You are on page 1of 11

Materials and Design 56 (2014) 174–184

Contents lists available at ScienceDirect

Materials and Design


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

Effect of strain rate and temperature on strain hardening behavior


of a dissimilar joint between Ti–6Al–4V and Ti17 alloys
S.Q. Wang a,b, J.H. Liu a,⇑, D.L. Chen b,⇑
a
State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, 127 Youyi Road, Xi’an 710072, PR China
b
Department of Mechanical and Industrial Engineering, Ryerson University, 350 Victoria Street, Toronto, Ontario M5B 2K3, Canada

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 evaluate the influence of strain rate and temperature on the tensile proper-
Received 21 August 2013 ties, strain hardening behavior, strain rate sensitivity, and fracture characteristics of electron beam
Accepted 3 November 2013 welded (EBWed) dissimilar joints between Ti–6Al–4V and Ti17 (Ti–5Al–4Mo–4Cr–2Sn–2Zr) titanium
Available online 13 November 2013
alloys. The welding led to significant microstructural changes across the joint, with hexagonal close-
packed martensite (a0 ) and orthorhombic martensite (a00 ) in the fusion zone (FZ), a0 in the heat-affected
Keywords: zone (HAZ) on the Ti–6Al–4V side, and coarse b in the HAZ on the Ti17 side. A distinctive asymmetrical
Titanium alloy
hardness profile across the dissimilar joint was observed with the highest hardness in the FZ and a lower
Electron beam welding
Strain hardening behavior
hardness on the Ti–6Al–4V side than on the Ti17 side, where a soft zone was present. Despite a slight
Strain rate reduction in ductility, the yield strength (YS) and ultimate tensile strength (UTS) of the joints lay in-
Temperature between the two base metals (BMs) of Ti–6Al–4V and Ti17, with the Ti17 alloy having a higher strength.
While the YS, UTS, and Voce stress of the joints increased, both hardening capacity and strain hardening
exponent decreased with increasing strain rate or decreasing temperature. Stage III hardening occurred in
the joints after yielding. The hardening rate was strongly dependent on the strain rate and temperature.
As the strain rate increased or temperature decreased, the strain hardening rate increased at a given true
stress. The strain rate sensitivity evaluated via both common approach and Lindholm approach was
observed to decrease with increasing true strain. The welded joints basically failed in the Ti–6Al–4V
BM near the HAZ, and the fracture surfaces exhibited dimple fracture characteristics at different
temperatures.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction welding (EBW) is considered to be an ideal process to join titanium


and its alloys. The high vacuum inside the chamber where the pro-
Titanium and its alloys have been increasingly used as one of cess is carried out protects the hot metal from contamination.
key structural materials in the aerospace applications due to their Moreover, a narrow and deep joint can be achieved with a high
excellent combination of properties such as high strength-to- beam power density, when compared to the conventional welding
weight ratio, high fatigue life, toughness, excellent resistance to processes [9]. A relatively small heat-affected zone (HAZ) is ob-
corrosion [1–5]. The structural application of titanium alloys tained in this process while producing lower thermal stresses com-
unavoidably involves welding and joining [6]. The welding of tita- pared to other welding techniques [9,10]. Therefore, EBW is
nium alloy is fairly challenging, since it becomes very reactive with recognized as a promising process for joining relatively thick,
most atmospheric gases such as oxygen, nitrogen, carbon, and refractory metals and their dissimilar combinations. The previous
hydrogen at temperatures above 550 °C and particularly in the studies about the electron beam welded (EBWed) titanium alloy
molten state [7]. While these interstitial elements increase the joints mainly involved the effects of heat treatment and welding
strength, they reduce ductility and toughness and cause severe parameters on the microstructure, hardness, tensile properties,
embrittlement [7,8]. Poor preparation and cleaning of the base me- and fatigue properties [11–18]. For example, Thomas et al. [14]
tal (BM) before welding, poor shielding of the joint or impurities studied the effect of pre- and post-weld heat treatment on the
present in the shielding gas during welding can also cause serious mechanical properties of EBWed Ti–6AI–4V joints. Pederson et al.
contamination [7]. To overcome these problems, electron beam [15] reported the effect of different beam scanning patterns on
the microstructure and mechanical properties of EBWed Ti–6Al–
4V joints. Fu et al. [16] and Wang et al. [17,18] further evaluated
⇑ Corresponding authors. Tel.: +1 416 979 5000x6487; fax: +1 416 979 5265
the strain-controlled low cycle fatigue behavior of EBWed titanium
(D.L. Chen), tel.: +86 (029) 8849 2624 (J.H. Liu).
E-mail addresses: jinhliu@nwpu.edu.cn (J.H. Liu), dchen@ryerson.ca (D.L. Chen).
alloy joints. However, it is unclear about the strain hardening

0261-3069/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.matdes.2013.11.003
S.Q. Wang et al. / Materials and Design 56 (2014) 174–184 175

behavior of EBWed titanium alloy joints, although a few studies of 6 mm were machined perpendicularly to the welding direction
have been reported on the strain hardening behavior of other types using electro-discharge machining (EDM). The gauge area was
of welded joints including laser welded DP980 steel joints [19] and ground up to #600 SiC papers to remove the EDM cutting marks
Mg joints [20,21], and friction stir welded Al joints [22] and Mg and to achieve a smooth and consistent surface. Tensile tests were
joints [23–27]. Therefore, it is necessary to evaluate the strain performed using a computerized United tensile testing machine at
hardening behavior of the EBWed titanium alloy joints. strain rates between 102 s1 and 105 s1 at varying test temper-
Strain hardening (or work hardening) is one of the most impor- atures from 40 to 180 °C. Three specimens in each case were
tant considerations in the evaluation of plastic deformation of tested to check the reproducibility of the test data. Fracture sur-
metallic materials [23,28]. The strength, ductility, toughness and faces after tensile testing were examined using SEM to identify
deformability of materials are intimately related to strain harden- the fracture mechanisms.
ing characteristics [29]. For this reason, many investigations have
been carried out on the strain hardening behavior and physical
3. Results and discussion
mechanism of conventional metallic materials [30–32]. The strain
hardening behavior of cubic metals is fairly well understood, and
3.1. Microstructure
the accumulation of a forest of dislocations is the predominant
hardening mechanism [31,32]. HCP metals present a more complex
Fig. 1 shows the microstructures of BMs of Ti–6Al–4V and Ti17,
case due to their low symmetry which restricts the number of slip
respectively. Ti–6Al–4V alloy was composed of globular a grains
systems and their strong plastic anisotropy [31,33,34]. Hitherto
and colonies of secondary elongated a grains surrounded by inter-
many studies have been reported on the strain hardening behavior
granular b phase (Fig. 1(a)), while Ti17 consisted of fine a plates
of Ti alloys [35–38]. For example, Honarmandi and Aghaie-Khafri
embedded in the b matrix (Fig. 1(b)). Fig. 2 shows a significant
[37], and Rao and Tangri [38] studied the strain hardening behavior
microstructural change occurred in the fusion zone (FZ) and HAZ
of Ti–6Al–4V and TiAlx alloy, respectively, by means of compres-
of the joint after EBW. The FZ was narrow (2 mm wide) and
sion testing at high temperatures and observed different stages
was mainly composed of acicular and fine martensite a0 and a00
of strain hardening, while Bystrzanowski et al. [36] investigated
phase (Fig. 2(c)) due to the fast cooling during EBW. Akhonin
the tensile flow behavior of Ti–46Al–9Nb sheet material and ob-
et al. [39] also observed the fine acicular microstructure in the FZ
served only stage III strain hardening at high temperatures. In spite
of EBWed titanium alloy joints. This was, however, in contrast to
of many studies on the strain hardening behavior of titanium alloys
the microstructure of EBWed Ti–5Al–5V–5Mo–3Cr joint where
and other alloys, as well as the strain hardening behavior of several
the FZ consisted of large grains of b phase due to the high cooling
welded joints including laser welded DP980 steel joints [19] and
rate [40] and the presence of b-stabilizing alloying elements
Mg joints [20,21], and friction stir welded Al joints [22] and Mg
[18,40,41]. Unlike the similar EBWed joint of a titanium alloy,
joints [23–27], to the authors’ knowledge no information on the
the microstructures in the HAZ of the dissimilar Ti–6Al–4V/Ti17
strain hardening behavior of EBWed dissimilar titanium alloy
joint on both sides became very different, i.e., coarse b phase
joints is available in the open literature. It is unknown what strain
(Fig. 2(b)) present on the Ti17 side and acicular martensite a0
hardening stages would occur, and how the strain rate and test
formed on the Ti–6Al–4V side (Fig. 2(d)). The formation of b phase
temperature affect the strain hardening behavior in an EBWed dis-
in the HAZ on the Ti17 side was apparently associated with the
similar joint of titanium alloys. This study was, therefore, aimed at
more b-stabilizing alloying elements (e.g., Cr, Mo [42], Sn, Zr
evaluating the tensile properties and identifying the strain harden-
[43]) existing in the Ti17 alloy as shown in Table 2. Similar large
ing behavior in an EBWed dissimilar joint between Ti–6Al–4V and
b grains in the HAZ of EBWed Ti–5Al–5V–5Mo–3Cr joints were also
Ti17 alloys at varying strain rates and temperatures.
reported in [40,41] as a result of a fairly large amount of b-stabiliz-
ing alloying elements present in the alloy. It was observed that
2. Material and experimental procedure martensite a0 in the FZ was much finer than that in the HAZ on
the Ti–6Al–4V side, as seen from Fig. 2(c) and (d). More detailed re-
The materials used in the present study were forged Ti–6Al–4V sults on the microstructural changes in the dissimilar Ti–6Al–4V/
and Ti17 titanium alloys of 10 mm in thickness, with their chemi- Ti17 joint after EBW could be seen in our earlier report [18]. The
cal compositions listed in Tables 1 and 2, respectively. Both alloys HAZ microstructure on the Ti–6Al–4V side of the present dissimilar
were machined into the plates with a dimension of joint resembled that of EBWed Ti–6Al–4V similar joint reported by
140 mm  80 mm  10 mm, and then mechanically and chemi- Lu et al. [44].
cally cleaned before welding. EBW was carried out using HDZ-
15B EBW machine with an accelerating voltage (V) of 60 kV, a
3.2. Microhardness
beam current (Ib) of 68 mA, a focus current (If) of 2230 mA, and a
welding speed (v) of 500 mm/min.
Fig. 3 shows a Vickers microhardness profile along the mid-
Metallographic samples were cut from the EBWed workpieces
thickness across the dissimilar welded joint between Ti–6Al–4V
perpendicular to the welding direction, then ground, polished
and Ti17 alloys. It is seen that a distinctive asymmetrical hardness
and etched using Keller’s reagent (12 ml HF, 36 ml HNO3 and
profile across the dissimilar weld was present with an average
42 ml H2O). Microstructures were examined via scanning electron
hardness value of 330 HV for Ti–6Al–4V BM and 375 HV for
microscopy (SEM) using a JEOL JSM-6380LV microscope equipped
Ti17 BM, respectively. Compared with the Ti–6Al–4V BM, the high-
with Oxford energy dispersive X-ray spectroscopy (EDS) system
er hardness in the Ti17 BM was related to the overall higher alloy-
and 3D fractographic analysis capacity.
ing elements in this alloy (Table 2) and the ensuing microstructure
Sub-sized tensile specimens of 140 mm long with a gauge
(Fig. 1(b)) where the presence of a large number of fine a plates of-
length of 25 mm (or a parallel length of 32 mm) and gauge width
fered a strong resistance to the motion of dislocations during plas-
Table 1 tic deformation. The appreciably higher hardness value in the FZ
Chemical composition of Ti–6Al–4V titanium alloy. was obviously attributed to the formation of fine martensite in
Element Al V Fe C N H O Ti the FZ, as shown in Fig. 2(c). Similar results were also reported
by Wang and Wu [12] and Lu et al. [44] in the EBWed Ti–6Al–4V
Content (wt.%) 6 4 0.3 0.1 0.05 0.015 0.2 Balance
joint. As seen from Fig. 3, the HAZ was wider on the Ti17 side than
176 S.Q. Wang et al. / Materials and Design 56 (2014) 174–184

Table 2
Chemical composition of Ti17 titanium alloy.

Element Al Mo Cr Sn Zr Fe C N H O Ti
Content (wt.%) 5 4 4 2 2 0.3 0.05 0.04 0.012 0.15 Balance

(a) (b)

α α
β
β

Fig. 1. Microstructures of (a) Ti–6Al–4V BM and (b) Ti17 BM.

(a) Top
(b)

b c d

Ti17 BM Ti-6Al-4V BM

Bottom 2mm

(c) (d)

Fig. 2. Microstructures of the joint, (a) overall view of the cross section, (b) HAZ of Ti17, (c) center of FZ, and (d) HAZ of Ti–6Al–4V.

on the Ti–6Al–4V side, corresponding well to Fig. 2(a). It was also basically smooth and continuous stress–strain curves. Semiatin
observed that the hardness value in the HAZ on the Ti–6Al–4V side and Bieler [46] and Martin et al. [47] have also reported similar
was slightly higher than that of the Ti–6Al–4V BM. However, the
hardness value in the HAZ on the Ti17 side was visibly lower than 500
that of the Ti17 BM, indicating that a soft zone emerged. Similar re-
Vickers hardness, HV

sult was also reported by Tan et al. [45] on an EBWed joint of Ti–
22Al–25Nb and TC11 alloys. The occurrence of such a soft zone
400
was attributed to the presence of relatively soft and coarse b phase
soft zone
in the HAZ on the Ti17 side, as shown in Fig. 2(b). While the hard-
ness in the soft zone increased with increasing distance from the
FZ boundary up to that of Ti17 BM, it still had a slightly higher 300
hardness value than that of Ti–6Al–4V BM.

Ti-6Al-4V-BM HAZ FZ HAZ Ti17-BM


3.3. Tensile properties
200
-8 -6 -4 -2 0 2 4 6 8
3.3.1. Effect of strain rate and temperature on YS and UTS
Distance from weld center, mm
Fig. 4(a) shows typical stress–strain curves obtained for BMs of
Ti–6A1–4V and Ti17, and their dissimilar joint at a strain rate of Fig. 3. Hardness profile across the mid-thickness of the dissimilar welded joint
1  103 s1 at room temperature. The BMs and joint showed between Ti–6Al–4V and Ti17 alloys.
S.Q. Wang et al. / Materials and Design 56 (2014) 174–184 177

smooth and continuous stress–strain curves for Ti–6Al–4V and own stress–strain curve as indicated in the Fig. 4(b), where BM2
Ti17 alloys. However, this was in contrast to the tensile stress– is stronger or harder than BM1 (In the present study, BM1 corre-
strain curves obtained in a laser-welded joint of DP600 steel [19], sponds to the Ti–6Al–4V alloy, and BM2 corresponds to the Ti17 al-
where an obvious yield point phenomenon occurred at all the loy). Upon loading, BM1, BM2, and the dissimilar joint initially
strain rates due to the interstitial diffusion in the steel during the deform elastically as indicated by the linear portion (OA) of the
laser welding process. It is also seen from Fig. 4(a) that the Ti– curves in Fig. 4(b). Beyond point A at eys1, BM1 will start to yield
6Al–4V alloy was more ductile than the Ti17 alloy, despite its low- and deform plastically since it has the lowest hardness value (as
er flow stress than that of Ti17 alloy. This was directly associated shown in Fig. 3). An intuitive thinking would be that if failure oc-
with the microstructure and grain morphology (Fig. 1). The lamel- curs in the BM1, the dissimilar joint would yield at point A. How-
lar microstructure was usually preferable for the strength, fracture ever, due to the smooth and continuous plastic deformation and
toughness, fatigue crack propagation, and oxidation behavior, strain hardening of BM1, no visible deviation from the initial linear
while the globular microstructure was better for the ductility and portion on the stress–strain curve of the dissimilar joint could be
fatigue crack initiation [48,49]. As seen from Fig. 1, Ti–6Al–4V alloy seen until point B at eys2 is reached where the HAZ on the BM2 side
consisted of more globular a, while Ti17 exhibited a lamellar starts to yield due to the presence of soft zone (SZ) as indicated in
microstructure as mentioned above. Therefore, the ductility of the Fig. 4(b). Then the dissimilar joint deforms plastically, and the
Ti–6Al–4V alloy was higher than that of Ti17 alloy, while the plastic strain is contributed mainly by BM1 and moderately by the
strength of Ti–6Al–4V alloy was lower than that of Ti17 alloy. HAZ (or SZ) on the BM2 side. Therefore, the stress–strain curve of
In general, the strength of a welded joint is supposed to be low- the dissimilar joint is positioned slightly above BM1 (or Ti–6Al–
er than that of the pertinent BMs, as also reported by Wang et al. 4V BM). Whether the dissimilar joint could reach point C (the yield
[12] and Thomas et al. [14]. However, a special or seemingly abnor- point of BM2 at eys3) depends on the difference of yield strengths
mal characteristic was observed in Fig. 4(a), i.e., the stress–strain between BM2 and BM1 (or BM2 and HAZ on the BM2 side). If this
curve of the dissimilar joint lay in-between those of the two BMs difference is big, point C may never be reached even though the
of Ti–6Al–4V and Ti17. It should be noted that this behavior was dissimilar joint fails. It would be reasonable to assume that no
reproducible in the present EBWed dissimilar joint since three yielding or plastic deformation occurs in the FZ due to the presence
samples were tested in all cases in this study. Similar abnormal of fine martensite (Fig. 2(c)) and the resulting high hardness
behavior was also reported by Farabi et al. [19] in the laser-welded (Fig. 3). Indeed, the above discussion could be further understood
DP600 joints and by Ma et al. [50] in the linear friction welded Ti– via a concept so-called the limiting strength ratio (LSR) [51], which
6Al–4V/Ti–6.5Al–3.5Mo–1.5Zr–0.3Si joints. To better understand corresponds to a special condition assuming that the yield strength
the above-observed behavior, a schematic diagram is plotted in of BM2, rys;BM2 , is identical to the ultimate tensile strength of BM1,
Fig. 4(b), where an additional schematic illustration of a dissimilar rUTS;BM1 . That is,
welded joint consisting of BM1 and BM2 is embedded within rys;BM2
Fig. 4(b) as well. Let us first consider that both BMs have their LSR ¼ ¼ 1: ð1Þ
rUTS;BM1
1200 It should be noted that based upon the force balance requirement,
(a) i.e., FBM1 = FBM2 (Fig. 4(b)), the LSR simply becomes the ratio of
Engineering stress, MPa

1000 the cross-sectional area of the weaker material BM1 to that of the
stronger material BM2. In the present study, the cross-sectional
800 area or thickness of both BMs is the same, and the actual strength
ratio of the two BMs of Ti17 and Ti–6Al–4V (rys;Ti17 /rUTS;Ti—6Al—4V ) ex-
600 ceeds significantly the LSR = 1, as shown in Fig. 4(a). It follows that
the BM2 (or Ti17) with a higher strength will be limited to elastic
400 deformation only even though the BM1 (or Ti–6Al–4V) with a lower
Ti-6Al-4V-BM strength exceeds its UTS.
200 Ti17-BM Fig. 5 shows the typical stress–strain curves of the Ti–6Al–4V/
Joint Ti17 dissimilar joints tested at different strain rates at room tem-
0 perature and at different temperatures at a constant strain rate
0 2 4 6 8 10
of 1  103 s1, respectively. As seen from Fig. 5(a), the flow curves
Engineering strain, %
shifted higher with increasing strain rate, leading to a higher yield
strength (YS) and ultimate tensile strength (UTS). Venkatesh et al.
(b) [52] also reported a similar changing trend of strain rate on the
BM2 tensile curves for a Ti–6Al–4V ELI (extra low interstitials) alloy
subjected to different heat treatment processes. It is seen from
C× Joint
Fig. 5(b) that with increasing test temperature the curves shifted
Stress (σ)

BM1 lower significantly which led to a lower strength but a better duc-


tility. This suggested a strong dependence of the flow behavior of
the dissimilar joint on temperature. These results were similar to
FZ
those of Ward et al. [53] for a Ti–25Al–10Nb–3V–1Mo alloy. It is
FBM1 FBM2 also observed that in all the cases a small slope of the curves of
BM1 BM2 the joint in the plastic range was present, indicating a relatively
HAZ HAZ (SZ)
low strain-hardening rate which is common in most titanium al-
O loys [54].
Strain ( )
ys1 ys2 ys3 Fig. 6 shows the change of YS and UTS of two BMs and their dis-
similar joint with the stain rate at room temperature. As seen from
Fig. 4. (a) Typical tensile stress–strain curves of Ti–6Al–4V BM, Ti17 BM, and their
joint, respectively, at a strain rate of 1  103 s1 at room temperature and (b) a
the figure, the YS and UTS of both BMs and their dissimilar joint all
schematic illustration of the stress–strain curves of a dissimilar joint and their increased with increasing strain rate, suggesting a strong strain-
pertinent BMs. rate dependence of both YS and UTS. For example, the average YS
178 S.Q. Wang et al. / Materials and Design 56 (2014) 174–184

1200 1200
(a) (a)
Engineering stress, MPa

1000

800 1000

YS, MPa
600

400 800
1 ×10 ² s ¹
1 ×10 ³ s ¹ Ti-6Al-4V-BM
200 1 ×10 s ¹ Ti17-BM
1 ×10 s ¹ Joint
0 600
0 2 4 6 8 10 1E-6 1E-5 1E-4 1E-3 1E-2 1E-1
Engineering strain, % Strain rate, s-1

1200 1200
(b) (b)
Engineering stress, MPa

1000

1000

UTS, MPa
800

600

800
400
Ti-6Al-4V-BM
-40°C
Ti17-BM
200 20°C
Joint
180°C
600
0
1E-6 1E-5 1E-4 1E-3 1E-2 1E-1
0 2 4 6 8 10
Strain rate, s-1
Engineering strain, %
Fig. 6. Effect of strain rate on (a) YS and (b) UTS of the BMs and dissimilar joints at
Fig. 5. Tensile stress–strain curves of Ti–6Al–4V/Ti17 dissimilar joints tested at (a) room temperature.
different strain rates at room temperature, and (b) different temperatures at a
constant strain rate of 1  103 s1.

in the flow stress with increasing temperature confirmed that the


alloy was temperature sensitive. Chiou et al. [61] reported that Ti–
of the dissimilar joint increased from about 800 MPa at a strain 5.2Zr–2.9Al–1.1Mo–0.35Fe alloy was highly sensitive to the test
rate of 1  105 s1 to about 916 MPa at a strain rate of temperature in the ranges of 25–500 °C and 700–900 °C, respec-
1  102 s1, and the corresponding average UTS of the dissimilar tively, in a dynamic impact test. Vanderhasten et al. [62] also pre-
joint increased from about 865 MPa to about 965 MPa in the same sented the same trend for Ti–6Al–4V alloy from room temperature
strain range. A similar strain rate dependence of the strength was up to about 650 °C, and they believed that the deformation mech-
also reported for Ti–6Al–4V alloy by López et al. [55]. Zhou and anism was associated with the classical cold and warm deforma-
Chew [56] reported that the strain rate dependent behavior was tion. However, Ul’yanov and Moskalenko [63] studied the
caused by the change of governing plastic deformation mechanism deformation mechanism at low temperature from 269 °C to
from dislocation slip at low strain rates to twinning at the high 40 °C and noted that twinning was the deformation mechanism.
strain rate for Ti–6Al–4V alloy. Indeed the increase of the strength Indeed, it was observed that the effects of strain rate and temper-
with increasing strain rate could be explained by the increase in
the density of dislocations and the occurrence of twinning in tita-
nium alloys [57–60]. It is also seen from Fig. 6 that the YS and UTS
of the dissimilar joint lay just in-between those of the BMs of Ti– 1200
6Al–4V and Ti17, with Ti17 alloy having a higher strength at a gi- YS
ven strain rate. As an example, the average YS of the joint increased UTS
by 9% compared to the Ti–6Al–4V alloy, and decreased by 12%
Strength, MPa

compared to the Ti17 alloy tested at a strain rate of 1  103 s1 1000
at room temperature. It should be noted that the error bars given
on the data points in Fig. 6 and the subsequent figures indicated
the standard deviation based on the test results of three samples
in all cases. 800
Fig. 7 shows the YS and UTS of the Ti–6Al–4V/Ti17 dissimilar
joint as a function of test temperature at a strain rate of
1  103 s1. With increasing temperature both YS and UTS of
the dissimilar joint decreased markedly. The average YS and UTS 600
-100 -50 0 50 100 150 200 250
of the joint at 180 °C decreased by 22% and 19%, respectively, com-
pared to those at room temperature at a strain rate of 1  103 s1, Temperature, oC
while they had an increase of 12% and 9% at 40 °C, respectively, Fig. 7. Effect of test temperature on YS and UTS of the Ti–6Al–4V/Ti17 dissimilar
compared to those at room temperature. In general, the reduction joints at a strain rate of 1  103 s1.
S.Q. Wang et al. / Materials and Design 56 (2014) 174–184 179

0.11 change in the dislocation density during plastic deformation in the


(a) Room temp.
uniform deformation stage from YS and UTS [23,67–69]. The same
trend in relation to the strain rate was also reported by Zhao et al.
Hardening capacity

0.09 [67,70] in NiAl-matrix composites and Xuan et al. [71] in 50 vol%


TiCx/Al and TiB2/Al composites tested in compression.
Fig. 9 shows the effect of strain rate and temperature on the
0.07 strain hardening exponent of the dissimilar joints evaluated on
the basis of a modified relationship proposed by Afrin et al. [23],

0.05 r  ry ¼ K  ðe  ey Þn ; ð3Þ
 
where ry is the YS, ey is the yield strain, K is a constant and n is the
strain hardening exponent. In Eq. (3) the elastic component in both
0.03
1E-6 1E-5 1E-4 1E-3 1E-2 1E-1 stress and strain is indeed excluded. It is seen that with increasing
strain rate at both room temperature and 180 °C or with decreasing
Strain rate, s-1
temperature, n values decreased. This was also reported to be the
0.12 case by Venkatesh et al. [52] for Ti–6Al–4V ELI alloy in different
(b) Strain rate 1×10-3 s-1 heat treatment conditions. The change of strain hardening exponent
0.1 with strain rate and temperature corresponded well to that of hard-
ening capacity with strain rate and temperature (Fig. 8).
Hardening capacity

0.08 Another way of characterizing the strain hardening behavior is


the strain hardening rate or the slope of true stress–true strain
0.06 (h = dr/de), which is plotted as a function of true stress in Fig. 10
showing the effect of strain rate and temperature in the dissimilar
0.04 joints. It is of interest to note that only hardening stage III was ob-
served in the welded joints in all the cases. The presence of stage III
0.02 was also observed by Bystrzanowski et al. [36] in a Ti–46Al–9Nb
alloy, and Choudhary and Palaparti [72] in a 9Cr–1Mo ferritic steel,
0 especially at high temperatures. However, it was significantly dif-
-100 -50 0 50 100 150 200 250 ferent from Mg alloy which had hardening stage II, III and IV
Temperature, oC [23,24,36,68,69]. This may be related to the magnitude of stacking
fault energy (SFE). The SFE of titanium (0.35 J/m2 [73,74]) was
Fig. 8. Effect of (a) strain rate, and (b) test temperature on the hardening capacity of
higher than that of magnesium (0.125 J/m2 [75]), and Kocks and
the Ti–6Al–4V/Ti17 dissimilar joints.

0.8
ature were interrelated [64]. Decreasing temperature had a similar
effect to increasing strain rate, as shown in Figs. 5–7. These trends
(a) Room temperature
180°C
were also in agreement with the results reported by Domiaty [65], 0.7
Isaac and Choudhary [66].
0.6
n*

3.3.2. Effect of strain rate and temperature on strain hardening


behavior 0.5
The hardening capacity, Hc, of a material may be simply defined
as follows [23],
0.4
rUTS  ry rUTS
Hc ¼ ¼  1: ð2Þ
ry ry 0.3
1E-6 1E-5 1E-4 1E-3 1E-2 1E-1
Following Eq. (2), the evaluated hardening capacity of the dis-
Strain rate, s-1
similar joint tested at different strain rates and temperatures is
plotted in Fig. 8. With increasing strain rate or decreasing test tem-
0.6
perature, the hardening capacity decreased. Based on Eq. (2) the (b) Strain rate 1×10-3 s-1
hardening capacity of a material was related to its YS. As seen from
Fig. 6, the slope of the YS is slightly steeper than that of the UTS 0.5
with increasing strain rate, therefore the Hc decreased with
increasing strain rate as can be seen from Fig. 8. For example, the
n*

YS increased from about 800 MPa to about 916 MPa when the 0.4
strain rate increased from 1  105 s1 to 1  102 s1, and the cor-
responding value of Hc decreased from 0.08 to 0.05. The YS in-
creased from about 690 MPa to about 986 MPa with decreasing 0.3
temperature from 180 °C to 40 °C at a constant strain rate of
1  103 s1, and the resulting value of Hc decreased from 0.09 to
0.2
0.02. Thus the test temperature had a more significant effect on -100 -50 0 50 100 150 200 250
the hardening capacity compared to the strain rate. This could be
Temperature, oC
better seen from Fig. 8(a), where the change of hardening capacity
with increasing strain rate reached a plateau (or flat level). The var- Fig. 9. Effect of (a) strain rate at different temperatures and (b) test temperature
iation of Hc was also related to the grain size in a material and the on n.
180 S.Q. Wang et al. / Materials and Design 56 (2014) 174–184

20000 1100
(a) 1 ×10 ² s ¹ (a)
Strain hardening rate, MPa

1 ×10 ³ s ¹
16000

Voce stress, MPa


1 ×10 s ¹
1 ×10 s ¹ 1000
12000

8000
900

4000

800
0 1E-6 1E-5 1E-4 1E-3 1E-2 1E-1
700 800 900 1000 1100 1200
Strain rate, s-1
True stress, MPa
1200
20000
-40°C (b)
(b)
Strain hardening rate, MPa

20°C

Voce stress, MPa


16000
180°C 1000

12000

800
8000

4000
600
-100 -50 0 50 100 150 200 250
0
600 700 800 900 1000 1100 1200 Temperature, ºC
True stress, MPa Fig. 11. Effect of (a) strain rate at room temperature and (b) test temperature at a
strain rate of 1  103 s1 on the Voce stress.
Fig. 10. The strain-hardening rate vs. true stress at different (a) strain rates at room
temperature, and (b) temperatures at a strain rate of 1  103 s1.
7
(a)
Mecking [31] noted that the presence of stage III could limit the ex-
tent of stage II or eliminate it as a separate stage, especially at high 6.9
temperatures and for materials of higher SFE. Therefore only stage
σ (MPa)

III was observed in the present dissimilar joints.


It is also observed that the strain hardening rate is strongly 6.8
lnσ

dependent on the strain rate and temperature as can be deduced


from Fig. 10. For example, h decreased almost linearly with increas-
6.7 1% True strain
ing flow stress. As the strain rate increased or the temperature de-
3% True strain
creased, the strain-hardening rate h increased at a given true stress.
5% True strain
This could be understood on the basis of the following equation 6.6
[31,76], -14 -12 -10 -8 -6 -4 -2
ln (Strain rate), s -1
Rd
h ¼ ho  r; ð4Þ
e_ 1=q 1050
(b)
where Rd and q are temperature dependent constants, which are 1000
independent of stress and strain rate [31], h is the strain-hardening
Ture stress, MPa

rate in stage III, and ho is considered as a constant. It is clear that as 950


the strain rate increased, the strain-hardening rate h increased. In
addition, according to the Voce law, the strain hardening rate and 900
true stress can also be expressed as [31],
850
ho 1% True strain
h ¼ ho  r; ð5Þ 800
rv 3% True strain
5% True strain
where rv is referred to as Voce stress or saturation stress at high 750
-6 -5 -4 -3 -2 -1
strains corresponding to the condition where the instantaneous
strain hardening rate is equal to zero, i.e., h = 0. The evaluated rv log (Strain rate), s -1
on the basis of Eq. (5) and Fig. 10 is plotted in Fig. 11. The Voce Fig. 12. The plot used to evaluate the strain-rate sensitivity index, (a) m, via the
stress had a similar change to that of the YS and UTS of the dissim- common approach, and (b) mL, via the Lindholm’s approach for the joints tested at
ilar joints as a function of strain rate or temperature (Figs. 6 and 7). room temperature.
S.Q. Wang et al. / Materials and Design 56 (2014) 174–184 181

0.03 That is, with increasing strain rate or decreasing temperature, the
Strain-rate sensitivity, m (a) Voce stress increased.

3.3.3. Effect of strain rate and temperature on strain rate sensitivity


0.02
Fig. 12(a) shows the curves of ln(r) vs. ln(strain rate) for the dis-
similar joints at different true strain values at room temperature,
where the slope indicates the strain rate sensitivity, m, in the com-
0.01
mon approach [77,78]. The obtained strain rate sensitivity is
shown in Fig. 13(a). It is seen that the strain rate sensitivity de-
creased with increasing true strain at room temperature. Similar
results have also been reported by Lin and Chen [79], and Shen
0 et al. [80] for Mg alloy and ultrafine-crystalline Cu.
0 1 2 3 4 5 6 7 The Lindholm approach [81] is also used to evaluate the strain
True strain, % rate sensitivity based on the following equation:

r ¼ ro ðeÞ þ r1 ðeÞ logðe_ Þ; ð6Þ


60
(b) where the Lindholm strain rate sensitivity, mL, is the slope, r1 ðeÞ, of
Strain-rate sensitivity, mL

:
50 r vs. logðeÞ based on the above equation. A typical plot of r vs.
logðeÞ at different true strain values is shown in Fig. 12(b), and
_
40
the evaluated mL [or r1 ðeÞ] is presented in Fig. 13(b). Such a strain
rate sensitivity also decreased as the true strain increased. In spite
30
of the similar change, the Lindholm strain rate sensitivity values
were much larger than the strain rate sensitivity values evaluated
20
via the common approach. Similar results were also reported in
10 Ti–6Al–4V [82], extruded AZ31 magnesium alloy [79] and ultra-
fine-crystalline Cu [80].
0
0 1 2 3 4 5 6 7
3.4. Fractography
True strain, %
The tensile failure of the Ti–6Al–4V/Ti17 dissimilar joint sam-
Fig. 13. Strain-rate sensitivity as a function of true strain for the joints tested at
ples was observed to occur in the Ti–6Al–4V BM, which had the
room temperature, evaluated using (a) common approach and (b) Lindholm’s
approach. lowest hardness across the welded joint as seen from the microh-
ardness profile in Fig. 3. No samples failed in the FZ, indicating that

Fig. 14. SEM images showing the fracture surfaces of the joints tested at room temperature at a strain rate of (a) 1  102 s1, and (b) 1  105 s1, where the failure occurred
in the Ti–6Al–4V BM.

Fig. 15. SEM images showing the fracture surfaces of the joints tested at 180 °C at a strain rate of (a) 1  102 s1, and (b) 1  105 s1, where the failure occurred in the Ti–
6Al–4V BM.
182 S.Q. Wang et al. / Materials and Design 56 (2014) 174–184

Fig. 16. SEM images showing the fracture surfaces of the joints tested at 40 °C at a strain rate of 1  103 s1 (a) an overall view, and (b) magnified area of the dashed box in
(a), where the failure occurred in the Ti–6Al–4V BM.

the strength of FZ was much higher than that of Ti–6Al–4V BM due 5. The strain rate sensitivity evaluated via both the common
to the formation of martensite (Fig. 2(c)) and the resultant high approach and Lindholm approach decreased as the true
hardness (Fig. 3). Similar results were also reported on the tensile strain increased.
tests of high current MIG welded Al alloy [83,84] and linear friction 6. The welded joints failed in the Ti–6Al–4V BM side near the
welded Ti–6Al–4V/Ti–6.5Al–3.5Mo–1.5Zr–0.3Si joints [51], where HAZ, and the fracture surfaces were characterized by typi-
the failure occurred in the BM as well. Fig. 14 shows typical images cal dimple fracture characteristics at different
of the fracture surfaces of the joints tested at strain rates of temperatures.
1  102 s1 and 1  105 s1 at room temperature. The fracture
was predominantly characterized by the dimple-like ductile frac-
ture characteristics with a large number of tear ridges at both high
and low strain rates. Similar dimple fracture characteristics were Acknowledgements
also observed in the joints tested at the higher temperature as
shown in Fig. 15 and in [52] for a Ti–6Al–4V ELI alloy. At a lower The authors would like to thank the Natural Sciences and Engi-
temperature of 40 °C, similar dimples appeared as well. The frac- neering Research Council of Canada (NSERC) for the financial sup-
ture surfaces of the joints near the crack initiation seemed to have port, and Northwestern Polytechnical University (NWPU), Xi’an,
facet-like features as shown in Fig. 16(a) and (b), and the dimen- China for providing test materials. One of the authors (D.L. Chen)
sions of the individual facets were also consistent with the primary is also grateful for the financial support by the Premier’s Research
a grain size as shown in Fig. 1(a). Similar results were also ob- Excellence Award (PREA), NSERC-Discovery Accelerator Supple-
served by Pilchak et al. [85] for a near a titanium alloy. ment (DAS) Award, Canada Foundation for Innovation (CFI), and
Ryerson Research Chair (RRC) program. The authors would also like
4. Conclusions to thank Messrs. Q. Li, A. Machin, J. Amankrah and R. Churaman for
easy access to the laboratory facilities of Ryerson University and
Tensile tests were conducted on dissimilar Ti–6Al–4V/Ti17 tita- their assistance in the experiments.
nium alloy joints made with EBW, and the effect of strain rate and
temperature on the UTS, YS, strain hardening behavior, strain rate
sensitivity, and fracture mechanisms of the joints were evaluated. References
The following conclusions can be drawn from this investigation:
[1] Huda Z, Edi P. Materials selection in design of structures and engines of
supersonic aircrafts: a review. Mater Des 2013;46:552–60.
1. The microstructure across the dissimilar welded joint [2] Zhu XJ, Tan MJ, Zhou W. Enhanced superplasticity in commercially pure
exhibited a considerable change, mainly consisting of HCP titanium alloy. Scripta Mater 2005;52:651–5.
martensite a0 and orthorhombic martensite a00 in the FZ, [3] Holmquist M, Recina V, Pettersson B. Tensile and creep properties of diffusion
bonded titanium alloy IMI 834 to gamma titanium aluminide IHI Alloy 01A.
a0 in the HAZ of Ti–6Al–4V side, and b in the HAZ of Ti17 Acta Mater 1999;47(6):1791–9.
side. The martensite in the FZ was finer than that in the [4] Zhang XH, Liu DX. Effects of temperature, slip amplitude, contact pressure on
HAZ of Ti–6Al–4V. fretting fatigue behavior of Ti811 alloys at elevated temperatures. Acta Metall
Sin (Engl Lett) 2009;22(2):131–7.
2. While the ductility reduced slightly, the UTS and YS of the [5] Li SJ, Murr LE, Cheng XY, Zhang ZB, Hao YL, Yang R, et al. Compression fatigue
joint lay in-between the two BMs of Ti–6Al–4V and Ti17, behavior of Ti–6Al–4V mesh arrays fabricated by electron beam melting. Acta
with Ti17 alloy being stronger. The YS, UTS, Voce stress of Mater 2012;60:793–802.
[6] Chen SH, Zhang MX, Huang JH, Cui CJ, Zhang H, Zhao XK. Microstructures and
the joint increased with increasing strain rate or decreasing mechanical property of laser butt welding of titanium alloy to stainless steel.
temperature. Mater Des 2014;53:504–11.
3. With increasing strain rate or decreasing temperature, both [7] Saresh N, Gopalakrishna Pillai M, Mathew J. Investigations into the effects of
electron beam welding on thick Ti–6Al–4V titanium alloy. J Mater Process
hardening capacity and strain hardening exponent
Technol 2007;192–193:83–8.
decreased. It was also observed that temperature has a [8] Barreda JL, SantamarmHa F, Azpiroz X, Irisarri AM, Varona JM. Electron beam
more significant effect on the hardening capacity than the welded high thickness Ti6Al4V plates using filler metal of similar and different
strain rate. composition to the base plate. Vacuum 2001;62:143–50.
[9] Gao XL, Zhang LJ, Liu J, Zhang JX. A comparative study of pulsed Nd:YAG laser
4. Only stage III hardening occurred after yielding in the joint. welding and TIG welding of thin Ti6Al4V titanium alloy plate. Mater Sci Eng A
The hardening rate was strongly dependent on the strain 2013;559:14–21.
rate and temperature. As the strain rate increased or tem- [10] Oh JK, Kim NJ, Lee SG. High-cycle fatigue properties of investment cast Ti–6Al–
4V alloy welds. J Mater Sci Lett 2001;20:2183–7.
perature decreased, the strain hardening rate increased at [11] Lu W, Li XY, Lei YP, Shi YW. Study on the mechanical heterogeneity of electron
a given true stress. beam welded thick TC4-DT joints. Mater Sci Eng A 2012;540:135–41.
S.Q. Wang et al. / Materials and Design 56 (2014) 174–184 183

[12] Wang SG, Wu XQ. Investigation on the microstructure and mechanical [43] Tong YX, Guo B, Zheng YF, Chung CY, Ma LW. Effects of Sn and Zr on the
properties of Ti–6Al–4V alloy joints with electron beam welding. Mater Des microstructure and mechanical properties of Ti–Ta-based shape memory
2012;36:663–70. alloys. J Mater Eng Perform 2011;20(4–5):762–6.
[13] Suresh D, Meshram T, Mohandas A. Comparative evaluation of friction and [44] Lu W, Shi YW, Lei YP, Li XY. Effect of electron beam welding on the
electron beam welds of near a titanium alloy. Mater Des 2010;31:2245–52. microstructures and mechanical properties of thick TC4-DT alloy. Mater Des
[14] Thomas G, Ramachandra V, Ganeshan R, Vasudevan R. Effect of pre- and post- 2012;34:509–15.
weld heat treatments on the mechanical properties of electron beam welded [45] Tan LJ, Yao ZK, Zhou W, Guo HZ, Zhao Y. Microstructure and properties of
Ti–6Al–4V alloy. J Mater Sci 1993;28:4892–9. electron beam welded joint of Ti–22Al–25Nb/TC11. Aerosp Sci Technol
[15] Pederson R, Niklasson F, Skystedt F, Warren R. Microstructure and mechanical 2010;14:302–6.
properties of friction- and electron-beam welded Ti–6Al–4V and Ti–6Al–2Sn– [46] Semiatin SL, Bieler TR. The effect of alpha platelet thickness on plastic flow
4Zr–6Mo. Mater Sci Eng A 2012;552:555–65. during hot working of Ti–6Al–4V with a transformed microstructure. Acta
[16] Fu PF, Liu FJ, Mao ZY, Li JW. Effects of electron beam local heat treatment on Mater 2001;49:3565–73.
fatigue properties for Ti–6Al–4V alloy joints. Int Technol Innovation Conf [47] Martin G, Nazé L, Cailletaud G. Numerical multi-scale simulations of the
2006:72–5. mechanical behavior of b-metastable titanium alloys Ti5553 and Ti17.
[17] Wang SQ, Liu JH, Chen DL. Tensile and fatigue properties of electron beam Procedia Eng 2011;10:1803–8.
welded dissimilar joints between Ti–6Al–4V and BT9 titanium alloys. Mater [48] Facchini L, Molinari A, Höges S, Wissenbach K. Ductility of a Ti–6Al–4V alloy
Sci Eng A 2013;584:47–56. produced by selective laser melting of prealloyed powders. Rapid Prototyp J
[18] Wang SQ, Liu JH, Chen DL. Strain-controlled fatigue properties of dissimilar 2010;16:450–9.
welded joints between Ti–6Al–4V and Ti-17 alloys. Mater Des [49] Nalla RK, Boyce BL, Campbell JP, Peters JO, Ritchie RO. Influence of
2013;49:716–27. microstructure on high-cycle fatigue of Ti–6Al–4V: bimodal vs. lamellar
[19] Farabi N, Chen DL, Zhou Y. Tensile properties and work hardening behavior of structures. Metall Mater Trans A 2002;33A:899–918.
laser-welded dual-phase steel joints. J Mater Eng Perform 2012;21:222–30. [50] Ma TJ, Zhong B, Li WY, Yang SQ, Yang CL. On microstructure and mechanical
[20] Chowdhury SM, Chen DL, Bhole SD, Powidajko E, Weckman DC, Zhou Y. properties of linear friction welded dissimilar Ti–6Al–4V and Ti–6.5Al–3.5Mo–
Microstructure and mechanical properties of fiber and diode laser welded 1.5Zr 0.3Si joint. Sci Technol Weld Join 2012;17:9–12.
AZ31 magnesium alloy. Metall Mater Trans A 2011;42(7):1974–89. [51] Shao HP, Gould J, Albright C. Laser blank welding high-strength steels. Metall
[21] Chowdhury SH, Chen DL, Bhole SD, Powidajko E, Weckman DC, Zhou Y. Fiber Mater Trans B 2007;38B:321–31.
laser welded AZ31 magnesium alloy: effect of welding speed on [52] Venkatesh BD, Chen DL, Bhole SD. Effect of heat treatment on mechanical
microstructure and mechanical properties. Metall Mater Trans A properties of Ti–6Al–4V ELI alloy. Mater Sci Eng A 2009;506:117–24.
2012;43(6):2133–47. [53] Ward CH, Thompson AW, Williams JC. Elevated temperature tensile behavior
[22] Xu WF, Liu JH, Chen DL, Luan GH, Yao JS. Improvements of strength and of Ti–25Al–10Nb–3V–1Mo. Metall Mater Trans A 1995;26A:703–20.
ductility in aluminum alloy joints via rapid cooling during friction stir welding. [54] Salem AA, Kalidindi SR, Doherty RD. Strain hardening of titanium: role of
Mater Sci Eng A 2012;548:89–98. deformation twinning. Acta Mater 2003;51:4225–37.
[23] Afrin N, Chen DL, Cao X, Jahazi M. Strain hardening behavior of a friction stir [55] López JG, Peirs J, Verleysen P, Degrieck J. Effect of small temperature variations
welded magnesium alloy. Scripta Mater 2007;57:1004–7. on the tensile behaviour of Ti–6Al–4V. Procedia Eng 2011;10:2330–5.
[24] Chowdhury SM, Chen DL, Bhole SD, Cao X, Powidajko E, Weckman DC, et al. [56] Zhou W, Chew KG. The rate dependent response of a titanium alloy subjected
Tensile properties and strain-hardening behavior of double-sided arc welded to quasi-static loading in ambient environment. J Mater Sci 2002;37:5159–65.
and friction stir welded AZ31B magnesium alloy. Mater Sci Eng A [57] Jaworski AW, Ankem S. The effect of a phase on the deformation mechanisms
2010;527:2951–61. of b titanium alloys. J Mater Eng Perform 2005;14(6):755–60.
[25] Chowdhury SH, Chen DL, Bhole SD, Cao X, Wanjara P. Friction stir welded AZ31 [58] Song H, Zhang S, Cheng M. Subtransus deformation mechanisms of TC11
magnesium alloy: Microstructure, texture, and tensile properties. Metall titanium alloy with lamellar structure. Trans Nonferr Met Soc China
Mater Trans A 2013;44(1):323–36. 2010;20:2168–73.
[26] Chowdhury SM, Chen DL, Bhole SD, Cao X. Tensile properties of a friction stir [59] Boehlert CJ, Chen Z, Chakkedath A, Gutiérrez-Urrutia I, Llorca J, Bohlen J, et al.
welded magnesium alloy: effect of pin tool thread orientation and weld pitch. In situ analysis of the tensile deformation mechanisms in extruded Mg–1Mn–
Mater Sci Eng A 2010;527(21–22):6064–75. 1Nd (wt%). Philos Mag 2013;93(6):598–617.
[27] Afrin N, Chen DL, Cao X, Jahazi M. Microstructure and tensile properties of [60] Wen H, Wang Y, Li ZR, Xia YM. Influences of temperature and strain rate on
friction stir welded AZ31B magnesium alloy. Mater Sci Eng A 2008;472(1– deformation twinning of polycrystalline titanium. Chin J Nonferr Met
2):179–86. 2008;18:1440–5.
[28] Chen XH, Lu L. Work hardening of ultrafine-grained copper with nanoscale [61] Chiou ST, Tsai HL, Lee WS. Effects of strain rate and temperature on the
twins. Scripta Mater 2007;57:133–6. deformation and fracture behaviour of titanium alloy. Mater Trans
[29] Meyers MA, Chawla KK. Mechanical behavior of materials. Upper Saddle River 2007;48(9):2525–33.
(NJ): Prentice Hall, Inc.; 1999. [62] Vanderhasten M, Rabet L, Verlinden B. Ti–6Al–4V: deformation map and
[30] Chu D, Morris JW. The influence of microstructure on work hardening in modelisation of tensile behaviour. Mater Des 2008;29:1090–8.
aluminum. Acta Mater 1996;44:2599–610. [63] Ul’yanov RA, Moskalenko VA. Characteristics of plastic deformation of
[31] Kocks UF, Mecking H. Physics and phenomenology of strain hardening: the FCC titanium at low temperatures. Metallovedenie i Termicheskaya Obrabotka
case. Prog Mater Sci 2003;48:171–273. Metallov 1966;10:48–51.
[32] Kocks UF. The relation between polycrystal deformation and single-crystal [64] Hosford WF. Mechanical behavior of materials. New York: Cambridge
deformation. Metall Trans A 1970;1:1121–33. University Press; 2005.
[33] Patel HA, Chen DL, Bhole SD, Sadayappan K. Microstructure and tensile [65] El-Domiaty A. Effect of strain, strain rate and temperature on formability of Ti–
properties of thixomolded magnesium alloys. J Alloys Comp 2010;496(1– 6Al–4V alloy. J Mater Process Technol 1992;32:243–51.
2):140–8. [66] Isaac SE, Choudhary BK. Universal scaling of work hardening parameters in
[34] Fan CL, Chen DL, Luo AA. Dependence of the distribution of deformation twins type 316L(N) stainless steel. Mater Sci Eng A 2010;527:7457–60.
on strain amplitudes in an extruded magnesium alloy after cyclic deformation. [67] Zhao HL, Qiu F, Jin SB, Jiang QC. Effect of different strain rates on compression
Mater Sci Eng A 2009;519(1–2):38–45. property and work-hardening behavior for the NiAl-matrix composite with
[35] Luo J, Li MQ, Yu WX, Li H. The variation of strain rate sensitivity exponent and 1.7 wt.% NbB2 and NbxC. Mater Sci Eng A 2012; 534:22–5.
strain hardening exponent in isothermal compression of Ti–6Al–4V alloy. [68] Liu TT, Pan FS, Zhang XY. Effect of Sc addition on the work-hardening behavior
Mater Des 2010;31(2):741–8. of ZK60 magnesium alloy. Mater Des 2013;43:572–7.
[36] Bystrzanowski S, Bartels A, Clemens H, Gerling R. Characteristics of the tensile [69] Chen XH, Pan FS, Mao JJ, Wang JF, Zhang DF, Tang AT, et al. Effect of heat
flow behavior of Ti–46Al–9Nb sheet material-analysis of thermally activated treatment on strain hardening of ZK60 Mg alloy. Mater Des 2011;32:1526–30.
processes of plastic deformation. Intermetallics 2008;l16:717–26. [70] Zhao HL, Qiu F, Jin SB, Jiang QC. Compression properties and work-hardening
[37] Honarmandi P, Aghaie-Khafri M. Hot deformation behavior of Ti–6Al–4V alloy effect of the NiAl-matrix composite with TaB2 and TaB. Intermetallics
in b phase field and low strain rate. Metall Microstruct Anal 2013;2:13–20. 2012;27:1–5.
[38] Rao PP, Tangri K. Yielding and work hardening behaviour of titanium [71] Xuan QQ, Shu SL, Qiu F, Jin SB, Jiang QC. Different strain-rate dependent
aluminides at different temperatures. Mater Sci Eng A 1991;132:49–59. compressive properties and work-hardening capacities of 50 vol% TiCx/Al and
[39] Akhonin SV, Mishchenko RN, Petrichenko IK. Investigation of the weldability of TiB2/Al composites. Mater Sci Eng A 2012;538:335–9.
titanium alloys produced by different methods of melting. Mater Sci [72] Choudhary BK, Palaparti DPR. Comparative tensile flow and work hardening
2006;42:323–9. behaviour of thin section and forged thick section 9Cr–1Mo ferritic steel in the
[40] Sabol JC, Pasang T, Misiolek WZ, Williams JC. Localized tensile strain framework of Voce equation and Kocks–Mecking approach. J Nucl Mater
distribution and metallurgy of electron beam welded Ti–5Al–5V–5Mo–3Cr 2012;430:72–81.
titanium alloys. J Mater Process Technol 2012;212:2380–5. [73] Morán-López JL, Mejía-Lira F, Sanchez JM, editors. Structural and phase
[41] Pasang T, Sánchez Amaya JM, Tao Y, Amaya-Vazquez MR, Botana FJ, Sabol JC, stability of alloys. New York (NY): Plenum Press; 1992. p. 65–86.
Misiolek WZ, Kamiya O. Comparison of Ti–5Al–5V–5Mo–3Cr welds performed [74] Akhtar A, Teghtsoonian E. Prismatic slip in a-titanium single crystals. Metall
by laser beam, electron beam and gas tungsten arc welding. Procedia Eng Trans A 1975;6A:2201–8.
2013;63:397–404. [75] Ravi KNV, Blandin JJ, Desrayaud C, Montheillet F, Suéry M. Grain refinement in
[42] Eisenbartha E, Veltena D, Müllera M, Thullb R, Bremea J. Biocompatibility of b- AZ91 magnesium alloy during thermomechanical processing. Mater Sci Eng A
stabilizing elements of titanium alloys. Biomater 2004;25(26):5705–13. 2003;359:150–7.
184 S.Q. Wang et al. / Materials and Design 56 (2014) 174–184

[76] Kuhlmann-Wilsdorf D. Theory of Work Hardening 1934–1984. Metall Trans A [82] Jain M, Chaturvedi MC, Richards NL, Goel NC. Strain rate sensitivity effects
1985;16A:2091–108. with forming characteristics of superplastic Ti–6Al–4V. Mater Sci Eng A
[77] Shibata N, Goto A, Choi S-Y, Mizoguchi T, Findlay SD, Yamamoto T, et al. Direct 1991;138:205–11.
imaging of reconstructed atoms on TiO2 (1 1 0) surfaces. Science 2008;322:570–3. [83] Makoto I, Michinori O, Makoto S, Junichi K, Jiro M, Yoshio N. Creep and creep
[78] Fujii H, Suzuki HG. Effect of cooling rates on microstructures of b treated a+b rupture of welded joints of 5083 aluminum alloy plates. J Jpn Inst Light Metals
titanium alloys. Tetsu-To-Hagane/J Iron Steel Inst Jpn 1991;77:1481–8. 1998;48:479–83.
[79] Lin XZ, Chen DL. Strain hardening and strain-rate sensitivity of an extruded [84] Makoto I, Michinori O, Makoto S, Junichi K, Jiro M, Yoshio N. Creep and creep
magnesium alloy. J Mater Eng Perform 2008;17:894–901. rupture of welded joints of 5083 alloy plates. Joints in Aluminium-Inalco: 7th
[80] Shen YF, Lu L, Dao M, Suresh S. Strain rate sensitivity of Cu with nanoscale International Conference; 1998. p. 311–19.
twins. Scripta Mater 2006;55:319–22. [85] Pilchak AL, Porter WJ, John R. Room temperature fracture processes of a near-a
[81] Lindholm US. Some experiments with split Hopkinson pressure bar. J Mech titanium alloy following elevated temperature exposure. J Mater Sci
Phys Solids 1964;12:317–35. 2012;47:7235–53.

You might also like