of hybrid Mg/Al2O3 composites with improved
properties using microwave assisted rapid sintering route. J. Mater. Sci. 2005, 40, 3395–3402.
13. Han, B.Q.; Dunand, D.C. Microstructure and mechanical properties of magnesium containing high
volume fractions of yttria dispersoids. Mater. Sci. Eng. A 2000, 277, 297–304.
14. Goh, C.S.; Wei, J.; Lee, L.C.; Gupta, M. Development of novel carbon nanotube reinforced
magnesium nanocomposites using the powder metallurgy technique. Nanotechnology 2006, 17, 7–12.
15. Zhong, X.L.; Wong, W.L.E.; Gupta, M. Enhancing strength and ductility of magnesium by
integrating it with aluminum nanoparticles. Acta Mater. 2007, 55, 6338–6344.
16. Goh, C.S.; Wei, J.; Lee, L.C.; Gupta, M. Effects of fabrication techniques on the properties of
carbon nanotube reinforced magnesium. Solid State Phenom. 2006, 111, 179–182.
17. Liu, S.Y.; Gao, F.P.; Zhang, Q.Y.; Zhu, X.; Li, W.Z. Fabrication of carbon nanotubes reinforced
AZ91D composites by ultrasonic processing. Trans. Nonferr. Met. Soc. China 2010, 20, 1222–1227.
18. Goh, C.S.; Wei, J.; Lee, L.C.; Gupta, M. Simultaneous enhancement in strength and ductility by
reinforcing magnesium with carbon nanotubes. Mater. Sci. Eng. A 2006, 423, 153–156.
19. Tun, K.S.; Gupta, M. Improving mechanical properties of magnesium using nano-yttria reinforcement
and microwave assisted powder metallurgy method. Comp. Sci. Technol. 2007, 67, 2657–2664.
20. Ramakrishnan, N. An analytical study on strengthening of particulate reinforced metal matrix
composites. Acta Mater. 1996, 44, 69–77.
21. Zhang, Z.; Chen, D.L. Contribution of Orowan strengthening effect in particulate-reinforced metal
matrix nanocomposites. Mater. Sci. Eng. A 2008, 483–484, 148–152.
22. Nardone, V.C.; Prewo, K.M. On the strength of discontinuous silicon carbide reinforced aluminum
composites. Scr. Metall. 1986, 20, 43–48.
23. Moghadam, A.D.; Schultz, B.F.; Ferguson, J.B.; Omrani, E.; Rohatgi, P.K.; Gupta, N. Functional
metal matrix composites-self-lubricating, self-healing, and nanocomposites-An outlook. JOM
2014, 66, 872–881.
24. Lurie, S.; Belov, P.; Volkov-Bogorodsky, D.; Tuchkova, N. Nanomechanical modeling of the
nanostructures and dispersed composites. Comp. Mater. Sci. 2003, 28, 529–539.
25. Lin, K.; Law, E.; Pang, S. Effects of interphase regions of particulate-reinforced metal matrix
nanocomposites using a discrete dislocation plasticity model. J. Nanomech. Micromech. 2014,
doi:10.1061/(ASCE)NM.2153-5477.0000098.
26. Meenashisundaram G.K.; Gupta, M. Synthesis and characterization of high performance low
volume fraction TiC reinforced Mg nanocomposites targeting biocompatible/structural applications.
Mater. Sci. Eng. A 2015, 627, 306–315.
Materials 2015, 8 5151

27. Park, M.S. An enhanced mean field material model incorporating dislocation strengthening for
particle reinforced metal matrix composites. J. Mech. Sci. Tech. 2014, 28, 2587–2594.
28. Bocchini, G.F. The influence of porosity on the characteristics of sintered materials. Int. J.
Pow. Metal. 1986, 22, 185–186.
29. Brassell, G.W.; Horak, J.A.; Butler, B.L. Effects of porosity on strength of carbon-carbon composites.
J. Comp. Mater. 1975, 9, 288–296.
30. Ray, S. Synthesis of cast metal matrix particulate composites. J. Mater. Sci. 1993, 28, 5397–5413.
31. Viswanathan, V.; Laha, T.; Balani, K.; Agarwal, A.; Seal, S. Challenges and advances in
nanocomposite processing techniques. Mater. Sci. Eng. R 2006, 54, 121–285.
32. Ahmad, S.N.; Hashim, J.; Ghazali, M.I. Effect of porosity on tensile properties of cast particle
reinforced MMC. J. Comp. Mater. 2005, 41, 575–589.
33. Molina, J.M.; Prieto, R.; Narciso, J.; Louis, E. The effect of porosity on the thermal conductivity of
Al-12 wt. % Si/SiC composites. Scr. Mater. 2009, 60, 582–585.
34. Zhong, X.L.; Gupta, M. Development of lead-free Sn-0.7Cu/Al2O3 nanocomposite solders with
superior strength. J. Phys. D Appl. Phys. 2008, 41, doi:10.1088/0022-3727/41/9/095403.
35. Thein, M.A.; Li, L.; On, L.M.; Liu, T.; Liu, Z.; Ringer, S.P. Creep behavior of in-situ formed aln
reinforced Mg–5Al nanocomposite at low temperatures and low stresses. Nanosci. Nanotechnol. Lett.
2009, 1, 204–207.
36. Paul, B.; Banerji, P. Grain structure induced thermoelectric properties in PbTe nanocomposite,
Nanosci. Nanotechnol. Lett. 2009, 1, 208–212.
37. Xiao, L.; Chen, D.L.; Chaturvedi, M.C. Shearing of gamma double prime precipitates and formation
of planar slip bands in Inconel 718 during cyclic deformation. Scr. Mater. 2005, 52, 603–607.
38. Hall, E.O. The deformation and ageing of mild steel: III discussion of results. Proc. Phys. Soc. B
1951, 64, 747–753.
39. Petch, N.J. The cleavage strength of polycrystals. J. Iron Steel Inst. 1953, 173, 25–28.
40. Van Aken, D.C.; Krajewski, P.E.; Vyletel, G.M.; Allison, J.E.; Jones, J.W. Recrystallization and
grain growth phenomena in a particle-reinforced aluminum composite. Metall. Mater. Trans. A
1995, 26, 1394–1405.
41. Doherty, R.D.; Hughes, D.A.; Humphreys, F.J.; Jonas, J.J.; Juul Jensen D.; Kassner, M.E.;
King, W.E.; McNelley, T.R.; McQueen, H.J.; Rollett, A.D. Current issues in recrystallization:
A review. Mater. Sci. Eng. A 1997, 238, 219–274.
42. Hassold, G.N.; Holm, E.A.; Srolovitz, D.J. Effects of particle size on inhibited grain growth.
Scr. Metall. Mater. 1990, 24, 101–106.
43. Stearns, L.C.; Harmer, M.P. Particle-inhibited grain growth in Al2O3-SiC: I, experimental results.
J. Am. Ceram. Soc. 1996, 79, 3013–3019.
44. Nishizawa, T.; Ohnuma, I.; Ishida, K. Examination of the Zener relationship between grain size and
particle dispersion. Mater. Trans. JIM 1997, 38, 950–956.
45. Nes, E.; Ryum, N.; Hunderi, O. On the Zener drag. Acta Metall. 1985, 33, 11–22.
46. Mirza, F.A.; Chen, D.L. An analytical model for predicting the yield strength of particulate-reinforced
metal matrix nanocomposites with consideration of porosity. Nanosci. Nanotechnol. Lett. 2012, 4,
794–800.
Materials 2015, 8 5152

47. Shao, J.C.; Xiao, B.L.; Wang, Q.Z.; Ma, Z.Y.; Yang, K. An enhanced FEM model for particle size
dependent flow strengthening and interface damage in particle reinforced metal matrix composite.
Comp. Sci. Technol. 2011, 71, 39–45.
48. Dunand, D.C.; Mortensen, A. On plastic relaxation of thermal stresses in reinforced metals.
Acta Metall. Mater. 1991, 39, 127–139.
49. Sankaranarayanan, S.; Nayak, U.P.; Sabat, R.K.; Suwas, S.; Almajid, A.; Gupta, M. Nano-ZnO
particle addition to monolithic magnesium for enhanced tensile and compressive response.
J. Alloys Compd. 2014, 615, 211–219.
50. Alizadeh, M.; Beni, H.A. Strength prediction of the ARBed Al/Al2O3/B4C nano-composites using
Orowan model. Mater. Res. Bull. 2014, 59, 290–294.
51. Akbarpour, M.R.; Salahi, E.; Hesari, F.A.; Kim, H.S.; Simchi, A. Effect of nanoparticle content on
the microstructural and mechanical properties of nano-SiC dispersed bulk ultrafine-grained Cu
matrix composites. Mat. Des. 2013, 52, 881–887.
52. Kim, C.-S.; Sohn, I.; Nezafati, M.; Ferguson, J.B.; Schultz, B.F.; Bajestani-Gohari, Z.;
Rohatgi, P.K.; Cho, K. Prediction models for the yield strength of particle-reinforced unimodal pure
magnesium (Mg) metal matrix nanocomposites (MMNCs). J. Mater. Sci. 2013, 48, 4191–4204.
53. Sanaty-Zadeh, A. Comparison between current models for the strength of particulate-reinforced
metal matrix nanocomposites with emphasis on consideration of Hall-Petch effect. Mater. Sci. Eng. A
2012, 531, 112–118.
54. Saada, G. Hall-Petch revisited. Mater. Sci. Eng. A 2005, 400–401, 146–149.
55. Li, J.C.M. Petch relation and grain boundary sources. Trans. Metall. Soc. AIME 1963, 227, 239–247.
56. Louchet, F.; Weiss, J.; Richeton, T. Hall-Petch law revisited in terms of collective dislocation
dynamics. Phys. Rev. Lett. 2006, 97, doi:10.1103/PhysRevLett.97.075504.
57. Liu, T.M.; Chao, C.G. Effect of magnesium on mechanical properties of alumina-fiber-reinforced
aluminum matrix composites formed by pressure infiltration casting. Mater. Sci. Eng. A 1993, 169,
79–84.
58. Ahmad, S.N.; Hashim, J.; Ghazali, M.I. The effects of porosity on mechanical properties of cast
discontinuous reinforced metal- matrix composite. J. Comp. Mater. 2005, 39, 451–466.
59. Geranmayeh, A.R.; Mahmudi, R.; Kangooie, M. High-temperature shear strength of lead-free
Sn-Sb-Ag/Al2O3 composite solder. Mater. Sci. Eng. A 2011, 528, 3967–3972.
60. Tohgo, K.; Itoh, Y.; Shimamura, Y. A constitutive model of particle-reinforced composites taking
account of particle size effects and damage evolution. Comp. Part A 2010, 41, 313–321.
61. Goh, C.S.; Gupta, M.; Wei, J.; Lee, L.C. Characterization of high performance Mg/MgO
nanocomposites. J. Comp. Mater. 2007, 41, 2325–2335.
62. Asgharzadeh, H.; Simchi, A.; Kim, H.S. Hot deformation of ultrafine-grained Al6063/Al2O3
nanocomposites. J. Mater. Sci. 2011, 46, 4994–5001.
63. Zehtab Yazdi, A.; Bagheri, R.; Zebarjad, S.M.; Razavi Hesabi, Z. Incorporating aspect ratio in a
new modeling approach for strengthening of MMCs and its extension from micro to nano scale.
Adv. Comp. Mater. 2010, 19, 299–316.
64. Kwon, Y.K.; Berber, S.; Tomanek, D. Thermal contraction of carbon fullerenes and nanotubes.
Phys. Rev. Lett. 2004, 92, 015901–015904.
Materials 2015, 8 5153

65. Morsi, K.; Esawi, A. Effect of mechanical alloying time and carbon nanotube (CNT) content on the
evolution of aluminum (Al) CNT composite powders. J. Mater. Sci. 2007, 42, 4954–4959.
66. Wong, W.L.E.; Gupta, M. Improving overall mechanical performance of magnesium using
nano-alumina reinforcement and energy efficient microwave assisted processing route. Adv. Eng.
Mater. 2007, 9, 902–909.
67. Hassan, S.F.; Tan, M.J.; Gupta, M. High-temperature tensile properties of Mg/Al2O3 nanocomposite.
Mater. Sci. Eng. A 2008, 486, 56–62.
68. Hassan, S.F.; Gupta, M. Effect of length of Al2O3 particulates on microstructural and tensile
properties of elemental Mg. Mater. Sci. Eng. A 2006, 425, 22–27.
69. Hassan, S.F.; Gupta, M. Development of high performance magnesium nano-composites using
nano- Al2O3 as reinforcement. Mater. Sci. Eng. A 2005, 392, 163–168.
70. De Castro, V.; Leguey, T.; Muñoz, A.; Monge, M.A.; Pareja, R. Relationship between hardness and
tensile tests in titanium reinforced with yttria nanoparticles. Mater. Sci. Eng. A 2005, 400–401,
345–348.
71. Callister, W.D.; Rethwisch, D.G. Materials Science and Engineering: An Introduction, 9th ed.;
John Wiley and Sons: New York, NY, USA, 2014.
72. Bever, M.B. Encyclopedia of Materials Science and Engineering; Pergamon Press: New York, NY,
USA, 1986.
73. Kingery, W.D. Introduction to Ceramics, 2nd ed.; John Wiley and Sons: New York, NY, USA,
1976.

© 2015 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article
distributed under the terms and conditions of the Creative Commons Attribution license
(http://creativecommons.org/licenses/by/4.0/).
Materials Science & Engineering A 575 (2013) 65–73

Contents lists available at SciVerse ScienceDirect

Materials Science & Engineering A


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

Low cycle fatigue of a rare-earth containing extruded magnesium alloy


F.A. Mirza a, D.L. Chen a,n, D.J. Li b, X.Q. Zeng b,1
a
Department of Mechanical and Industrial Engineering, Ryerson University, 350 Victoria Street, Toronto, Ontario, Canada M5B 2K3
b
The State Key Laboratory of Metal Matrix Composites, School of Materials Science and Engineering, Shanghai Jiao Tong University, 800 Dongchuan Road,
Shanghai 200240, PR China

art ic l e i nf o a b s t r a c t

Article history: The application of ultra-lightweight magnesium alloys inevitably involves fatigue resistance under cyclic
Received 26 January 2013 loading. The present study was aimed at evaluating strain-controlled cyclic deformation behavior and the
Received in revised form relevant effect of microstructure in a rare-earth (RE) element containing extruded Mg–10Gd–3Y–0.5Zr
12 March 2013
(GW103K) alloy. The microstructure of this alloy consisted of fine equiaxed grains with an average grain
Accepted 14 March 2013
size of about 12 μm and a large number of RE-containing precipitates. Unlike the RE-free extruded
Available online 27 March 2013
magnesium alloys, this alloy exhibited essentially cyclic stabilization and symmetrical hysteresis loops
Keywords: without tension–compression asymmetry due to the presence of the relatively weaker texture and the
Magnesium alloy suppression of twinning activities arising from the fine grain size and especially RE-containing precipitates.
Cyclic deformation
A detailed analysis for understanding the obstructive role of the precipitate to twinning has been presented.
Low cycle fatigue
While this alloy had a lower cyclic strain hardening exponent than the RE-free extruded magnesium alloys,
Twinning–detwinning
Texture it had a longer fatigue life which can also be described by the Coffin–Manson law and Basquin's equation.
Tension–compression yield asymmetry Fatigue crack was observed to initiate from the specimen surface with some cleavage-like facets at the
initiation site. Crack propagation was basically characterized by fatigue striations in conjunction with
secondary cracks.
& 2013 Elsevier B.V. All rights reserved.

1. Introduction the availability of slip systems and leads to preferred orientations


(or textures) in the deformation process [14–16]. The commonly
Due to the increasing global energy demand and high con- reported tension–compression yield asymmetry and mechanical
sciousness of environmental protection in recent years [1,2], anisotropy in wrought magnesium alloys are the main result of
lightweighting has become a key strategy in the automotive and those preferred orientations [17–19]. For the structural compo-
aerospace industries [1,3,4]. It has been reported that the fuel nents subjected to alternating cyclic loading, such a mechanical
efficiency of passenger vehicles can be enhanced by 6–8% for each anisotropy could result in irreversibility of cyclic deformation
10% reduction in weight [5]. Magnesium alloy, as the lightest which may exert an unfavorable influence on the material perfor-
structural metallic material, has been increasingly used in the mance. This problem could be conquered through texture mod-
transportation industry to reduce the weight of motor vehicles [6]. ification via an alloy composition adjustment. It has recently been
The structural application of magnesium alloys involves unavoid- reported that magnesium alloys containing rare-earth (RE) ele-
ably fatigue and cyclic deformation characteristics due to the fact ments can develop more random textures during hot extrusion
that structural components in the vehicles experience dynamic [20–22]. The addition of RE elements in magnesium alloys leads to
loading, which results in the occurrence of fatigue failure [7–9]. a decrease of the overall texture sharpness or intensity, allowing
Hence, an understanding of fatigue and cyclic deformation of the activation of easy basal slip to a higher extent, compared to the
magnesium alloy is critical for the design and durability evaluation RE-free wrought magnesium alloys [21,23,24]. For example, Stan-
of engineering components. ford and Barnett [24] reported that the anisotropy of yield strength
Wrought alloys exhibit superior fatigue properties than casting of a rolled Mg–Zn–RE alloy was reversed as compared with the
counterparts and are appropriate for studying the intrinsic fatigue traditional Mg–Zn alloy. Despite the fact that the addition of RE
mechanisms of magnesium alloys [10–13]. However, the hexago- elements sheds some light on the alteration in the mechanical
nal close-packed (hcp) crystal structure of magnesium alloys limits anisotropy, the potential advantage of such RE–Mg alloys as
structural components under cyclic loading condition has not
n
been well appreciated. The earlier studies were conducted mainly
Corresponding author. Tel.: +1 416 979 5000x6487; fax: +1 416 979 5265.
E-mail addresses: dchen@ryerson.ca (D.L. Chen),
on the high cyclic fatigue properties of the RE-containing magne-
xqzeng@sjtu.edu.cn (X.Q. Zeng). sium alloys [14,25–27], such as Yang et al. [14] evaluated very
1
Tel.: +021 5474 2911; fax: +021 3420 2794. high-cycle fatigue behavior of an extruded Mg–12Gd–3Y–0.5Zr

0921-5093/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2013.03.041
66 F.A. Mirza et al. / Materials Science & Engineering A 575 (2013) 65–73

magnesium alloy at an ultrasonic frequency of 20 kHz and a stress 3. Results


ratio of R¼ −1 and Yang et al. [27] reported the effect of addition of
different amounts of RE elements on high-cycle fatigue behavior of 3.1. Microstructure and texture
die-cast magnesium alloy AZ91D. Studies on the strain-controlled
low cycle fatigue behavior of RE-containing magnesium alloys are Fig. 1(a) shows a typical microstructure of as-extruded GW103K
very limited [28,29]. It is unclear how and to what extent the RE alloy where the extrusion direction (ED) is indicated. It is seen that
elements can change the tensile–compressive yield asymmetry uniform equiaxed grains with an average grain size of about 12 μm
which generally occurs in the RE-free extruded magnesium alloys, were obtained in this alloy due to the occurrence of dynamic
whether RE–Mg alloys exhibit cyclic hardening or softening, and recrystallization (DRX) in the hot extrusion process at 400 1C. The
via what mechanisms the RE elements affect the cyclic deforma- grain size was fairly small in comparison with the common
tion behavior and fatigue life. Thus, an understanding about the extruded magnesium alloys, such as AZ31 and AM30 [8,11–13,30].
influence of grain size and texture on the cyclic deformation This was due to the role of added RE elements and zirconium where
characteristics and twin formation due to the addition of RE Zr mainly restricted the grain growth [31]. A similar role of grain
elements is necessary. The objective of the present work was to refinement by RE elements, e.g., gadolinium [32], yttrium [33],
evaluate the cyclic deformation behavior of an extruded Mg– cerium [34], neodymium [35], erbium [36], has been reported as
10Gd–3Y–0.5Zr (GW103K) alloy, and determine the fatigue life well. Besides, as seen from Fig. 1(a) no twins were present in the
under varying strain amplitudes, with particular attention to the un-deformed samples. Fig. 1(b) shows a typical SEM back-scattered
effect of the RE element addition on the twin formation and twin- electron image of as-extruded sample where many rare-earth
particle interaction. containing particles can be seen, as indicated by arrows. To confirm
the presence of RE-rich particles in the as-extruded alloy, EDS line
scan was performed as shown in Fig. 1(c) and (d), where the middle
2. Material and experimental procedure rectangle-like particle contained Gd and the right circular particle
contained Zr. Similar microstructure was presented in [37] and
The material investigated was a recently developed extruded three different phases were observed: α-Mg solid solution matrix
magnesium alloy with the following composition (wt%): 10Gd, 3Y, phase with supersaturated Gd+Y elements, (Gd+Y)-rich eutectic
0.5Zr, and Mg (balance). The received alloy bars with a diameter compound which had a higher Gd+Y content than the matrix, and
of 20 mm were extruded at 400 1C with an extrusion ratio of 9:1. intracrystalline Zr-rich cores. Similar microstructures were also
Microstructural examinations were performed using an optical reported in as-extruded Mg–11.90Gd–0.81Y–0.44Zr alloy by Zhang
microscope (OM), and scanning electron microscope (SEM) JSM- et al. [38], as-cast samples of Mg–7Gd–3Y–0.4Zr (GW73K) alloy by
6380LV equipped with Oxford energy dispersive X-ray spectroscopy Liang et al. [39], and cast Mg–10Gd–3Y–Zr alloy by Wang et al. [40].
(EDS) system. Standard metallographic sample preparation techni- Fig. 2 shows a (0002) pole figure of the extruded GW103K alloy
ques were used with an etchant based on an acetic picral solution evaluated using MTEX software. A relatively weak basal texture (tilted
containing 4.2-g picric acid, 10-ml acetic acid, 10-ml H2O, and 70-ml towards the radial direction (RD)) with a maximum intensity of 2.1 m.
ethanol. The texture was determined by measuring incomplete pole r.d. (multiples of random distribution) after de-focusing correction was
figures between Ψ¼01 to 751 in the back reflection mode using a observed, where the c-axes of most grains were aligned perpendicular
PANalytical X-ray diffractometer with Cu Kα radiation at 45 kV and to the ED. The presence of such a weaker texture, in comparison with
40 mA and analyzed using MTEX software, and defocusing due to the extruded AM30 [41] and rolled AZ31 [42], was another benefit of
the rotation of sample holder was corrected using experimentally the RE elements added into magnesium alloys, as also reported by
determined data obtained from the diffraction of magnesium Stanford and Barnett [24] who observed that microalloying with RE
powders received from Magnesium Electron. Sub-sized fatigue elements can weaken texture in the forming process. Similar weaker
samples in accordance with ASTM E8 standard were machined with texture in Mg-based alloys with the addition of RE elements was
the loading axis parallel to the extrusion direction (ED). The samples reported [14,20,43,44]. In the current extruded GW103K alloy, grain
had a gauge length of 25 mm (or a parallel length of 32 mm) and a boundary nucleation, coalescence of low-angle boundaries and shear
width of 6 mm. The thickness of the samples was 6 mm as well. The band assisted nucleation are absent (as seen in Fig. 1(a)) due to the
gage section of fatigue samples was ground progressively along the presence of a large number of RE-containing particles (Fig. 1(b)).
loading direction with emery papers up to grit number of 600 to Zhu et al. [43] considered that only particle-stimulated nucleation was
remove the machining marks and to achieve a smooth surface. a factor influencing texture weakening. Since the fine particles cannot
Strain-controlled, pull–push type fatigue tests were conducted deform compatibly with the matrix during extrusion, stresses could
using a computerized Instron 8801 fatigue testing system at a zero concentrate easily near the precipitates and subsequent plastic
mean strain (strain ratio Re ¼−1, completely reversed strain cycle), relaxation would occur to relieve the stress build-up around the
a constant strain rate of 1 " 10−2 s−1 and room temperature of precipitates, resulting in local lattice rotations in the regions adjacent
25 1C. Triangular loading waveform was applied during the tests. to the precipitates [14,45]. Indeed, Humphreys [46] observed that the
Low cycle fatigue tests were performed at total strain amplitudes high RE content with a mixed addition resulted in the precipitation of
of 0.2%, 0.4%, 0.6%, 0.8%, 1.0%, and 1.2%, and at least two samples a large number of the second-phase particles, which can effectively
were tested at each level of the strain amplitudes. The strain- hinder the movement of dislocations and build up a stress concentra-
controlled tests at lower strain amplitude levels were continued tion near the precipitates, in turn promote the particle-stimulated
up to 10,000 cycles, then the tests were changed to load control at nucleation of dynamic recrystalliazation and diminish the basal
a frequency of 50 Hz using sine cyclic waveform. For the sake of texture. In addition, the weakening of the texture was also associated
comparison, some samples of extruded RE-free AM30 magnesium with the appearance of deformation bands containing twins and the
alloy (with a composition of 3.4 wt% Al, 0.33 wt% Mn, 0.16 wt% Zn, restriction of grain growth [47].
0.0026 wt% Fe, 0.0006 wt% Ni, 0.0008 wt% Cu and balance Mg)
were also tested. The fracture surfaces of fatigued specimens were
examined via scanning electron microscope (SEM) to identify 3.2. Hysteresis loops and tensile properties
fatigue crack initiation sites and propagation characteristics.
The residual twins in the regions near the fracture surface were Fig. 3 shows typical stress–strain hysteresis loops of the first,
observed as well. second, and mid-life cycles at a total strain amplitude of 1.2% and
F.A. Mirza et al. / Materials Science & Engineering A 575 (2013) 65–73 67

Precipitate
Precipitate

ED

Intensity, Counts x1000


Mg
3

1 Zr
Gd
0
0 5 10 15 20 25 30
Distance , m
Fig. 1. Microstructure of extruded GW103K alloy, (a) OM image, and (b) SEM back-scattered electron image, (c) SEM back-scattered electron image indicating an EDS line
scan position, and (d) the corresponding EDS line scan results.

360

270

180
ED
90
Stress, MPa

RD 0

-90

AM30-1st cycle
-180 AM30-2nd cycle
AM30-Mid-life cycle
-270 GW103K-1st cycle
GW103K-2nd cycle
GW103K-Mid-life cycle
Fig. 2. (0002) pole figure of extruded GW103K alloy, where ED stands for the -360
extrusion direction and RD indicates the radial direction. -1.5 -1 -0.5 0 0.5 1 1.5
Total strain, %
strain ratio of Re ¼−1 for the extruded GW103K and AM30 alloys, Fig. 3. Typical stress-strain hysteresis loops of different cycles at a total strain
respectively. The yield strength at a strain rate of 1 " 10−2 s−1 amplitude of 1.2% and strain ratio of Re ¼ −1 for the extruded AM30 and Mg–10Gd–
3Y–0.5Zr (GW103K) alloy, respectively.
could be easily determined from the initial quarter of the first-
cycle hysteresis loops. The present GW103K alloy exhibited a
higher tensile yield strength than the extruded AM30 alloy, i.e., similar to those of face-centered cubic (fcc) metals (e.g., Al, Cu, and
#240 MPa vs. # 200 MPa [41]. This reflected an obvious beneficial Ni) as a result of the dislocation slip-dominated deformation in
role of RE element addition in the magnesium alloy. The obtained most materials [49]. Furthermore, the maximum and minimum
tensile properties were in agreement with those reported in an peak stresses showed no significant change among the first,
extruded Mg–8Al–xRE alloy [48], where the improved tensile second, and mid-life cycle loops in the GW103K alloy. In contrast,
properties due to the addition of RE elements were presented as the situation in the extruded AM30 alloy was very different, where
well. In particular, it is seen from Fig. 3 that, unlike the extruded the hysteresis loops were skewed although the mid-life hysteresis
RE-free AM30 alloy, the current GW103K alloy exhibited nearly loop appeared slightly less distorted than the first and second
symmetrical hysteresis loops in shape, which were somewhat cycle hysteresis loops. This was predominantly due to the activity
68 F.A. Mirza et al. / Materials Science & Engineering A 575 (2013) 65–73

of twinning in compression in the descending phase and subse- a physical quantity that results in several damaging processes and
quent detwinning in the ascending phase [8,9,11–13,30,50–52]. influences the internal microstructure which is closely related to the
strain resistance and eventually the fatigue life [12]. The variation of
the plastic strain amplitude ðΔεp =2Þ during cyclic deformation is
3.3. Fatigue behavior
shown in Fig. 5 for different total strain amplitudes. Again, it is seen
that cyclic stabilization basically occurred in the present RE-
Fig. 4 shows the evolution of stress amplitudes with respect to
containing GW103K alloy at all strain amplitudes applied up to as
the number of cycles at different applied strain amplitudes on a
high as 1.2%. As a result, the addition of RE elements was able to
semi-log scale. As the total strain amplitude increased, the stress
change significantly the cyclic deformation behavior of the magne-
amplitude increased and the fatigue life of the material decreased.
sium alloy.
Unlike the RE-free extruded AM30 [11] and AZ31 [8,11–13] where
The fatigue life (i.e., the number of cycles to failure, Nf) as a
cyclic stabilization occurred only at lower strain amplitudes of about
function of the applied total strain amplitudes ðΔεt =2Þ of the
0.1% and 0.2%, the GW103K alloy exhibited cyclic stabilization until
extruded GW103K alloy is plotted in Fig. 6, along with the
failure up to a strain amplitude of 1.0%. Even at a strain amplitude of
experimental data reported in the literature for various extruded
1.2%, only slight change, i.e., an initial slight cyclic hardening within
magnesium alloys [11,12,42,53–55] for comparison. Run-out data
the first three cycles and then minor cyclic softening, could be seen
points were indicated by arrows pointing horizontally at or more
from Fig. 4. Such a change was indeed trivial, as also seen from the
than 107 cycles. The alloy basically showed an improved fatigue life
evolution of hysteresis loops mentioned above (Fig. 3). In the low
than the RE-free extruded magnesium alloys [8,11,12,42,53–55].
cycle fatigue tests, plastic strain amplitude has been considered as
As described in refs. [8,9,11,12,56–59] the total strain amplitude
400
Total strain amplitudes 1.2% 1.4
GW103K-present study
1.0%
AM30 [11]
300 0.8% 1.2
AZ31 [12]
Stress amplitude, MPa

0.6%
AZ31 [42]
0.4% 1.0
ZK60 [53]
0.2% AZ61-ED [54]
200 0.8
εt/2, %

AZ61-ETD [54]
AM30 [55]
0.6

100
0.4

0.2
0
1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 0.0
Number of cycles, N 1E+1 1E+3 1E+5 1E+7
Number of cycles to failure, Nf
Fig. 4. Stress amplitude vs. the number of cycles at different total strain amplitudes
applied. Fig. 6. Total strain amplitude as a function of the number of cycles to failure for the
extruded GW103K alloy, in comparison with the data reported in the literature for
various extruded Mg alloys.

0.75
Total strain amplitudes 1.2% -1
1.0%
Total
0.8%
Plastic strain amplitude, %

0.55 -1.5 Plastic


0.6% YT = -0.1888X T - 1.5297
R² = 0.9762 Elastic
0.4%
-2
0.2%
0.35
-2.5
Log

YP = -0.4449XP - 1.3191
R² = 0.975
-3
0.15
Y E= -0.1087X E - 1.9342
-3.5 R² = 0.9555

-0.05 -4
1E+0 1E+1 1E+2 1E+3 1E+4 1E+5 1 2 3 4 5 6
Number of cycles, N Log ( f )

Fig. 5. Plastic strain amplitude vs. the number of cycles at different total strain Fig. 7. Evaluation of fatigue parameters in the form of log elastic, plastic and total
amplitudes applied. strain amplitudes vs. log number of reversals to failure, respectively.
F.A. Mirza et al. / Materials Science & Engineering A 575 (2013) 65–73 69

could be expressed as elastic strain amplitude and plastic strain fatigue life or the number of cycles to failure, s′f is the fatigue
amplitude, i.e., strength coefficient, b is the fatigue strength exponent, ε′f is the
fatigue ductility coefficient, and c is the fatigue ductility exponent.
Δεt Δεe Δεp s′f ð2Nf Þb The elastic strain component is referred to as Basquin's equation
¼ þ ¼ þ ε′f ð2N f Þc ; ð1Þ
2 2 2 E and the plastic strain component is known as the Coffin–Manson
where E is the Young's modulus (for the present alloy the average relation. Fig. 7 shows the elastic, plastic, and total strain amplitudes
value obtained during fatigue testing was # 44.5 GPa), Nf is the plotted as a function of the number of reversals to failure. The
values of the strain amplitudes were taken from the mid-life cycles,
Table 1 even though cyclic stabilization or saturation was present in this
Low cycle fatigue parameters obtained for the extruded GW103K alloy. alloy at all the strain amplitudes up to the applied strain amplitude
of 1.2% (Figs. 3–5). The fatigue life parameters obtained on the basis
Low cycle fatigue parameters Extruded GW103K
of Eq. (1) are summarized in Table 1. It is seen that while the
Cyclic strain hardening exponent, n′ 0.20 obtained fatigue parameters were well within the range in other
Cyclic strength coefficient, K′, MPa 754 fatigued magnesium alloys reported in the literature [8,11,12,50],
Fatigue strength coefficient, s′f, MPa 521 the cyclic strain hardening exponent n′ of GW103K alloy was lower
Fatigue strength exponent, b −0.11 than that of extruded AZ31 alloy [8,11,12]. This corresponded well
Fatigue ductility coefficient, ε′f 0.05
Fatigue ductility exponent, c −0.44
to the higher yield strength and cyclic stabilization characteristics of
this alloy (Figs. 4 and 5).

Fig. 8. SEM images of overall fracture surfaces of the extruded specimens fatigued at a total strain amplitude of (a) 0.4% and (b) 1.0%.

Fig. 9. SEM micrographs of the fracture surface near crack initiation of the extruded specimens fatigued at a total strain amplitude of 0.4% ((a) secondary electron image and
(b) back-scattered electron image) and (c) 1.0% (secondary electron image).
70 F.A. Mirza et al. / Materials Science & Engineering A 575 (2013) 65–73

Fig. 10. SEM micrographs of the fatigue crack propagation region of the extruded specimens fatigued at a total strain amplitude of (a) 0.4% and (b) 1.0%.

3.4. Fractography process (rolling or extrusion) of magnesium alloys makes most


grains orientated in such a manner that their basal planes are
Fig. 8 shows an overall view of typical fracture surfaces at a lower nearly parallel to the rolling/extrusion direction. When a load is
magnification for the specimens tested at a strain amplitude of 0.4% applied in the extrusion or rolling direction, the activation stress is
and 1.0%, respectively. It is seen that fatigue crack initiated basically relatively low for the extension twinning (in compression) or basal
from the specimen surface. On a close examination at a higher slip (in tension), and is relatively high for the prismatic and
magnification (Fig. 9(a)), fatigue crack indeed initiated from near- pyramidal slip systems in magnesium alloys at room temperature
surface defects. From a back-scattered electron image (Fig. 9(b)), [63]. Then under cyclic loading a fairly large portion of the cyclic
which was taken at the same location as that of Fig. 9(a), it is seen strain would be undertaken by the twinning–detwinning activities
that the RE-containing precipitates were uniformly distributed on the [11,12]. As seen from the near fracture surface area in Fig. 11(a) and (b),
fracture surface. The initiation site at a strain amplitude of 1.0% had dense residual twins appeared near the fracture surface of AM30 alloy.
more obvious features of cleavage-like facets (Fig. 9(c)) than that at a The formation of these twins was attributed to an insufficient number
strain amplitude of 0.4% (Fig. 9(a) and (b)). Fatigue crack propagation of slip systems in magnesium alloys deformed at room temperature
was basically characterized by fatigue striations observed at higher arising from the hcp structure with a low crystal symmetry, in
magnifications (Fig. 10(a) and (b)), in conjunction with some sec- conjunction with the presence of strong texture as discussed above.
ondary cracks. In addition, the fatigue striations were perpendicular However, twins could barely be visible in the near fracture
to the crack propagation direction. These characteristics were more surface area of GW103K alloy (Fig. 11(c) and (d)). The absence of
obvious at higher strain amplitudes, as see in Fig. 10(b). twins suggested that twinning was not a dominating deformation
mode in the cyclic deformation of GW103K alloy, and the basal slip
of dislocations would become a main deformation mechanism as
4. Discussion corroborated by the nearly symmetrical hysteresis loops (Fig. 3),
somewhat like the situation of the common fcc metals. This was
4.1. Texture mainly due to the effect of grain size and precipitates in the
GW103K alloy, to be discussed in the following section.
In polycrystalline alloys, the yield asymmetry is directly related to
the presence of crystallographic texture which dominates the orienta- 4.3. Effect of grain size and precipitates
tion of slip or twinning planes and directions relative to the externally
applied stress [60]. Normally the yield asymmetry is less obvious in An important factor that influences the twinning in tension-
cast magnesium alloys which tend to have weak textures compared compression cyclic deformation is grain size. As seen from Figs. 1(a),
with wrought magnesium alloys [24]. Yin et al. [61] also reported that and 11(c), the grains in the GW103K alloy were equiaxed and fairly
randomly textured casting samples exhibit almost no yielding asym- small due to the presence of RE elements and zirconium, in
metry. As mentioned earlier in Section 3.1, the present RE-containing comparison with the extruded Mg alloys, such as AZ31 and AM30
GW103K alloy showed a relatively weak texture (Fig. 2) due to the [8,11–13,30]. It was reported that grain size had a significant effect
presence of second-phase particles and finer grain size which were on the tendency of twinning since the energy required to form twin
responsible for the nearly symmetrical hysteresis loops in Fig. 3. interfaces was particularly high in the fine-grained magnesium
Generally, the initial deformation asymmetry behavior of metals is alloy [64]. In addition, there is a general agreement that a small
associated with the Basuschinger effect [56]. From Fig. 3 it is clearly grain size inhibits twinning somewhat like the slip of dislocations
seen that the hysteresis loops of AM30 alloy showed a strong [41,51,52,57–61]. Generally, the grain size dependence of the
Basuschinger-like effect, which concurred with the results reported activation of deformation mechanisms in polycrystals could be
by other investigators [11,12,56,62]. The ratio of the compressive-to- described in terms to the Hall–Petch relationship ðs ¼ so þ kd
−1=2
Þ
tensile yield stress was # 0.55 for AM30 alloy, whereas it was almost [56], where s is the yield strength, so is a friction stress for the
1 for GW103K alloy, indicating that there was almost no difference in movement of dislocations, and d is the average grain size. The slope
the tensile and compressive yielding due to the microalloying with RE k is called as the “Hall–Petch strength coefficient”. The effect of the
elements (Fig. 3). Thus, it could be concluded that the addition of RE grain size or the content of alloying elements on twinning was
elements played an important role in eliminating or diminishing the studied using Hall–Petch analysis by Bohlen et al. [65], who also
yield asymmetry of the material via texture randomization. observed the distinctive tension/compression asymmetry in the
yield behavior of magnesium alloy, similar to the hysteresis loops of
4.2. Twinning–detwinning during cyclic deformation AM30 alloy shown in Fig. 3. While the asymmetry in the tensile and
compressive yield strengths reflects a different contribution of
Twinning has an important effect on the plastic deformation twinning and slip to the macroscopic strain, the difference between
of magnesium alloys [8–13,15,30,41,59]. In general, the forming the tensile and compressive yield strengths generally decreases
F.A. Mirza et al. / Materials Science & Engineering A 575 (2013) 65–73 71

Fig. 11. Optical micrographs and SEM images in the areas near the fracture surface at a strain amplitude of 1.2%, showing the distribution of residual twins in the fatigued
samples of AM30 alloy ((a) and (b)) and GW103K alloy ((c) and (d)), respectively.

with decreasing grain size [65]. Thus, the grain size has a strong [0001]
influence on the yield stress associated with both the slip of (Matrix)
dislocations and twinning. The twinning stress was indeed reported {10 1 2}
twinning plane K1
to increase with decreasing grain size more rapidly than the stress
required to activate slip [61,66,67]. It follows that for the fine- Matrix 2 [10 1 1]
grained alloys dislocation slip would dominate the deformation no *[10 1 1] 7.4° Twin
2
matter whether it is in compression or in tension, and the yield
asymmetry would be absent or reduced in extent due to the 86.3° Shear
suppression of twinning activities, as also seen from the hysteresis
loops of GW103K alloy in Fig. 3.
In addition to the effect of grain size, it has also been reported Precipitate 86.3°
that the presence of precipitates modifies the process of twinning
in magnesium alloys [68]. The interaction between twinning and Twin basal plane
precipitates has been considered to arise primarily from the
difficulty experienced by migrating twin boundaries during the
twin propagation through a densely distributed precipitates [60].
43.15°
To better understand the interaction between a twin and a plate- Matrix [10 1 0]
shaped precipitate, a schematic illustration showing the obstruc- (Matrix)
tion of a precipitate/particle to twinning is plotted in Fig. 12. Let us Matrix basal plane
first establish a coordinate system with the Y-axis being parallel to Fig. 12. A schematic illustration of the interaction between a f1 0 1 2g twin and a
the c-axis [0001] of an hcp unit cell in the magnesium matrix and plate-shaped precipitate in the GW103K alloy. (For interpretation of the references
the X-axis being oriented in the direction of ½1 0 1 0( which is to color in this figure, the reader is referred to the web version of this article.)

positioned on the basal plane. In such a way the f0 0 0 2g basal


plane of the magnesium matrix becomes perpendicular to the sur-
face of “paper” (or “computer screen”) and parallel to the X-axis, as perpendicular to the surface of the “paper” or “computer screen”
indicated in Fig. 12. Then the precipitate can be plotted vertically as well) or the twinning direction η1 ½1 0 1 1( residing on the
to the X-axis (the basal plane of the magnesium matrix), since it twinning plane K1 to be 43.151, as indicated in Fig. 12. In response
was observed to form on the prismatic planes of the magnesium to the twinning shear and the symmetry requirement to generate
matrix in a triangular arrangement [37]. In the magnesium alloys, a mirror image of the atoms inside the twin versus those in the
f1 0 1 2g extension twinning is normally activated when a com- matrix with respect to the invariant twinning plane K1, an
pressive stress is applied perpendicular to the [0001] c-axis or operation equivalent to the rotation of the lattice atoms within
parallel to the f0 0 0 2g basal plane [9,69]. Using the magnesium the twin (i.e., the portion in-between the two red-colored lines in
lattice parameters of a ¼0.32 nm and c¼ 0.52 nm [70], one can Fig. 12) has to take place under an external (or internal) driving
readily calculate the angle between the f0 0 0 2g basal plane and force. Such a change is similar to re-orienting the basal planes of
the f1 0 1 2g twinning plane K1 (which is an invariant plane being the matrix (i.e., the fine horizontal lines parallel to the X-axis) to
72 F.A. Mirza et al. / Materials Science & Engineering A 575 (2013) 65–73

those indicated by the fine parallel lines inside the twin, which are strain amplitudes at room temperature. The following conclusions
inclined at an angle of 86.31 to the fine horizontal lines, as shown can be drawn from this investigation:
in Fig. 12. Due to the unique nature of twin characterized by the
perfectly symmetrical (or mirror) image of atoms with respect to 1. The microstructure of the RE-containing GW103K alloy con-
the twinning plane, the inclined fine parallel lines within the twin sisted of fine recrystallized and equiaxed grains with an
must have an atomic arrangement identical to that represented by average grain size of about 12 μm and a large number of RE-
the fine horizontal lines. Therefore, the inclined fine parallel lines containing precipitates.
inside the twin indeed stand for the f0 0 0 2g twin basal planes. It 2. Unlike the RE-free extruded magnesium alloys, the present
is clear that the occurrence of twinning leads to the formation of a GW103K alloy exhibited basically symmetrical hysteresis loops,
large “kink” or interruption of the regular atomic arrangement on suggesting the absence of the tensile–compressive yield asym-
the basal plane along the fine horizontal lines, with the kink size metry. This was predominantly due to the presence of the
depending on the twin width. At the same time, the conjugate relatively weaker crystallographic texture and the lack of
f1 0 1 2g extension twinning plane K2 (not shown in the schematic twinning–detwinning activities stemming from the fine grain
diagram of Fig. 12) in an adjoining unit cell would rotate counter- size and RE-containing precipitates.
clockwise from η2 ½1 0 1 1( to η2n ½1 0 1 1( with a rotating angle of 3. Strain-controlled fatigue tests also showed that the present RE-
7.41 relative to the red-colored twinning plane considered, as containing GW103K alloy exhibited essentially cyclic stabiliza-
plotted in Fig. 12, where η2 (the dashed arrow) is the twinning tion until failure up to a strain amplitude of 1.0%. The cyclic
direction in the adjoining unit cell before twinning, and η2n (the strain hardening exponent of the alloy was observed to be
solid arrow) is the twinning direction in the same adjoining unit lower than that of RE-free extruded magnesium alloys obtained
cell after twinning, with a fixed angle of 86.31 between η1 and η2. at the same strain rate of 1 " 10−2 s−1 also due to its finer grain
Such twinning in the magnesium matrix grains has also been size and RE-containing precipitates.
reported, e.g., in [60,65,67,70]. If the particle/precipitate is not 4. The fatigue life of the present alloy was observed to be longer
sheared during twinning, as the twinning ledge approaches the than that of the RE-free extruded magnesium alloys, which can
particle there exists a back stress stemming from the harder and be well described by the Coffin–Manson law and Basquin's
more rigid particle, which would impede the free shear of the equation.
moving twin ledge and must be overcome by the applied stress 5. SEM examinations revealed that fatigue cracks initiated from
[68]. The portion of the matrix in the absence of twinning due to the near-surface defects and the initiation site contained some
the blockage of the precipitate is indicated by the dashed lines in cleavage-like facets. The propagation was basically character-
Fig. 12. Furthermore, when the twinning shear reaches the inter- ized by fatigue striations perpendicular to the crack propaga-
face between the matrix and the precipitate, the movement of tion direction observed at higher magnifications in conjunction
atoms towards the precipitate should be accommodated by the with secondary cracks.
surrounding material, which is likely accomplished by the basal
slip of the matrix near the particle/matrix interface. However, the
basal slip in the matrix along the fine horizontal lines would also
be inhibited by the precipitate, thereby providing an opportunity Acknowledgments
of shear being accommodated by the f0 0 0 2g twin basal slip (i.e.,
along the fine parallel lines inside the twin), as indicated in Fig. 12. The authors would like to thank the Natural Sciences and
Then the need for the additional basal slip on a local scale around Engineering Research Council of Canada (NSERC) and AUTO21
the precipitate is likely to increase the stress required to form the Network of Centres of Excellence for providing financial support.
twin [68]. As a result, the second-phase particles in the GW103K This investigation involves part of Canada–China–USA Collabora-
alloy (Fig. 1(b)) would significantly modify the potential twin tive Research Project on the Magnesium Front End Research and
nucleation sites and retard the rate of twin formation. Similar Development (MFERD). The authors also thank Dr. A.A. Luo,
results were reported by Jain et al. [60]. It should also be noted General Motors Research and Development Center for the supply of
that other precipitates, e.g., the common eutectic β-Mg17Al12 extruded AM30 magnesium alloy. One of the authors (X.Q. Zeng)
particles present in the cast magnesium alloys and semi-solid would like to give thanks to Ministry of Science and Technology of
processed magnesium alloys would also generate a strong resis- China (MOST) and Natural Science Foundation of China (NSFC) for
tance to the twinning (and basal slip as well) [71–76]. More their financial support (Project nos. 2011BAE22B02, 2011DFA50907,
detailed studies on the twin-particle interactions in these cast and 51171113). One of the authors (D.L. Chen) is also grateful for the
magnesium alloys are needed. financial support by the Premier's Research Excellence Award (PREA),
The above results and analyses indicated that both the grain NSERC-Discovery Accelerator Supplement (DAS) Award, Canada
refinement and the presence of precipitates in the present GW103K Foundation for Innovation (CFI), and Ryerson Research Chair (RRC)
alloy due to the addition of RE elements played a crucial role in program. The authors would like to thank Dr. R. Tandon and
suppressing the occurrences of deformation twins which led to the Dr. B. Davies (Magnesium Elektron) for supplying magnesium pow-
reduction or elimination of the tension–compression asymmetry, as ders for the defocusing calibration. The authors would also like to
shown in Fig. 3. In brief, the addition of RE elements was able to thank Messrs. A. Machin, Q. Li, J. Amankrah and R. Churaman for easy
influence substantially the overall cyclic deformation behavior of the access to the laboratory facilities of Ryerson University and their
extruded magnesium alloys via the refinement of grains, the pre- assistance in the experiments.
sence of RE-containing precipitates, the relatively weaker crystal-
lographic texture, and the suppression of twinning activities.
References

[1] N.P. Gillett, V.K. Arora, K. Zickfeld, S.J. Marshall, W.J. Merryfield, Nature Geosci.
4 (2011) 83–87.
5. Conclusions [2] J. Murray, D. King, Nature 481 (2012) 433–435.
[3] T.M. Pollock, Science 328 (2010) 986–987.
[4] M. Wise, K. Calvin, A. Thomson, L. Clarke, B.B. Lamberty, R. Sands, S.J. Smith,
Strain-controlled low cycle fatigue tests were conducted on an A. Janetos, J. Edmonds, Science 324 (2009) 1183–1186.
extruded Mg–10Gd–3Y–0.5Zr (GW103K) alloy with different total [5] W.J. Joost, JOM 64 (2012) 1032–1038.
F.A. Mirza et al. / Materials Science & Engineering A 575 (2013) 65–73 73

[6] A.Α. Luo, JOM 54 (2002) 42–48. [44] J. Bohlen, M.R. Nurnberg, J.W. Senn, D. Letzig, S.R. Agnew, Acta Mater. 55
[7] V.V. Ogarevic, R.I. Stephens, Annu. Rev. Mater. Sci. 20 (1990) 141–177. (2007) 2101–2112.
[8] X.Z. Lin, D.L. Chen, Mater. Sci. Eng. A 496 (2008) 106–113. [45] F.J. Humphreys, Acta Metall. 27 (1979) 1801–1814.
[9] F.A. Mirza, D.L. Chen, in: S. Zhang, D.L. Zhao (Eds.), Aerospace Materials [46] F.J. Humphreys, Acta Mater. 45 (1997) 5031–5039.
Handbook, CRC Press, Taylor & Francis, New York, 2013, pp. 647–698. [47] K. Hantzsche, J. Bohlen, J. Wendt, K.U. Kainer, S.B. Yi, D. Letzig, Scr. Mater. 63
[10] L. Chen, C. Wang, W. Wu, Z. Liu, G.M. Stoica, L. Wu, P.K. Liaw, Metall. Mater. (2010) 725–730.
Trans. A 38 (2007) 2235–2241. [48] W.G. Yang, C.H. Koo, Mater. Trans. 44 (2003) 1029–1035.
[11] S. Begum, D.L. Chen, S. Xu, A.A. Luo, Metall. Mater. Trans. A 39 (2008) [49] S. Suresh, Fatigue of Materials, second ed., Cambridge University Press,
3014–3026. Cambridge, 1998.
[12] S. Begum, D.L. Chen, S. Xu, A.A. Luo, Int. J. Fatigue 31 (2009) 726–735. [50] U. Noster, B. Scholtes, Zeitschrift fur Metallkunde 94 (2003) 559–563.
[13] S. Begum, D.L. Chen, S. Xu, A.A. Luo, Mater. Sci. Eng. A 517 (2009) 334–343. [51] S. Hasegawa, Y. Tsuchida, H. Yano, M. Matsui, Int. J. Fatigue 29 (2007)
[14] F. Yang, F. Lv, X.M. Yang, S.X. Li, Z.F. Zhang, Q.D. Wang, Mater. Sci. Eng. A 528 1839–1845.
(2011) 2231–2238. [52] L. Wu, A. Jain, D.W. Brown, G.M. Stoica, S.R. Agnew, B. Clausen, D.E. Fielden,
[15] E.A. Ball, P.B. Prangnell, Scr. Metall. Mater. 31 (1994) 111–116. P.K. Liaw, Acta Mater. 56 (2008) 688–695.
[16] Y.N. Wang, J.C. Huang, Mater. Chem. Phys. 81 (2003) 11–26. [53] Q. Yu, J. Zhang, Y. Jiang, Q. Li, Int. J. Fatigue 36 (2012) 47–58.
[17] S. Kleiner, P.J. Uggowitzer, Mater. Sci. Eng. A 379 (2004) 258–263. [54] J.B. Jordon, J.B. Gibson, M.F. Horstemeyer, H. El Kadiri, J.C. Baird, A.A. Luo,
[18] S.M. Yin, H.J. Yang, S.X. Li, S.D. Wu, F. Yang, Scr. Mater. 58 (2008) 751–754. Mater. Sci. Eng. A 528 (2011) 6860–6871.
[19] S.H. Safi-Naqvi, W.B. Hutchinson, M.R. Barnett, Mater. Sci. Technol. 24 (2008) [55] T.J. Luo, Y.S. Yang, W.H. Tong, Q.Q. Duan, X.G. Dong, Mater. Des 31 (2010)
1283–1292. 1617–1621.
[20] N. Stanford, M.R. Barnett, Mater. Sci. Eng. A 496 (2008) 399–408. [56] G.E. Dieter, Mechanical Metallurgy, SI metric ed., McGraw-Hill Inc., New York,
[21] N. Stanford, G. Sha, J.H. Xia, S.P. Ringer, M.R. Barnett, Scr. Mater. 65 (2011) 1986.
919–921. [57] Z. Liu, Y.Y. Xu, Z.G. Wang, Y. Wang, Z.Y. Liu, Acta Metall. Sin. (English letters) 13
[22] D. Wu, R.S. Chen, W.N. Tang, E.H. Han, Mater. Des. 41 (2012) 306–313. (2000) 961–966.
[23] N. Stanford, D. Atwell, A. Beer, C.H. Davies, M.R. Barnett, Scr. Mater. 59 (2008) [58] S.H. Chowdhury, D.L. Chen, S.D. Bhole, E. Powidajko, D.C. Weckman, Y. Zhou,
772–775. Metall. Mater. Trans. A 43 (2012) 2133–2147.
[24] N. Stanford, M.R. Barnett, Scr. Mater. 58 (2008) 179–182. [59] H.A. Patel, D.L. Chen, S.D. Bhole, K. Sadayappan, Mater. Sci. Eng. A 528 (2010)
[25] L. Wu, Z. Yang, W. Xia, Z. Chen, L. Yang, Mater. Des. 36 (2012) 47–53. 208–219.
[26] H. Bayani, E. Saebnoori, J. Rare Earths 27 (2009) 255–258. [60] J. Jain, W.J. Poole, C.W. Sinclair, M.A. Gharghouri, Scr. Mater. 62 (2010)
[27] Y. Yang, Y.B. Liu, S.Y. Qin, Y. Fang, J. Rare Earths 24 (2006) 591–595. 301–304.
[28] F.H. Wang, J. Dong, Y.Y. Jiang, W.J. Ding, Mater. Sci. Eng. A 561 (2013) 403–410. [61] S.M. Yin, C.H. Wang, Y.D. Diao, S.D. Wu, S.X. Li, J. Mater. Sci. Technol. 27 (2011)
[29] C.Y. Ma, X. Che, B. Zhao, L.J. Chen, J. Northeast. Univ. 28 (2007) 144–149. 29–34.
[30] C.L. Fan, D.L. Chen, A.A. Luo, Mater. Sci. Eng. A 519 (2009) 38–45. [62] J.P. Nobre, U. Noster, M. Kornmeier, A.M. Dias, B. Scholtes, Key Eng. Mater.
[31] Z.-K. Peng, X.-M. Zhang, J.-M. Chen, Y. Xiao, H. Jiang, Mater. Sci. Technol. 21 230–232 (2002) 267–270.
(2005) 722–726. [63] M.R. Barnett, Metall. Mater. Trans. A 34 (2003) 1799–1806.
[32] W.P. Li, H. Zhou, Z.F. Li, J. Alloys Compd. 475 (2009) 227–232. [64] J. Koike, T. Kobayashi, T. Mukai, H. Watanabe, M. Suzuki, K. Maruyama,
[33] X.Y. Fang, D.Q. Yi, B. Wang, W.H. Luo, W. Gu, Trans. Nonferrous Met. Soc. China K. Higashi, Acta Mater. 51 (2003) 2055–2065.
16 (2006) 1053–1058. [65] J. Bohlen, P. Dobron, J. Swiostek, D. Letzig, F. Chmelik, P. Lukac, K.U. Kainer,
[34] K. Yu, S.T. Rui, J.M. Song, W.X. Li, L. Guo, Trans. Nonferrous Met. Soc. China 18 Mater. Sci. Eng. A 462 (2007) 302–306.
(2008) s39–s43. [66] M.R. Barnett, Z. Keshavarz, A.G Beer, D. Atwell, Acta Mater. 52 (2004)
[35] M.Z. Li, Y.Q. Wang, C. Li, X.G. Liu, B.S. Xu, Mater. Sci. Technol. 27 (2011) 5093–5103.
1138–1142. [67] A. Jain, O. Duygulu, D.W. Brown, C.N. Tome, S.R. Agnew, Mater. Sci. Eng. A 486
[36] J. Zhang, X.F. Zhang, F.S. Pan, C.P. Liu, Q.B. He, Mater. Sci. Forum 610–613 (2008) 545–555.
(2009) 810–814. [68] N. Stanford, M.R. Barnett, Mater. Sci. Eng. A 516 (2009) 226–234.
[37] S.M. He, X.Q. Zeng, L.M. Peng, X. Gao, J.F. Nie, W.J. Ding, J. Alloys Comp. 427 [69] Y.N. Wang, J.C. Huang, Acta Mater. 55 (2007) 897–905.
(2007) 316–323. [70] P.G. Partridge, Int. Mater. Rev. 12 (1967) 169–194.
[38] F. Zhang, K.X. Zhang, C.W. Tan, X.D. Yu, H.L. Ma, F.C. Wang, H.N. Cai, Trans. [71] H.A. Patel, N. Rashidi, D.L. Chen, S.D. Bhole, A.A. Luo, Mat. Sci. Eng. A 546
Nonferrous Met. Soc. China 21 (2011) 2140–2146. (2012) 72–81.
[39] S.Q. Liang, D.K. Guan, L. Chen, Z.H. Gao, H.X. Tang, X.T. Tong, Mater. Des. 32 [72] H.A. Patel, D.L. Chen, S.D. Bhole, K. Sadayappan, Mater. Des. 49 (2013)
(2011) 361–364. 456–464.
[40] Q. Wang, G. Wu, H.Q. Zheng, B. Chen, Y. Zheng, W. Ding, China Found. J. 7 [73] J.B. Clarke, Acta Metall. 16 (1968) 141–152.
(2010) 6–12. [74] M.A. Gharghouri, G.C. Weatherly, J.D. Embury, Philos. Mag. A 78 (1998)
[41] D. Sarker, D.L. Chen, Scr. Mater. 67 (2012) 165–168. 1137–1149.
[42] F. Lv, F. Yang, Q.Q. Duan, Y.S. Yang, S.D. Wu, S.X. Li, Z.F. Zhang, Int. J. Fatigue 33 [75] L.M. Wang, C.C. Chen, Mater. Lett. 67 (2012) 158–161.
(2011) 672–682. [76] N. Stanford, J. Geng, Y.B. Chun, C.H.J. Davies, J.F. Nie, M.R. Barnett, Acta Mater.
[43] R. Zhu, Y.J. Wu, J.T. Wang, K.C. Lu, Adv. Mater. Res. 320 (2011) 222–227. 60 (2012) 218–228.
Materials and Design 46 (2013) 411–418

Contents lists available at SciVerse ScienceDirect

Materials and Design


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

Effect of rare earth elements on deformation behavior of an extruded


Mg–10Gd–3Y–0.5Zr alloy during compression
F.A. Mirza a, D.L. Chen a,⇑, D.J. Li b, X.Q. Zeng b,⇑
a
Department of Mechanical and Industrial Engineering, Ryerson University, 350 Victoria Street, Toronto, Ontario M5B 2K3, Canada
b
The State Key Laboratory of Metal Matrix Composites, School of Materials Science and Engineering, Shanghai Jiao Tong University, 800 Dongchuan Road, Shanghai 200240, PR China

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

Article history: The aim of this study was to identify the influence of rare-earth (RE) elements on the strain hardening
Received 30 August 2012 behavior in an extruded Mg–10Gd–3Y–0.5Zr magnesium alloy via compression in the extrusion direction
Accepted 24 October 2012 at room temperature. The plastic deformation behavior of this RE-containing alloy was characterized by a
Available online 6 November 2012
rapidly decreasing strain hardening rate up to a strain level of about 4% (stage A), followed by a fairly flat
linear strain hardening rate over an extended strain range from 4% to 18% (stage B). Stage C was rep-
Keywords: resented by a decreasing strain hardening rate just before failure. The extent of twinning in this alloy was
Magnesium alloy
observed to be considerably less extensive than that in the RE-free extruded Mg alloys. The weaker crys-
Rare-earth element
Plastic deformation
tallographic texture, refined grain size, and second-phase particles arising from the addition of RE ele-
Strain hardening ments were responsible for the much higher strain hardening rate in stage A due to the increased
Twinning difficulty on the formation of twins and the slip of dislocations at lower strains, and for the occurrence
Slip of quite flat linear strain hardening in stage B at higher strains which was likely related to the dislocation
debris and twin debris (or residual twins) stemming from dislocation–twin interactions as well as the
interactions between dislocations/twins and second-phase particles and grain boundaries.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction their normal direction; in the extrusion the c-axes tend to align
transversely to the extrusion axis. These forming processes would
The extreme precipitation events and environmental problems result in the presence of a strong preferred orientation or texture in
under global warming, which are today recognized to stem from Mg alloys. Other mechanisms such as discontinuous recrystalliza-
anthropogenic environment-damaging emissions, are increasing tion due to grain boundary sliding (GBS) would also produce some
severely in recent years [1]. To lower the greenhouse gas emissions reorientations [7].
and promote the fuel efficiency in a time of environmental chal- It has recently been observed that magnesium alloys containing
lenge and impending energy crisis [2], lightweight vehicle design rare earth (RE) elements can develop more random textures during
has become an important goal in the automotive and aerospace hot extrusion [8,9]. The addition of RE elements in magnesium al-
industries [3]. Recently, the development and application of the ul- loys leads to a decrease of the overall texture sharpness allowing
tra-lightweight magnesium alloys have been significantly increas- the activation of easy basal slip to a higher extent, compared to
ing in the transportation sectors due to their low density, high other RE-free wrought magnesium alloys [8,10,11]. Furthermore,
strength-to-weight ratio, and superior damping capacity [4–6]. RE elements could result in the formation of a f1 2 1 1g texture
In contrast to the cubic metals, the deformation of magnesium component parallel to the extrusion direction [8,10,11]. In the
is fairly complex because its hexagonal close-packed (hcp) crystal presence of strong texture, wrought Mg alloys were prone to
structure has a limited number of active slip systems at room tem- over-concentrated twinning–detwinning activities which may
perature, with basal slip and twinning playing a key role in the aggravate the cyclic deformation irreversibility and increase the
plasticity of magnesium alloys [5,6]. Both basal slip and twinning propensity of crack initiation [5,6,12]. During cyclic deformation,
mechanisms tend to align the c-axes of grains with the direction the level of strain amplitudes applied at a strain ratio of R = ÿ1
of predominantly compressive strains during forming (rolling, was normally limited to a maximum value of about 1–2% in both
extrusion). In the magnesium sheet and plate, the c-axes align with tensile and compressive phases [5]. This would potentially limit
the occurrence of twinning to its full capacity in the compressive
⇑ Corresponding authors. Tel.: +1 416 979 5000x6487; fax: +1 416 979 5265 (D.L. phase during cyclic deformation. Although the addition of RE ele-
Chen), tel.: +86 21 5474 2301; fax: +86 21 3420 3730 (X.Q. Zeng). ments offers the possibility of effectively improved mechanical
E-mail addresses: dchen@ryerson.ca (D.L. Chen), xqzeng@sjtu.edu.cn (X.Q. Zeng). properties via texture weakening, it remains unclear what role

0261-3069/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.matdes.2012.10.040
412 F.A. Mirza et al. / Materials and Design 46 (2013) 411–418

the RE elements play in the formation of twins and how the twin- 3. Results and discussion
ning develops beyond the above strain level, whether the detwin-
ning would occur with increasing compressive strain, and how the 3.1. Microstructure and texture
twinning or detwinning affects the strain hardening behavior. This
study was aimed to identify the deformation behavior of an ex- Fig. 1a shows a typical optical image on the microstructure of
truded Mg–10Gd–3Y–0.5Zr alloy in compression along the extru- Mg–10Gd–3Y–0.5Zr alloy in the as-extruded condition where the
sion direction at different strain levels, with special attention to ED is indicated. For the sake of comparison, the microstructure of
the effect of the RE element addition on the twin formation. an extruded AM30 magnesium alloy is shown in Fig. 1b as well.
It is seen that uniform equiaxed grains with an average grain size
of about 12 lm were obtained in the Mg–10Gd–3Y–0.5Zr alloy
2. Experimental procedure
(Fig. 1a) due to the occurrence of dynamic recrystallization (DRX)
in the hot extrusion process at 400°C, which is fairly small in com-
The material selected was a recently developed extruded mag-
parison with AM30 alloy (Fig. 1b). This was due to the role of added
nesium alloy with the following composition (wt.%): 10Gd, 3Y,
RE elements and zirconium (Zr) where Zr mainly restricted the
0.5Zr, and Mg (balance). The received alloy bars were extruded at
grain growth [13]. A similar role of grain refinement by RE ele-
400 °C with a diameter of 20 mm and an extrusion ratio of 9:1.
ments, e.g., gadolinium [14], yttrium [15], cerium [16], neodymium
Microstructural examinations were performed using an optical
[17], erbium [18], has been reported as well. For example, an addi-
microscope (OM), scanning electron microscope (SEM, JEOL, JSM-
tion of Gd into AZ31 alloy resulted in the formation of Al2Gd
6380LV) equipped with Oxford energy dispersive X-ray spectros-
(which was still present after homogenization), and it played an
copy (EDS) system and three-dimensional (3D) surface/fracto-
important role in refining the grain size and improving the rolling
graphic analysis capacity, and transmission electron microscope
capability of the alloy [14]; the cerium exhibited a good grain
(TEM, JEM-2100) operated at 200 kV. Standard metallographic
refinement effect on the as-cast AZ31 alloy [16]; and remarkable
sample preparation techniques were used with an etchant based
grain size refinement of Mg–1.5Zn–0.1Zr alloy with an addition
on an acetic picral solution containing 4.2-g picric acid, 10-ml ace-
of erbium was observed through recrystallization during the roll-
tic acid, 10-ml H2O, and 70-ml ethanol. The average grain size was
ing process [18]. Besides, as seen from Fig. 1 no twins were visible
measured via a linear intercept method. Texture measurements
in the un-deformed samples. Fig. 2a shows a typical SEM image of
were performed using a PANalytical X-ray diffractometer with
as-extruded sample where it contained three different phases as
Cu Ka radiation at 45 kV and 40 mA. The texture were determined
identified by He et al. [19]: a-Mg solid solution matrix phase with
by measuring incomplete pole figures between w = 0 to 75° in a
supersaturated Gd + Y elements, (Gd + Y)-rich eutectic compound
back reflection mode, and analyzed using PANalytical X’Pert Tex-
which had a higher Gd + Y content than the matrix, and intracrys-
ture software. Cylindrical specimens with a diameter of 5 mm
talline Zr-rich cores. By means of EDS point analysis, the content at
and height of 8 mm were machined with the compression axis par-
points A, B, and C of Fig. 2a was estimated to be 81.7% Mg–4.9% Y–
allel to the extrusion direction (ED) of the as-received bars. Com-
13.4% Gd, 62.5% Mg–6.1% Y–31.1% Gd, and 32.4% Mg–67.6% Zr,
pression tests were conducted using a computerized Instron
respectively. Similar observations have been reported for as-cast
machine at a strain rate of 10ÿ3 sÿ1 and stopped at different
samples of Mg–7Gd–3Y–0.4Zr (GW73K) alloy by Liang et al. [20],
amounts of strain at room temperature. The deformed samples
cast Mg–10Gd–3Y–Zr alloy by Wang et al. [21], as-extruded Mg–
were cut along the compression axis diameter using a slow dia-
11.90Gd–0.81Y–0.44Zr alloy by Zhang et al. [22]. In addition, EDS
mond cutter, cold-mounted, ground, polished, and etched to exam-
line scan was performed as shown in Fig. 2b and c which also con-
ine the evolution of deformation twins during compression. A
firmed the presence of RE-rich particles in the as-extruded alloy,
point counting procedure according to ASTM: E562-11 was used
where the middle rectangle-like particle contained both Gd and
to evaluate the volume fraction of twins. Five digital micrographs
Y. Such RE-enrichment particles was also observed via TEM which
were taken from each sample at a magnification of 200, and then
was operated in a microdiffraction mode, as shown in Fig. 3a. Elec-
a grid of 31 points by 39 points was superimposed on the images.
tron microdiffration pattern in Fig. 3b shows the structure of the
The volume fraction was determined as a ratio of the number of
particle which was identified as cuboid-shaped (RE-rich) particles
points positioned within the twins to the total number of grid
(face-centered cubic, fcc with a = 0.56 nm). Similar results were re-
points. It should be noted that in evaluating the stress–strain
ported in a Mg–10Gd–2Y–0.5Zr alloy [19].
curves and strain hardening rates, the deformation amount of ma-
Fig. 4 shows a basal texture where the reflecting surface was
chine system itself was eliminated using a calibration curve so as
parallel to the extrusion direction (ED) and the radial direction
to obtain the actual or net deformation of test samples.

(a) (b)

ED ED

Fig. 1. Microstructures of (a) as-extruded Mg–10Gd–3Y–0.5Zr, and (b) as-extruded AM30 alloy.
F.A. Mirza et al. / Materials and Design 46 (2013) 411–418 413

normal direction and some grains with about half of the maximum
(a) C basal intensity were oriented toward the RD. The presence of such
a weaker texture was another benefit of the RE elements added
into Mg alloys, as also reported by Stanford and Barnett [11] who
observed that microalloying with RE elements can weaken the
A extrusion texture. A similar weaker texture in Mg-based alloys
with the addition of RE elements was reported [23]. In the current
Mg–10Gd–3Y–0.5Zr alloy, grain boundary nucleation, coalescence
of low-angle boundaries and shear band assisted nucleation are ab-
sent (as seen in Fig. 1). Zhu et al. [23] reported that only particle-
stimulated nucleation was considered to be a factor influencing
B texture weakening. Indeed, Humphreys [24] observed that the high
RE content and mixed addition resulted in the precipitation of a
large number of the second-phase particles. Such precipitates can
hinder the movement of the dislocations so that the stress could
concentrate easily near the precipitates, which promoted the par-
ticle-stimulated nucleation of dynamic recrystalliazation, and
weaken the basal texture. In addition, the weakening of the texture
was also associated with the appearance of deformation bands
containing twins and the restriction of grain growth [25].
(b)
3.2. Uniaxial compression behavior

The stress–strain curves of the extruded Mg–10Gd–3Y–0.5Zr al-


loy during compression along the ED are shown in Fig. 5a, where
the stress–strain curves of several samples deformed at varying
amounts of strain are plotted together to show the reproducibility
of the compression tests. An average compressive yield strength of
232 ± 4 MPa was obtained from the present multiple tests, which
was higher than the tensile yield strength of 219 MPa for a Mg–
10Gd–2Y–0.5Zr alloy [19]. While the tensile percent elongation
of Mg–10Gd–2Y–0.5Zr alloy was reported to be 25.6% for the rod
also extruded at 400°C [19], which was higher than the compres-
sive percent shortening (18.7%) of the present Mg–10Gd–3Y–
0.5Zr alloy, the ultimate compressive strength of the present alloy
(493 MPa) was significantly higher than the ultimate tensile
strength (305 MPa) of Mg–10Gd–2Y–0.5Zr alloy [19]. This might
2000 be attributed to the favorable role of one more percent yttrium
added, while all other elements remained the same. Furthermore,
(c) in comparison with the compressive test results of an extruded
AM30 alloy [26], the present RE-containing alloy had both higher
(b) ultimate compressive strength and ductility (as indicated by the
1500 Mg compressive percent shortening). In particular, the current Mg–
Intensity, Counts

10Gd–3Y–0.5Zr alloy exhibited over threefold higher compressive


yield strength than the extruded AM30 alloy, i.e., 232 MPa vs.
71 MPa [26]. Again this reflected the significant beneficial effect
1000
of RE elements added into Mg alloys. Furthermore, unlike the sig-
moidal (S-shaped) hardening curves observed in the twinning-
dominated deformation, such as in the extruded AM30 Mg alloy
[26], the compressive stress–strain curves in Fig. 5a showed
500 roughly a parabolic-like hardening, indicating a likely change in
the deformation mechanisms from the twinning-dominated defor-
Gd Y mation to the slip-dominated deformation. To better understand
the deformation behavior of the present alloy, the strain hardening
0 rate dr/de (where r is true stress) with respect to true strain e from
0 5 10 15 20 25 the specimen tested until failure is presented in Fig. 5b along with
Distance, µm the RE-free AM30 extruded Mg alloy [26]. Three stages of strain
hardening in the extruded Mg–10Gd–3Y–0.5Zr alloy could be dis-
Fig. 2. (a) SEM backscattered electron image indicating EDS point analysis locations tinguished from Fig. 5b. Stage A was characterized by a rapidly
(A, B, and C), (b) SEM backscattered electron image indicating EDS line scan decreasing strain hardening rate from the onset of plastic deforma-
position, and (c) the corresponding EDS line scan results in the as-extruded Mg–
tion up to a strain of 4%, followed by stage B with a fairly flat and
10Gd–3Y–0.5Zr alloy.
almost linear strain hardening rate over a large range of strain up
to 18%, and then stage C with an decreasing strain hardening rate
(RD). A basal plane fiber texture was observed, where the c-axis of again until failure. Three stages of strain hardening rate have been
most grains was aligned perpendicular to the ED. The peak inten- reported by Sarker and Chen [26] where stage A and stage C had a
sity of the basal plane split to a few degrees around the surface falling trend, as observed in the present study as well. However, as
414 F.A. Mirza et al. / Materials and Design 46 (2013) 411–418

(a) (b)

Fig. 3. Transmission electron micrographs of the as-extruded Mg–10Gd–3Y–0.5Zr alloy: (a) bright-field image taken along [0 0 1] zone axis of particle A, and (b) [0 0 1] zone
axis microdiffraction pattern recorded from particle A.

Intensity Color
1 0.231
26 1.194
51 2.158
76 3.122
100 4.048

ED

RD

Fig. 4. (0 0 0 2) Pole figure of as-extruded Mg–10Gd–3Y–0.5Zr alloy, where ED stands for the extrusion direction and RD indicates the radial direction.

seen from Fig. 5b, the value of strain hardening rate in stage A was of strain were examined. Fig. 6 shows the change of twin volume
much higher in the present extruded Mg–10Gd–3Y–0.5Zr alloy fraction with the applied compressive strain, together with some
than in the extruded AM30 alloy; stage B of the AM30 Mg alloy typical images showing the visual change of twins. It is seen from
was totally different from that of the present alloy. Similar trend image (a) in Fig. 6 that at a compressive strain of e = 0.9%, no twins
in the strain hardening rate to the Mg–10Gd–3Y–0.5Zr alloy has were visible, whereas at e = 3.2% a very small number of twins
also been observed in extruded Mg–Y–Nd alloy [27] and ZK60 were observed in some grains (image (b) in Fig. 6). The twins
Mg alloy [28], where the strain hardening rate started with a sim- formed during compression in the extruded Mg alloys along the
ilarly steep decrease at small strains and then approached a linear/ ED direction has been identified as f1 0 1  2g extension twins in a
steady state with a very low value over a large range of strain. The number of publications [23,26]. At a strain of e = 7.7%, more twins
significant difference in the strain-hardening rate in stages A and B were observed in more grains (image (c) in Fig. 6). When the spec-
between the present Mg–10Gd–3Y–0.5Zr and AM30 alloys could imen was deformed at a compressive strain of e = 12.3%, the twin
be understood on the basis of the amount of twins formed during volume fraction reached its maximum of 0.04 (image (d) in
compression and the subsequent interactions between twin and Fig. 6). Further compression beyond this strain value resulted in
twin, twin and dislocation, twin–precipitate, etc., which will be a decrease in the twin volume fraction or the disappearance of
discussed in the next section. twins, as seen in images (e) and (f) in Fig. 6. In addition, Fig. 7
shows the typical SEM micrographs at strain levels of 3.2%, and
3.3. Twin volume fraction 12.3%, respectively. It is obvious that the compressive deformation
at a strain of 12.3% resulted in more twins than that at a strain of
To understand the strain hardening behavior, the deformation 3.2%. The phenomenon of detwinning occurred during pure com-
characteristics of the samples compressed to different amounts pression in an extruded Mg alloy was recently reported in [26]. It
F.A. Mirza et al. / Materials and Design 46 (2013) 411–418 415

600 hardening [30]. Most grains re-orient to a stable orientation so that


(a) the effect of flow stress would be expected to saturate and not lead
500 to any new strain hardening at very large strains. As the addition of
Engineering stress, MPa

RE elements offers an avenue to control or change the crystallo-


graphic texture (as mentioned in Section 3.1), it would thus affect
400
the slip and twinning systems and change the deformation behav-
ior. It was observed that, in the case of a weak basal texture (Fig. 4),
300 it is expected that hai dislocation slip with a relatively low CRSS
would contribute more to accommodate the deformation and twin
up to failure
200 formation will be more difficult [31]. Moreover, Hantzsche et al.
12.3% [25] reported that the deformation texture changes at a higher Y
7.7% content, which is related to an increasing activity of hc + ai slip
100
3.2% and thus can also reflect in the twinning activity.
0.9% There is a general agreement that a smaller grain size inhibited
0
twinning [32] and grain size had a significant effect on the ten-
0 4 8 12 16 20
dency of twinning since the energy required to form twin inter-
Engineering strain, %
faces is particularly high in Mg [33]. Finer grains, which have a
7000 higher surface-to-volume ratio, increase the difficulty of twin
(b) This investigation
nucleation and thereby suppress twinning by limiting the sizes of
6000 the twins [34]. The addition of RE elements and Zr in the present
Strain hardening rate, MPa

AM30 [26] Mg–10Gd–3Y–0.5Zr alloy led to a fairly fine grain size, as seen from
5000 Fig. 1. Thus in the present alloy, the grain size would be another
important factor influencing the incidence of deformation twin-
4000 A ning. It was also reported that it is common for the twinning stress
to increase with decreasing grain size more rapidly than the stress
3000 required to activate slip [35]. As a result, the formation of deforma-
tion twins was more difficult in the Mg–10Gd–3Y–0.5Zr alloy as
2000 B compared with AM30 alloy [26] due to the grain refinement role
C of RE elements as mentioned in the above Section 3.1. In addition
1000
to the effect of texture and grain size, the presence of precipitates
0
has also been reported to impede the process of twinning in mag-
0 4 8 12 16 20 nesium alloys [36]. During plastic deformation as the twinning
True strain, % ledge approaches a particle there exists a back-stress arising from
the rigid particle inhibiting the free shear of the moving twin ledge.
Fig. 5. (a) Compressive stress–strain curves and (b) strain hardening rate as a This back-stress must be overcome by the applied stress. When the
function of true strain of as-extruded Mg–10Gd–3Y–0.5Zr alloy in the extrusion shear takes place at the interface between the matrix and the pre-
direction tested at a strain rate of 10ÿ3 sÿ1.
cipitate, the movement of atoms toward the precipitate must be
accommodated by basal slip away from the particle/matrix inter-
was explained through a twin–dislocation interaction model, face [36]. Then the need for additional basal slip on a local scale
where the moving dislocations during compression cut and de- around the precipitates is likely to increase the stress required to
stroyed twins, leading to the disappearance of twins [26]. form the twin. As a result, the second-phase particles in the Mg–
A comparison of the twin volume fraction as a function of the 10Gd–3Y–0.5Zr alloy (Figs. 2 and 3) would modify potential twin
applied compressive strain between the extruded Mg–10Gd–3Y– nucleation sites and hinder the rate of twin formation. Similar re-
0.5Zr alloy and extruded AM30 alloy is shown in Fig. 8. It is clear sults were reported in Ref. [28].
that twinning occurred in the present Mg–10Gd–3Y–0.5Zr alloy It was reported that at large compressive strains the spacing be-
was much less extensive than that in the extruded AM30 alloy tween the second-phase decreased [29]. If the second-phase is
[26]. The maximum volume fraction in the AM30 alloy (0.47), oc- hard enough to block the dislocation flow or twinning in the ma-
curred at a compressive strain of 2.5%, was considerably (or over trix, the mean free path for dislocation motion and twinning de-
one order of magnitude) higher than that in the Mg–10Gd–3Y– creases strongly with strain and leads to rises of the flow stress
0.5Zr alloy (0.04) that appeared at a compressive strain of by the Hall–Petch effect, thus contributing to the strain hardening
12.3%. Other investigators also observed the occurrence of a sig- [29,32]. In addition, strain hardening is in nature associated with
nificant amount of twinning in the extruded RE-free Mg alloys elastic and inelastic incompatibilities between matrix and particles
[5,11]. The presence of RE elements in the current Mg–10Gd–3Y– during plastic deformation. The direction of this resistance is oppo-
0.5Zr alloy played a key role in suppressing the excessive twinning site to the applied stress, and therefore opposes the bulk slip and
and delaying the appearance of the maximum twinning. It is also twinning, contributing to the overall strain hardening as well. Fur-
due to the addition of RE elements that resulted in a huge differ- thermore, the plastic relaxation of the matrix at the RE particle
ence in the strain hardening characteristics between Mg–10Gd– interface could create both deformation twins and a high local dis-
3Y–0.5Zr alloy and AM30 alloys, especially in stages A and B location density, which are obstacles to the moving dislocations,
(Fig. 5b). Explanations for this are given below. and would also affect the strain hardening [37]. Similar result
Plastic deformation of Mg alloy is mainly determined by the regarding the effect of second-phase particles on the strain harden-
hexagonal crystal structure and strongly affected by the crystallo- ing has also been reported in Ref. [25]. Therefore, the presence of
graphic texture. Thus, crystallographic texture would be one of the RE-containing particles in conjunction with finer grain sizes and
key factors influencing the deformation behavior of Mg–10Gd–3Y– weaker texture is responsible for the much higher strain hardening
0.5Zr alloy. As the orientation of each grain changes, the combina- rate in stage A in the Mg–10Gd–3Y–0.5Zr alloy (Fig. 5b), due to the
tion of active slip system changes [29,30]. When a change of active increased difficulty on the slip of dislocations and the formation of
slip systems occurs, the flow stress will rise which will change the twins.
416 F.A. Mirza et al. / Materials and Design 46 (2013) 411–418

0.08

0.07

(c) (d)
0.06
Twin volume fraction

0.05 20 µm 20 µm

(e)
(b)
0.04

20 µm
0.03 20 µm

(f)

0.02
(a)
20 µm

0.01

20 µm

0
0 5 10 15 20 25
Strain, %

Fig. 6. Variation of the volume fraction of twins with the applied strain during compression in the as-extruded Mg–10Gd–3Y–0.5Zr alloy, where some typical micrographs at
different strain levels are inserted to show the evolution of deformation twins.

Fig. 7. Typical SEM micrographs of as-extruded Mg–10Gd–3Y–0.5Zr alloy deformed in compression at a strain level of (a) 3.2%, and (b) 12.3%.

The concept of dislocation hardening could also be used to of dislocation segments in this region. If the effect of the accu-
explain the linear portion in stage B of the strain hardening rate mulating is to raise the flow stress within the cell walls of the
curve of Mg–10Gd–3Y–0.5Zr alloy, as shown in Fig. 5b. Several dislocation substructure, then superposition of flow stress would
models have been proposed to explain stage B (in some studies contributed in hardening and ultimately leads to a finite
which is referred to as stage IV) [30,38–40]. It was reported that hardening rate in stage B (fairly flat linear or sometimes con-
stage B might be associated with microscopic instabilities such stant). Similar observations of stage IV or stage B in different ex-
as microbands commonly observed at large strains in copper truded Mg alloys were reported in several studies [27,28]. The
[38]. The cell model proposed by Nix et al. [40] indicated that above discussion indicated that the addition of RE elements
plastic straining at low temperatures leads to the accumulation was able to affect significantly the deformation behavior of the
of dislocation debris such as dipoles and loops. The accumulation extruded Mg alloys by predominantly suppressing the twinning
of dipolar debris may occur as part of the glide process and it due to the finer grain size, the relatively more random crystallo-
may also occur as part of the dislocation rearrangements due graphic texture, and the presence of RE-containing precipitates.
to dynamic recovery [39]. However, the rate of dislocation However, more studies related to the dislocation–twin interac-
rearrangement inside the cell walls is expected to be affected tions, twin–particle interactions, twin–grain boundary interac-
by the debris left and would have forward internal stresses with- tions from both experimental and modeling/theoretical aspects
in the cell walls that lead to rearrangement or even annihilation are needed.
F.A. Mirza et al. / Materials and Design 46 (2013) 411–418 417

0.6 [2] Murray J, King D. Oil’s tipping point has passed. Nature 2012;481:433–5.
[3] Pollock TM. Weight loss with magnesium alloys. Science 2010;328:986–7.
This investigation [4] Zhang Y, Zeng XQ, Lu C, Ding WJ. Deformation behavior and dynamic
recrystallization of a Mg–Zn–Y–Zr alloy. Mater Sci Eng A 2006;428(1–2):91–7.
AM30 [26] [5] Begum S, Chen DL, Xu S, Luo AA. Effect of strain ratio and strain rate on low
Twin volume fraction

cycle fatigue behavior of AZ31 wrought magnesium alloy. Mater Sci Eng A
0.4 2009;517:334–43.
[6] Lin XZ, Chen DL. Strain controlled cyclic deformation behavior of an extruded
magnesium alloy. Mater Sci Eng A 2008;496(1–2):106–13.
[7] Agnew SR, Duygulu O. Plastic anisotropy and the role of non-basal slip in
magnesium alloy AZ31B. Int J Plast 2005;21:1161–93.
[8] Stanford N, Sha G, Xia JH, Ringer SP, Barnett MR. Solute segregation and texture
modification in an extruded magnesium alloy containing gadolinium. Scripta
0.2
Mater 2011;65:919–21.
[9] Wu D, Chen RS, Tang WN, Han EH. Influence of texture and grain size on the
room-temperature ductility and tensile behavior in a Mg–Gd–Zn alloy
processed by rolling and forging. Mater Des 2012;41:306–13.
[10] Stanford N, Atwell D, Beer A, Davies CH, Barnett MR. Effect of microalloying
with rare-earth elements on the texture of extruded magnesium-based alloys.
0 Scripta Mater 2008;59:772–5.
0 5 10 15 20 [11] Stanford N, Barnett MR. Effect of composition on the texture and deformation
Strain, % behaviour of wrought Mg alloys. Scripta Mater 2008;58:179–82.
[12] Zhang XP, Castagne S, Luo XF, Gu CF. Effects of extrusion ratio on the ratcheting
Fig. 8. A comparison of twin volume fraction as a function of the applied strain behavior of extruded AZ31B magnesium alloy under asymmetrical uniaxial
during compression between the extruded Mg–10Gd–3Y–0.5Zr alloy and extruded cyclic loading. Mater Sci Eng A 2011;528(3):838–45.
AM30 alloy. [13] Peng Z-K, Zhang X-M, Chen J-M, Xiao Y, Jiang H. Grain refining mechanism in
Mg–9Gd–4Y alloys by zirconium. Mater Sci Technol 2005;21(6):722–6.
[14] Li WP, Zhou H, Li ZF. Effect of gadolinium on microstructure and
rolling capability of AZ31 alloy. J Alloys Compd 2009;475(1–2):
4. Conclusions 227–32.
[15] Fang XY, Yi DQ, Wang B, Luo WH, Gu W. Effect of yttrium on microstructures
The deformation behavior of as-extruded Mg–10Gd–3Y–0.5Zr and mechanical properties of hot rolled AZ61 wrought magnesium alloy. Trans
Nonferrous Met Soc China 2006;16(5):1053–8.
magnesium alloy in compression along the extrusion direction [16] Yu K, Rui ST, Song JM, Li WX, Guo L. Effects of grain refinement on mechanical
was characterized by a rapidly decreasing strain hardening rate properties and microstructures of AZ31 alloy. Trans Nonferrous Met Soc China
up to a strain level of 4% (stage A), followed by a fairly flat linear 2008;18:s39–43.
[17] Jia SG, Liu P, Dong QM, Xia CQ, Tian BH, Ren FZ. Effect of rare earth neodymium
strain hardening rate over an extended strain range from 4% to on microstructure and properties of AZ31B magnesium alloy. Trans Mater
18% (stage B). Stage C was represented by a decreasing strain Heat Treat 2010;31(7):99–103.
hardening rate just prior to failure. The amount of twins observed [18] Zhang J, Zhang XF, Pan FS, Liu CP, He QB. Microstructure and mechanical
properties of an Mg–Zn–Zr alloy modified by rare earth of Erbium. Mater Sci
in the present alloy was much less extensive than that in the RE-
Forum 2009;610–613:810–4.
free extruded AM30 Mg alloy, with a higher strain value of [19] He SM, Zeng XQ, Peng LM, Gao X, Nie JF, Ding WJ. Microstructure and
12.3% corresponding to the maximum twin volume fraction. strengthening mechanism of high strength Mg–10Gd–2Y–0.5Zr alloy. J Alloys
Compd 2007;427:316–23.
The weaker crystallographic texture, refined grain size, and RE-
[20] Liang SQ, Guan DK, Chen L, Gao ZH, Tang HX, Tong XT. Precipitation and its
containing second-phase particles were responsible for the much effect on age-hardening behavior of as-cast Mg–Gd–Y alloy. Mater Des
higher strain hardening rate in stage A in the Mg–10Gd–3Y–0.5Zr 2011;32:361–4.
alloy due to the increased difficulty on the formation of twins [21] Wang Q, Wu G, Zheng HQ, Chen B, Zheng Y, Ding W. A comparative study of
Mg–Gd–Y–Zr alloy cast by metal mould and sand. China Foundry J
and the slip of dislocations at lower strains, and for the occurrence 2010;7(1):6–12.
of fairly flat linear strain hardening rate in stage B at higher strains [22] Zhang F, Zhang KX, Tan CW, Yu XD, Ma HL, Wang FC, et al. Microstructure and
which was likely associated with the presence of dislocation debris mechanical properties of Mg–Gd–Y–Zr alloys processed by equal channel
angular pressing. Trans Nonferrous Met Soc China 2011;21:2140–6.
and twin debris (or residual twins) resulting mainly from disloca- [23] Zhu R, Wu YJ, Wang JT, Lu KC. Mechanical anisotropy of extruded Mg–10Gd–
tion–twin interactions as well as the interactions between the dis- 2Y–0. 5Zr alloy. Adv Mater Res 2011;320:222–7.
locations/twins and second-phase particles and grain boundaries. [24] Humphreys FJ. Unified theory of recovery, recrystallization and grain growth,
based on the stability and growth of cellular microstructures – II. The effect of
second-phase particles. Acta Mater 1997;45(12):5031–9.
Acknowledgements [25] Hantzsche K, Bohlen J, Wendt J, Kainer KU, Yia SB, Letting D. Effect of rare earth
additions on microstructure and texture development of magnesium alloy
sheets. Scripta Mater 2010;63:725–30.
The authors would like to thank the Natural Sciences and Engi- [26] Sarker D, Chen DL. Detwinning and strain hardening of an extruded
neering Research Council of Canada (NSERC), Premier’s Research magnesium alloy during compression. Scripta Mater 2012;67(2):165–8.
[27] Xin RL, Song B, Zeng K, Huang GJ, Liu Q. Effect of aging precipitation on
Excellence Award (PREA), NSERC-DAS Award, AUTO21 Network mechanical anisotropy of an extruded Mg–Y–Nd alloy. Mater Des
of Centers of Excellence, and Ryerson Research Chair program for 2012;34:384–8.
providing financial support. This work was also co-funded by Tech- [28] Chen XH, Pan FS, Mao JJ, Wang JF, Zhang DF, Tang AT, et al. Effect of heat
treatment on strain hardening of ZK60 Mg alloy. Mater Des 2011;32:1526–30.
nology Support Project Program of the Ministry of Science and [29] Rollett AD, Kocks UF, Doherty RD. Stage IV work hardening in cubic metals.
Technology of China (No. 2011BAE22B02 and No. Symp on formability metall struct. The Metallurgical Society, AIME; 1987. p.
2011DFA50907), and the project (No. 51171113) received from 211–25.
[30] Haasen P. A cell theory for stage IV work hardening of metals and
the National Natural Science Foundation of China. This investiga-
semiconductors. J Phys France 1989;50:2445–54.
tion involves part of Canada–China–USA Collaborative Research [31] Yi S, Bohlen J, Heinemann F, Letzig D. Mechanical anisotropy and deep drawing
Project on the Magnesium Front End Research and Development behaviour of AZ31 and ZE10 magnesium alloy sheets. Acta Mater
2010;58:592–605.
(MFERD). The authors also thank Q. Li, C. Ma, A. Machin, J. Amank-
[32] Partridge PG. Crystallography and deformation modes of hexagonal close
rah, and R. Churaman for assisting in the experiments. packed metals. Met Rev 1967;12:169–94.
[33] Koike J, Kobayashi T, Mukai T, Watanabe H, Suzuki M, Maruyama K, et al. The
activity of non-basal slip systems and dynamic recovery at room
References temperature in fine-grained AZ31B magnesium alloys. Acta Mater
2003;51:2055–65.
[1] Shakun JD, Clark PU, He F, Marcott SA, Mix AC, Liu ZY, et al. Global warming [34] Dobron P, Chmelik F, Yi S, Parfenenko K, Letzig D, Bohlen J. Grain size effects on
preceded by increasing carbon dioxide concentrations during the last deformation twinning in an extruded magnesium alloy tested in compression.
deglaciation. Nature 2012;484:49–54. Scripta Mater 2011;65:424–7.
418 F.A. Mirza et al. / Materials and Design 46 (2013) 411–418

[35] Yin SM, Wang CH, Diao YD, Wu SD, Li SX. Influence of grain size and texture on [38] Kocks UF, Mecking H. Physics and phenomenology of strain hardening: the FCC
the yield asymmetry of Mg–3Al–1Zn Alloy. J Mater Sci Technol case. Prog Mater Sci 2003;48(3):171–273.
2011;27(1):29–34. [39] Argon AS, Haasen P. A new mechanism of work hardening in the late stages of
[36] Stanford N, Barnett MR. Effect of particles on the formation of deformation large strain plastic flow in fcc and diamond cubic crystals. Acta Met Mater
twins in a magnesium-based alloy. Mater Sci Eng A 2009;516:226–34. 1993;41(1):3289–306.
[37] Yakubtsov IA, Diak BJ, Sager CA, Bhattacharya B, MacDonald WD, Niewczas M. [40] Nix W, Gibeling J, Hughes D. Time-dependent deformation of metals. Met
Effects of heat treatment on microstructure and tensile deformation of Mg Trans 1985;16A:2215–26.
AZ80 alloy at room temperature. Mater Sci Eng A 2008;496:247–55.

You might also like