You are on page 1of 9

Journal of Materials Science & Technology 73 (2021) 52–60

Contents lists available at ScienceDirect

Journal of Materials Science & Technology


journal homepage: www.jmst.org

Research Article

Effect of strain rate and temperature on the deformation behavior in a


Ti-23.1Nb-2.0Zr-1.0O titanium alloy
Yi Yang a , Di Xu a , Sheng Cao b,∗ , Songquan Wu a,∗ , Zhengwang Zhu c , Hao Wang c , Lei Li d ,
Shewei Xin d , Lei Qu d , Aijun Huang e
a
School of Materials Science and Engineering, University of Shanghai for Science and Technology, Shanghai 200093, China
b
Department of Materials, University of Manchester, Oxford Road, Manchester M13 9PL, UK
c
Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China
d
Northwest Institute for Non-ferrous Metal Research, Xi’an 710016, China
e
Department of Materials Science and Engineering, Monash University, Clayton, VIC 3800, Australia

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

Article history: The compression behavior of a Ti-23.1Nb-2.0Zr-1.0O (at.%) alloy was investigated at strain rates from 0.1
Received 6 June 2020 s−1 to 1000 s−1 and temperatures from 100 ◦ C to 200 ◦ C on a Gleeble 3800 system and Split Hopkinson
Received in revised form 6 July 2020 Pressure Bar (SHPB) compressive tester. Optical microscopy, electron backscatter diffraction (EBSD), X-
Accepted 7 July 2020
ray diffraction (XRD) and transmission electron microscopy (TEM) were employed to characterize the
Available online 2 October 2020
microstructure evolution during the deformation. Numerous deformation phenomena, including dislo-
cation slip, twinning of both {332}113 and {112}111 modes, stress-induced ␣” martensite (SIM␣”)
Keywords:
and stress-induced ␻ (SI␻) transformations, were observed. The preferred activation of twinning and
Titanium alloy
Plastic deformation
SI␻ transformations was observed in the sample compressed at lower temperatures and/or higher strain
Strain rate rates. The underlying mechanism is that twinning and stress induced phase transformations are attribute
Temperature to higher stress concentrations at ␤ grain boundaries and additional energy supplied by a higher strain
Microstructure rate, as well as high stacking fault energy because of higher temperature.
© 2021 Published by Elsevier Ltd on behalf of The editorial office of Journal of Materials Science &
Technology.

1. Introduction strain rates from 0.001 s−1 to 1 s−1 at 750−850 ◦ C, and they found
that the flow stress significantly increases with the decreasing of
Metastable ␤ titanium alloys have been applied in a vari- deformation temperature and increasing of strain rate, which was
ety of applications including aerospace industries and biomedical explained by dislocation kinetics. Once the strain rate increases, the
devices due to their high strength, low Young’s modulus and velocity of mobile dislocations is required to increase, which causes
superior biocompatibility [1–6]. The deformation mechanism of higher applied stress. Moreover, some studies have proposed that
metastable ␤ alloys is considerably more complex than the most high-speed compression was responsible for the higher number
widely employed ␣ + ␤ titanium alloys. In general, the deformation density of twins, and it would increase the flow stress according to
mechanism can vary from stress-induced phase transformations to the dynamic Hall-Petch law [13–17]. Sadeghpour et al. [18] have
twining and then to dislocation sliding when the stability of ␤ phase investigated the compressed Ti-4Al-7Mo-3V-3Cr (wt.%) alloy at a
is increasing [7–9]. In our previous work [10,11] we found that the strain rate range from 0.7 × 10-4 to 0.7 × 10−1 s−1 , and they have
dominant deformation mechanism is also related to the extent of observed ␣” martensite in all compressed conditions. Ahmed et al.
plastic deformation. [19] have tested a metastable ␤ Ti-10V-3Fe-3Al-0.27O (wt.%) alloy
In recent years, the effect of loading parameters, i.e. strain rate at strain rates of 10-3 s−1 , 10−1 s−1 , 10 s−1 , and 102 s−1 at ambi-
and temperature, on the deformation mechanisms of the ␤ tita- ent temperature, and they found that {332}113 twinning was the
nium alloys have been investigated. Bai et al. [12] have studied dominant deformation mode when the strain rate was higher than
the hot compression behavior of Ti-3Zr-2Sn-3Mo-25Nb (wt.%) at 10 s−1 .
Most previous studies are mainly focused on quasi-static defor-
mation at a low strain rate. Because of the potential applications in
∗ Corresponding authors.
high strain rate conditions, like armor plates on military vehicles,
E-mail addresses: sheng.cao@yahoo.com (S. Cao), sqwu@alum.imr.ac.cn (S. Wu).
the dynamic deformation at a high strain rate of titanium alloys is

https://doi.org/10.1016/j.jmst.2020.09.030
1005-0302/© 2021 Published by Elsevier Ltd on behalf of The editorial office of Journal of Materials Science & Technology.
Y. Yang et al. Journal of Materials Science & Technology 73 (2021) 52–60

important for ␤ titanium alloys. A few studies have presented the deformation. However, compared to the abundant studies on quasi-
influence of loading temperatures and strain rates on the defor- static deformation at low strain rates, knowledge regarding the
mation behavior. Zhan et al. [14] have investigated the dynamic dynamic deformation at high strain rates of metastable ␤ titanium
response of Ti-25Nb-3Zr-3Mo-2Sn (wt.%) at a strain rate of 1000 s−1 alloys is still insufficient.
and a temperature range from 293 K to 873 K. They have revealed In the present work, we have studied a metastable ␤ titanium
that the loading temperature is an important factor in determining Ti-23.1Nb-2.0Zr-1.0O (at.%) alloy modified from Gum Metal [21,22].
the deformation mode. Xiao et al. [20] deformed a Ti-2Al-9.2Mo- The alloy has been compressed at a strain rate range of 0.1-1000 s−1
2Fe (wt.%) alloy at a strain rate of 3000 s−1 at room temperature, and at 100−200 ◦ C. In order to reveal the deformation mechanism, the
they have observed stress-induced ␣” martensite (SIM␣”), stress- effects of strain rate and temperature on the deformation behavior
induced ␻ phase (SI␻), {332}113 twins and dislocations after

Fig. 1. True strain-stress curves of specimens deformed at strain rate of (a) 0.1 s−1 , (b) 10 s−1 and (c) 1000 s−1 with temperatures varying from 100 ◦ C to 200 ◦ C; (d) The yield
strength vs. strain rate; (e) The yield strength vs. deformation temperature.

53
Y. Yang et al. Journal of Materials Science & Technology 73 (2021) 52–60

have been investigated by carrying out multiscale characteriza-


tions.

2. Material and methods

A 20 kg ingot was triple melted in a consumable arc-melting


furnace under Ar atmosphere. The measured chemical composition
was Ti-23.1Nb-2.0Zr-1.0O (at.%) with an electron/atom ratio (e/a) at
4.231, a bond order (Bo) of 2.867, a d electron-orbital energy level
(Md) at 2.451 and a Molybdenum equivalence ([Mo]e ) of 10.13,
which are close to those of Gum Metal [21,22]. The ␤ transus is cal-
culated at 564 ◦ C according to Sun’s work [23]. The ingot was forged
at 950 ◦ C, then rolled into a  10 mm bar at 800 ◦ C and drew into a
 5 mm bar at 600 ◦ C. The bar was subsequently solution-treated at
890 ◦ C for 30 min, followed by water quenching.  4.7 mm cylindri-
cal specimens with a length of 4.7 mm were compressed uniaxially
at strain rates of 0.1 s−1 , 10 s−1 and 1000 s−1 at 100 ◦ C, 150 ◦ C
and 200 ◦ C. The 0.1 s−1 and 10 s−1 compressive tests were con-
ducted on Gleeble 3800 and the 1000 s−1 compressive tests were
conducted by Split Hopkinson Pressure Bar (SHPB). Phase identifi-
cation was carried out by a Bruker D8 Advance diffractometer (XRD)
using a Cu K˛ source at 40 kV and 300 mA. Microstructural obser-
vations were carried out on a LEICA DMI8 optical microscope (OM).
Electron back-scattered diffraction (EBSD) data were acquired on
a Quanta 450 FEG scanning electron microscope (SEM) equipped
with an EBSD detector, and the EBSD data was then analyzed by
an OIM software to obtain the inverse pole figure (IPF). The EBSD
map was recorded at a step size of 0.35 ␮m. Transmission elec-
tron microscopy (TEM) analysis was performed on an FEI Tecnai
F30 transmission electron microscope operated at 300 kV. The TEM
foils were perpendicular to the compressed direction.

3. Results

3.1. Compression behavior

The true stress-strain curves of different compressed conditions


are shown in Fig. 1(a)-(c). For the strain rate of 0.1 s−1 (Fig. 1(a)) at Fig. 2. XRD profiles of specimens deformed (a) at 100 ◦ C with different strain rates
100 ◦ C, 150 ◦ C and 200 ◦ C, the yield strengths were 642 ± 8 MPa, and (b) at a strain rate of 1000 s−1 with different temperatures. The original condition
597 ± 4.5 MPa and 546 ± 8 MPa, respectively. For a higher strain is the sample before deformation as a comparison.
rate 10 s−1 (Fig. 1(b)) tested at the sample temperatures, the yield
strengths were increased to 864 ± 5 MPa, 787 ± 25 MPa and 701 The absence of martensite in the solution and water quenched sam-
± 2 MPa, respectively. A further increase in strain rate to 1000 s−1 ple (the original specimen in Fig. 2) and the presence of martensite
(Fig. 1(c)) led to enhanced yield strengths of 991 ± 15 MPa, 885 ± in the samples compressed at 100−200 ◦ C indicate that the start-
2.5 MPa and 769 ± 21 MPa for sample compressed at 100 ◦ C, 150 ing temperature of ␣” martensitic transformation (Ms ) has been
◦ C and 200 ◦ C, respectively.
elevated by the deformation.
The yield strength as functions of strain rate and temperature According to our previous work, the solution-treated specimen
are plotted in Fig. 1(d) and (e), respectively. Linear relationships generally shows a twin-free equiaxed ␤ grain microstructure with
were achieved in both cases. The difference was that yield strength a grain size at 20−100 ␮m [24]. In this study, optical images of
increased with strain rate, but it decreased with temperature. After the compressed specimens are shown in Fig. 3. After compression,
yielding, the alloy exhibited work hardening at the strain rate of there was no significant change in both the size and morphology of
0.1 s−1 , while it showed work softening for the higher strain rates the prior ␤ grains, while some coarse lath-like features appeared
at 10 s−1 and 1000 s−1 . Such a distinction could be the result of the in ␤ grains for the samples compressed at a relatively lower tem-
isothermal effect at a lower strain rate and the adiabatic effect at a perature and a higher strain rate, as indicated by the white arrows
higher strain rate. in Fig. 3. In ␤ titanium alloys, such features are generally defor-
mation twins. In order to further confirm them, EBSD was used
3.2. Microstructure characterization to characterize the coarse lath-like features. Taking the specimen
compressed at a strain rate of 1000 s−1 and 100 ◦ C as an example,
XRD profiles of the specimens deformed at 100 ◦ C with different the EBSD IPF map is shown in Fig. 4(a). The crystallographic orien-
strain rates and at a strain rate of 1000 s−1 with different temper- tations of the lath-like features were completely different from the
atures are shown in Fig. 2. ␤-Ti was the only phase detected in matrix ␤ grains as shown in the orientation map. Fig. 4(b) shows
the solution-treated specimen before deformation highlighted in the grain boundary and twin boundary map, and it suggests that
green. While all the compressed specimens contain ␣” martensite these lath-like features are {332}113 mechanical twins.
in addition to the ␤ phase. The formation of ␣” has been proposed For quantitative analysis, the number density of {332}113
to be related to the stress-induced martensitic transformation [13]. twins was calculated from over 2000 grains in a 5 mm2 region for

54
Y. Yang et al. Journal of Materials Science & Technology 73 (2021) 52–60

Fig. 3. Optical micrographs of specimens compressed at (a) 100 ◦ C, 0.1 s−1 ; (b)100 ◦ C, 10 s−1 ; (c) 100 ◦ C, 1000 s−1 ; (d) 150 ◦ C, 0.1 s−1 ; (e) 150 ◦ C, 10 s−1 ; (f) 150 ◦ C, 1000 s−1 ;
(g) 200 ◦ C, 0.1 s−1 ; (h) 200 ◦ C, 10 s−1 ; and (i) 200 ◦ C, 1000 s−1 . White arrows denote lath-like features.

Fig. 4. (a) EBSD IPF map and (b) grain boundaries and twin boundaries map of the specimen compressed at a strain rate of 1000 s−1 at 100 ◦ C. In (b), the blue and red lines
highlights the grain boundaries and {332}<113> twin boundaries respectively.

all the deformed conditions. As shown in Fig. 5, the twin number In order to reveal the deformation mechanism, TEM was
density shows a remarkable change with the strain rate and the employed to characterize the deformed microstructure for those
temperature. In the specimens compressed at a strain rate lower samples showed significant amount of twins. Fig. 6 shows the TEM
than 10 s−1 and a temperature higher than 150 ◦ C, {332}113 twin bright field (BF) images of tangled dislocations in the specimens
is not observed. At 100 ◦ C, the twin number densities for strain rates compressed at 100 ◦ C with a strain rates of 10 s−1 , and 100−200 ◦ C
of 0.01 s−1 , 10 s−1 and 1000 s−1 are 9 mm-2 , 108 mm-2 and 211 with a strain rate of 1000 s−1 . Dislocation networks were observed
mm-2 , respectively. When the strain rate is fixed at 1000 s−1 , the in all deformed specimens. This reveals that dislocation slip is one
twin number densities for 200 ◦ C, 150 ◦ C and 100 ◦ C are 0.6 mm-2 , of the dominant deformation mechanisms in these samples.
126 mm-2 and 211 mm-2 , respectively. It indicates that {332}113 In addition to dislocations, another type of {112}111 mechan-
twinning are preferred to occur at conditions compressed at low ical twin and SI␻ phase were also observed in TEM. As {112}111
temperatures and high strain rates. twins were much finer than the {332}113 twins, {112}111 was

55
Y. Yang et al. Journal of Materials Science & Technology 73 (2021) 52–60

{112}111 mechanical twins and SI␻ are much smaller than that
of {332}113 mechanical twins.
When the temperature increased to 150 ◦ C, the concomitant
{112}111 twins and SI␻ were observed. Fig. 9(a) shows the DF
image of zigzag-shaped {112}111 mechanical twins in the spec-
imen compressed at a strain rate of 1000 s−1 . In addition, a thin
layer of SI␻ is located on the twin boundary according to DF image
in Fig. 9(b). For the sample tested at the highest temperature of 200
◦ C with a strain rate of 1000 s−1 , a V-shaped {112}111 mechan-

ical twins containing SI␻ precipitates were observed. The TEM DF


result shows that the SI␻ precipitates formed within those defor-
mation twins is a representative case in those sample compressed
at moderate to high strain rates.

4. Discussion

Because of the small observed area in TEM, it is difficult


to achieve a quantitative analysis of the number density for
{112}111 mechanical twins and SI␻ for different deformation
conditions, while the number of {332}113 twins can be easily
Fig. 5. The number density of {332}113 twins in different deformed conditions. counted on OM, SEM and EBSD images as they have a substantially
larger size. Therefore, only the quantitative relationship between
the number density of {332}113 twins and deformation param-
eters is established in this study. The {332}113 twin number
not detectable in the EBSD map in Fig. 4. TEM dark-field (DF) images density increases with the strain rate in the range from 0.1 s−1 to
of these mechanical twins and SI␻ are shown in Figs. 7–10. Fig. 7(a) 1000 s−1 , but it decreases with the temperature in the range from
shows the DF image of plate-like {112}111 mechanical twins 100 ◦ C to 200 ◦ C. {332}113 twinning becomes more significant at
accompanied by SI␻ in the specimen compressed at 100 ◦ C with higher strain rate and lower temperature conditions. Meanwhile,
a strain rate of 10 s−1 . Fig. 7(b) shows the DF image of the same the compressive yield strength has a similar trend as shown in
plate-like SI␻ by using another reflection. Fig. 1(d) and (e).
For the sample compressed at the same temperature of 100 ◦ C In general, a high stress concentration is required for twin nucle-
but a higher strain rate of 1000 s−1 , individual late-like {112}111 ation. It has been well-known that grain boundaries usually act as
mechanical twins and plate-like SI␻ were observed as shown in obstacles for dislocation slip, and the dislocations pile-up in the
Fig. 8(a) and (c), respectively. In addition, the plate-like SI␻ was grain boundary front results in a high stress concentration during
observed within the {332}113 mechanical twins according to straining. Therefore twins normally nucleate at grain boundaries,
Fig. 8(e) and (f). This indicates that {332}113 twinning happens where the dislocation density and the stress concentration are the
first, and SI␻ is formed afterwards. The thicknesses of the observed highest. Ahmed et al. [16] have proposed that the number of twin

Fig. 6. TEM bright field (BF) images of deformed structures in the specimens compressed at strain rate of (a) 10 s−1 at 100 ◦ C, (b) 1000 s−1 at 100 ◦ C, (c) 1000 s−1 at 150 ◦ C
and (d) 1000 s−1 at 200 ◦ C.

56
Y. Yang et al. Journal of Materials Science & Technology 73 (2021) 52–60

Fig. 7. TEM dark-field (DF) images of deformation microstructure in the specimen compressed at 100 ◦ C with a strain rate of 10 s−1 . (a) DF image using the circled reflection
((011)␤t //(1011)␻) in (c); (b) DF image using the squared reflection ((1100)␻) in (c); (c) the selected area diffraction pattern along [311]␤m //[311]␤t //[1123]␻. m: matrix,
t: twin.

nucleation site is dependent on the dislocation density, and the ary front, which then leads to an increase in the number
relation can be established by the following equation [16]: density of twins, including both {332}113 and {112}111
2 twins.
nnuc = (2x¯T t/l )nf (1) In Fig. 5, another experiment observation was twin number den-
where nnuc is the number of twins, x¯T is the length of twin embryo, t sity reduced with temperature. In plastic deformation, twinning
is the thickness of twin embryo, l is the average dislocation spacing, and dislocation slip are two dominant deformation mechanisms
2 for metals and alloys [28–32], and ability to form twins is pro-
nf is the number of dislocations. Therefore, 2x¯T t/l is related to the
posed to be controlled by the stacking fault energy (SFE). It has
area of the twin embryo.
been reported that twinning preferred to occur in ␤ Ti alloys with
In addition, it has been proposed that a high strain rate leads to
moderate SFE, while dislocation slip preferred to occur when the
a high dislocation density [25–27], which can be expressed as [26]:
SFE is higher. Xing et al. [33] have calculated the SFE of a Ti-Nb

ε˙p = mbm v (2) alloy by using first principles and have found that the SFE of the ␤
phase increases with the e/a ratio (4.24–5.0). In other words, the

where ε˙p is plastic strain rate, m can be interpreted as the Schmid SFE of ␤ phase increases with the Nb content. Experimental obser-
factor, b is the magnitude of the Burgers vector, m is the mobile vations have found that twining is more likely to happen in Ti-Nb
dislocation density, and v is the velocity of mobile dislocations. with a lower Nb content [34–37], which also suggest that twin-
Therefore, it is reasonable to suggest that an enhanced ning happens in a relatively lower SFE regime. Moreover, Tadmor
strain rate increases the number density of dislocations et al. [38] have proposed a model to reveal the competitive rela-
and the associated stress concentration at the grain bound- tionship between dislocation slip and twinning. This model shows

57
Y. Yang et al. Journal of Materials Science & Technology 73 (2021) 52–60

Fig. 8. TEM dark-field (DF) images of deformation microstructure in the specimen compressed at 100 ◦ C with a strain rate of 1000 s−1 . (a) DF image of {112}111 twins using
the circled reflection ((011)␤t ) in (b); (b) the selected area diffraction pattern along [311]␤m //[311]␤t ; (c) DF image using the squared reflection ((1100)␻) in (d); (d) the
selected area diffraction pattern along [110]␤m //[1120]␻; (e) DF image using the circled reflection ((110)␤t ) in (g); (f) DF image using the squared diffraction spot ((1100)␻)
in (g); (g) the selected area diffraction pattern along [110]␤m //[110]␤t //[1120]␻. m: matrix, t: twin.

that the transition between slip and twinning is affected by the alloys [40], us decreases with temperature, but ut increases with
unstable-SFE (us , the energy required to form a partial dislocation) temperature. This suggests that dislocation slip is preferred at high
and the unstable twinning-energy (ut , the energy required to form temperatures, while twinning as a deformation mode is favorable
a twin). In previous studies on HCP metals including Mg [39] and Ti at low temperatures. The change of {332}113 twin number den-

58
Y. Yang et al. Journal of Materials Science & Technology 73 (2021) 52–60

Fig. 9. TEM dark-field (DF) images of deformation microstructure in the spec- Fig. 10. TEM dark-field (DF) images of deformation microstructure in the spec-
imen compressed at 150 ◦ C with a strain rate of 1000 s−1 . (a) DF image imen compressed at 200 ◦ C with a strain rate of 1000 s−1 . (a) DF image
using the circled reflection ((002)␤t //(2021)␻) in (c); (b) DF image using the using the circled reflection ((110)␤t //(1101)␻) in (c); (b) DF image using the
squared reflection ((1100)␻) in (c); (c) the selected area diffraction pattern along squared reflection ((1100)␻) in (c); (c) the selected area diffraction pattern along
[210]␤m //[210]␤t //[1126]␻. m: matrix, t: twin. [110]␤m //[110]␤t //[1120]␻. m: matrix, t: twin.

sity with the temperature in the present work agrees well with this
theory. perature, and {112}111 twinning occurs during tensile test when
SI␻ transformation and {112}111 twinning have been the temperature reduces to liquid nitrogen temperature [44]. It is
reported to have a similar shear mechanism. A 1/2[111] perfect dis- worth to mention that only a single variant of ␻ is observed after
location dissociates into three 1/6[111] partials (one plane apart), compression at low temperatures and high strain rates, and this is
related to the athermal ␻ reversion during the deformation [45,46],
and a correlative glide among them on (112) plane can form a
which facilitates the formation of {112}111 twins [47]. There-
three-layer (112)[111] microtwin. Similarly, a 1/2[111] perfect dis- fore, an observation of concomitant formed {112}111 twins and
location dissociates into two 1/12[111] partials and one 1/3[111] SI␻ precipitates was found in TEM results. In the present work, as
partial (one plane apart), and a correlative glide among them on strain rate increasing, high stress concentration at the grain bound-
(112) plane can form a three-layer micro-␻ [34]. Therefore, the ary leads to an increase in the number density of {112}111 twins
formation of {112}111 twins are usually accompanied by SI␻ pre- and a concomitant promotion in SI␻ precipitation. As temperature
cipitation [41]. Both {112}111 twinning and SI␻ transformation increasing, perfect dislocation is hard to dissociate because of the
have also been found preferred to occur at high strain rates and increasingly high SFE. The chance for {112}111 twins and and SI␻
low temperatures. It has been reported that {112}111 twins and to occur reduces.
SI␻ could form via increasing strain rate or reducing temperature In the specimens compressed at the highest strain rate of 1000
in bcc Ta alloys. For instance, {112}111 twins [42] and SI␻ [43] s−1 , the morphology of {112}<111> twins changes from straight to
have been observed after shock-wave deformation at room tem- zigzag-shaped and V-shaped with an increased deformation tem-

59
Y. Yang et al. Journal of Materials Science & Technology 73 (2021) 52–60

perature according to Figs. 8–10. In one of our previous work [36], [5] W.S. Lee, T.H. Chen, H.H. Hwang, Metall. Mater. Trans. A 39 (2008) 1435–1448.
a model based on Lagerlöf’s twinning theory and cross-slip theory [6] A. Kreitcberg, V. Brailovski, S. Prokoshkin, J. Mater. Process. Technol. 252
(2018) 821–829.
of screw dislocations has been proposed for the formation mecha- [7] J. Hwang, S. Kuramoto, T. Furuta, K. Nishino, T. Saito, J. Mater. Eng. Perform. 14
nism of zigzag-shaped and V-shaped {112}<111> twins. This model (2005) 747–754.
involves the dissociation of 1/2[111] dislocation to three 1/6[111] [8] M. Abdel-Hady, K. Hinoshita, M. Morinaga, Scr. Mater. 55 (2006) 477–480.
[9] R.J. Talling, R.J. Dashwood, M. Jackson, S. Kuramoto, D. Dye, Scr. Mater. 59
partials, and the following cross-slip on adjacent {112} planes as (2008) 669–672.
discussed in a previous paragraph. The occurrence of cross-slip [10] Y. Yang, S.Q. Wu, G.P. Li, Y.L. Li, Y.F. Lu, K. Yang, P. Ge, Acta Mater. 58 (2010)
requires a constriction of the extended dislocation into a new 1/2 2778–2787.
[11] Y. Yang, G.P. Li, G.M. Cheng, H. Wang, M. Zhang, F. Xu, K. Yang, Scr. Mater. 58
< 111> screw dislocation. Both the dislocation dissociation and
(2008) 9–12.
the subsequent cross-slip requires a moderate SFE. Therefore in [12] X.F. Bai, Y.Q. Zhao, W.D. Zeng, Z.Q. Jia, Y.S. Zhang, Mater. Sci. Eng. A 598 (2014)
the present work, the change in morphology of {112}<111> twins 236–243.
[13] H.Y. Zhan, W.D. Zeng, G. Wang, D. Kent, M. Dargusch, Scr. Mater. 107 (2015)
from straight into zigzag-shaped and V-shaped reflects that SFE
34–37.
increases with temperature. [14] H.Y. Zhan, G. Wang, D. Kent, M. Dargusch, Acta Mater. 105 (2016)
The formation of ␣” during straining has been known as 104–113.
stress-induced martensitic transformation [13]. According to a [15] Y.H. Lin, S.M. Wu, F.H. Kao, S.H. Wang, J.R. Yang, C.C. Yang, C.S. Chiou, Mater.
Sci. Eng. A 528 (2011) 2271–2276.
computational thermodynamic study on displacive transforma- [16] M. Ahmed, D. Wexler, G. Casillas, D.G. Savvakin, E.V. Pereloma, Acta Mater.
tions in titanium alloys [48], Zr, normally as an ␣ stablizer, has 104 (2016) 190–200.
an anomalous ␤ stablizing effect in a Ti-Nb-Zr ternary system, and [17] X. Ji, S. Emura, X.H. Min, K. Tsuchiya, Mater. Sci. Eng. A 707 (2017)
701–707.
the Ms of such a ternary system is generally near room temper- [18] S. Sadeghpour, S.M. Abbasi, M. Morakabati, J. Alloys Compd. 650 (2015)
ature. Recently, a few studies [49,50] have provided insight into 22–29.
the phase stability through combined thermodynamics and kinet- [19] M. Ahmed, D. Wexler, G. Casillas, O.M. Ivasishin, E.V. Pereloma, Acta Mater. 84
(2015) 124–135.
ics considerations. In the current study, deformation supplies extra [20] J.F. Xiao, Z.H. Nie, C.W. Tan, G. Zhou, R. Chen, M.R. Li, X.D. Yu, X.C. Zhao, S.X.
energy and increases the Ms [51] for the SI␣” transformation from Hui, W.J. Ye, Y.T. Lee, Mater. Sci. Eng. A 751 (2019) 191–200.
room temperature to above 200 ◦ C. In addition, a higher strain rate [21] T. Saito, T. Furuta, J.H. Hwang, S. Kuramoto, K. Nishino, N. Suzuki, R. Chen, A.
Yamada, K. Ito, Y. Seno, Mater. Sci. Forum 426-432 (2003) 681–688.
means more energy supplied and a higher Ms achieved, therefore
[22] T. Saito, T. Furuta, J.H. Hwang, S. Kuramoto, K. Nishino, N. Suzuki, R. Chen, A.
more SIM␣” phase is expected. Temperature has a negative effect Yamada, K. Ito, Y. Seno, Science 300 (2003) 464–467.
on martensitic transformation. As a result, the propensity of SIM␣” [23] S.Y. Sun, C. Deng, Titanium Ind. Prog. 28 (2011) 21–25.
[24] L. Qu, Y. Yang, Y.F. Lu, L. Feng, J.H. Ju, P. Ge, W. Zhou, D. Han, D.H. Ping, Scr.
decreases with an increased temperature.
Mater. 69 (2013) 389–392.
[25] J.W. Edington, Philos. Mag. 19 (1969) 1189–1206.
5. Conclusion [26] G.Z. Voyiadjis, F.H. Abed, Mech. Mater. 37 (2005) 355–378.
[27] T. Akanuma, H. Matsumoto, S. Sato, A. Chiba, I. Inagaki, Y. Shirai, T. Maeda, Scr.
Mater. 67 (2012) 21–24.
In this study, a Ti-23.1Nb-2.0Zr-1.0O (at.%) alloy was com- [28] N. Bernstein, E.B. Tadmor, Phys. Rev. B 69 (2004) 2507–2519.
pressed at a strain rate range of 0.1-1000 s−1 and temperatures [29] P. Li, K.M. Xue, Y. Lu, J.R. Tan, J. Mater. Sci. Technol. 19 (2003) 161–163.
between 100 ◦ C and 200 ◦ C. The effect of strain rate and temperature [30] K.A. Ofei, L. Zhao, J. Sietsma, J. Mater. Sci. Technol. 29 (2013) 161–167.
[31] V.S.Y. Injeti, Z.C. Li, B. Yu, R.D.K. Misra, Z.H. Cai, H. Ding, J. Mater. Sci. Technol.
on the deformation behavior was investigated. In the specimens 34 (2018) 745–755.
deformed at a higher strain rate and/or a lower temperature, multi- [32] F.Y. Lui, P. Yang, L. Meng, F. Cui, H. Ding, J. Mater. Sci. Technol. 27 (2011)
ple deformation modes were observed, which includes dislocation 257–265.
[33] H. Xing, Transmission Electron Microscopy Study on Microstructure of ␤ Type
slip, {332}113 twinning, {112}111 twinning, stress-induced ␣” Ti-Alloys, Ph.D. Thesis, Shanghai Jiao Tong University, Shanghai, 2008 (in
martensitic and stress-induced ␻ phase transformations. Within Chinese).
the {112}111 twins, stress-induced ␻ precipitates were observed. [34] H. Xing, J. Sun, Appl. Phys. Lett. 93 (2008) 669–672.
[35] R.J. Talling, R.J. Dashwood, M. Jackson, D. Dye, Acta Mater. 57 (2009)
Twinning and SI␻ transformations were favorably activated at 1188–1198.
lower temperatures and/or higher strain rates. Such trends are [36] Y. Yang, G.P. Li, H. Wang, S.Q. Wu, L.C. Zhang, Y.L. Li, K. Yang, Scr. Mater. 66
proposed to arise from a higher stress concentration at ␤ grain (2012) 211–214.
[37] S. Kuramoto, T. Furuta, J.H. Hwang, K. Nishino, T. Saito, J. Jpn. Inst. Met. 69
boundaries and an extra nucleation energy supplied by a higher
(2005) 953–961.
strain rate, as well as a lower stacking fault energy at a lower tem- [38] E.B. Tadmor, N. Bernstein, J. Mech. Phys. Solids 52 (2004) 2507–2519.
perature. [39] L.L. Liu, Generalized Stacking Fault Energy, Twinning and Dislocations of
Materials at Finite Temperature Based on the First-Principles Study and
Quasiharmonic Approximation, Ph.D. Thesis, Chongqing University, 2015 (in
Acknowledgements Chinese).
[40] M. Wei, First-Principles Study of Displacive Phase Transitions and Plastic
Deformation Mechanisms in ␤ Ti-V Alloys, Ph.D. Thesis, Shanghai Jiao Tong
This research is financially supported by the internal funding
University, 2017 (in Chinese).
source from University of Shanghai for Science and Technology, [41] Y. Takemoto, M. Hida, A. Sakakibara, J. Jpn. Inst. Met. 57 (1993) 1471–1472.
Frontier and Key Projects of the Chinese Academy of Sciences (No. [42] L.E. Murr, M.A. Meyers, C.-S. Niou, Y.J. Chen, S. Pappu, C. Kennedy, Acta Mater.
QYZDJ-SSW-JSC031-01) and the China National Key Laboratory 45 (1997) 157–175.
[43] L.M. Hsiung, D.H. Lassila, Scr. Mater. 38 (1998) 1371–1376.
Foundation of Science and Technology on Materials under Shock [44] R.C. Koo, J. Less Common Met. 4 (1962) 138–144.
and Impact (No. 6142902190501). [45] L. Ren, W. Xiao, C. Ma, R. Zheng, L. Zhou, Scr. Mater. 156 (2018) 47–50.
[46] Y. Fu, W. Xiao, D. Kent, M.S. Dargusch, J. Wang, X. Zhao, C. Ma, Scr. Mater. 187
(2020) 285–290.
References [47] L. Ren, W. Xiao, D. Kent, M. Wan, C. Ma, L. Zhou, Scr. Mater. 184 (2020) 6–11.
[48] J.-Y. Yan, G.B. Olson, J. Alloys Compd. 673 (2016) 441–454.
[1] M. Niinomi, Sci. Technol. Adv. Mater. 4 (2003) 445–454. [49] Q. Luo, Y. Guo, B. Liu, Y. Feng, J. Zhang, Q. Li, K. Chou, J. Mater. Sci. Technol. 44
[2] S. Naka, Curr. Opin. Solid State Mater. Sci. 1 (1996) 333–339. (2020) 171–190.
[3] H.X. Jin, K.X. Wei, J.M. Li, J.Y. Zhou, W.J. Peng, Chin. J. Nonferrous Met. 25 [50] Q. Luo, C. Zhai, D. Sun, W. Chen, Q. Li, J. Mater. Sci. Technol. 35 (2019)
(2015) 280–292. 2115–2120.
[4] J. Chen, F.C. Ma, P. Liu, C.H. Wang, X.K. Liu, W. Li, Q.Y. Han, J. Mater. Eng. [51] M. Song, S.Y. He, K. Du, Z.Y. Huang, T.T. Yao, Y.L. Hao, S.J. Li, R. Yang, H.Q. Ye,
Perform. 28 (2019) 1410–1418. Acta Mater. 118 (2016) 120–128.

60

You might also like