Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Materials Science & Engineering A 640 (2015) 443–448

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Ultrafine-grained magnesium–lithium alloy processed by high-


pressure torsion: Low-temperature superplasticity and potential
for hydroforming
Hirotaka Matsunoshita a,1, Kaveh Edalati a,b,n, Mitsuaki Furui c,2, Zenji Horita a,b
a
Department of Materials Science and Engineering, Faculty of Engineering, Kyushu University, Fukuoka 819-0395, Japan
b
WPI, International Institute for Carbon-Neutral Energy Research (WPI-I2CNER), Kyushu University, Fukuoka 819-0395, Japan
c
Graduate School of Science and Engineering for Research, University of Toyama, Toyama 930-8555, Japan

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

Article history: A Mg–Li alloy with 8 wt% Li was processed by severe plastic deformation (SPD) through the process of
Received 20 April 2015 high-pressure torsion (HPT) to achieve ultrafine grains with an average grain size of  500 nm. Tensile
Received in revised form testing with an initial strain rate of 10  3 s  1 showed that the alloy exhibited superplasticity at a tem-
28 May 2015
perature of 323 K or higher. Tensile testing in boiling water confirmed that the specimens were elongated
Accepted 29 May 2015
to 350–480% at 373 K under the initial strain rates of 10  3 s  1 to 10  2 s  1 with a strain rate sensitivity of
Available online 9 June 2015
 0.3. The current study suggests that not only superplastic forming but also superplastic hydroforming
Keywords: should be feasible after the grain refinement using the HPT method.
Mg–Li alloys & 2015 Elsevier B.V. All rights reserved.
Superplastic deformation
Ultrafine-grained materials (UFG)
Severe plastic deformation (SPD)
High-pressure torsion (HPT)
Strain rate sensitivity

1. Introduction Following the pioneering works by Valiev et al. [8–10], the


occurrence of superplasticity was reported in different HPT-pro-
Severe plastic deformation (SPD) methods are promising cessed aluminum alloys (Al-1420 [11–14], Al-1421 [15], Al-1570
techniques to produce ultrafine-grained materials [1,2]. Among [16], Al-2024 [17], Al-7034 [18], Al-7075 [19,20], Al–3%Mg–0.2%Sc
different SPD methods, the method of high-pressure torsion (HPT) [21] and Al–6%Cu–0.4%Zr [22]), magnesium alloys (Zk60 [23,24],
is currently receiving increasing attention for both experimental AZ61 [25], Mg–9%Al [26] and Mg–10%Gd [27]), titanium alloys (Ti–
studies [3,4] and theoretical modeling [5,6]. The HPT method was 6%Al–4%V [12,28–30]), aluminum–copper alloys (Al–33%Cu [31]),
originally introduced by Bridgman in 1935 for investigation of copper alloys (Cu–0.1%Zr [32]), lead–tine alloys (Pb–62%Sn [33]),
shear strength and phase transformation in different materials zinc–aluminum alloys (Zn–22%Al [33–37]), bismuth–tin alloys (Bi–
under high pressure [7]. The HPT method received limited atten- 42%Sn [38]) and intermetallics (Ni3Al [11–13,39–42]). The occur-
tion until 1988, when Valiev et al. reported that the grain refine- rence of superplasticity was reported not only after HPT proces-
ment achieved by HPT processing can result in superplasticity in sing of thin disc specimens [8–42], but also after HPT processing of
an Al–4%Cu–0.5%Zr alloy [8]. In 1991 and 1993, Valiev et al. re- an Al–3%Mg–0.2%Sc alloy using the ring samples [43], bulk sam-
ported superplastic behavior in several other HPT-processed alloys ples [44,45] as well as after continuous high-pressure torsion [46].
such as Mg–l.5%Mn–0.3%Ce [9] and Ti–6.5%A1–3.2%Mo–0.3%Si Despite numerous reports on superplastic behavior of SPD-
[10]. processed materials (see reviews in [47–51]), there are few reports
on superplastic hydroforming of SPD-processed materials. A gen-
n
eral requirement for superplastic hydroforming in water is that the
Corresponding author at: Department of Materials Science and Engineering,
Faculty of Engineering, Kyushu University, Fukuoka 819-0395, Japan. material should exhibit a tensile elongation of at least 400% before
Fax: þ81 92 802 2992 . breaking at temperatures close to 373 K (with a strain rate sensi-
E-mail address: kaveh.edalati@zaiko6.zaiko.kyushu-u.ac.jp (K. Edalati). tivity of  0.5 [52]). Mg–Li alloys which have a high strength to
1
Present affiliation: Department of Materials Science and Engineering, Kyoto
weight ratio and are used for transportation systems, are among
University, Sakyo-ku, Kyoto 606-8501, Japan.
2
Present affiliation: Department of Mechanical Engineering, School of En- the candidates for superplastic hydroforming [53]. There are cur-
gineering, Tokyo University of Technology, Hachioji 192-0982, Japan. rently wide activities on SPD processing of Mg alloys (see a recent

http://dx.doi.org/10.1016/j.msea.2015.05.103
0921-5093/& 2015 Elsevier B.V. All rights reserved.
444 H. Matsunoshita et al. / Materials Science & Engineering A 640 (2015) 443–448

review in [54]) including a few reports on grain refinement in


HPT-processed pure Mg [55–57] and Mg–Li alloys [58–60]. The
occurrence of superplasticity was reported in an Mg–8%Li alloy
after processing by equal-channel angular pressing (ECAP) [61,62].
The current study was conducted to examine the superplastic
hydroforming potential of an HPT-processed Mg–8%Li alloy in hot
water.

2. Experimental materials and methods

An Mg–8%Li alloy (in wt%) was received in the form of an ex-


truded rod with 10 mm diameter (extruded at 373 K with a speed
of 1 mm/s and an extrusion ratio of 25:1). The microstructure of
samples before and after extrusion is shown in Fig. 1(a) and (b),
respectively, indicating the coexistence of α (Mg-rich phase with a
hcp structure) phase and eutectic α þ β (Li-rich phase with bcc
structure) phases in consistency with the Mg–Li phase diagram
Fig. 2. Appearance of tensile testing facility for pulling samples in boiling water.
[63]. The rod was sliced to discs with thicknesses of 0.9 mm using
a wire-cutting electrical discharge machine.
HPT was conducted using a facility consisting of upper and
microstructures. First of all, the disc samples were polished to a
lower anvils having a shallow hole of 10 mm diameter and
mirror-like surface and the Vickers microhardness was measured
0.25 mm depth at the center. Each sample was placed on the hole
along the radii from the center to edge at 8 different radial di-
and the upper and lower anvils were rotated with respect to each
rections with 1 mm increments. For each hardness measurement,
other at room temperature with a rotation speed of 1 rpm under a
a load of 50 g was applied for 15 s. Second, the disc samples were
pressure of 3 GPa. The rotation was terminated after 5 turns.
ground and polished to reduce the total thickness to 0.6 mm.
The disc samples subjected to HPT processing were evaluated
Thereafter, tensile specimens having 1 mm gauge length and
in terms of Vickers microhardness, tensile properties and
1 mm width were cut from the samples using the wire-cutting
electrical discharge machining. Each tensile specimen was moun-
ted on grips and pulled to failure at selected temperatures in the
range of 298–473 K with an initial strain rate of 1  10  3 s  1 to
5  10  2 s  1. To obtain the average mechanical properties, two to
four tensile tests were conducted on the samples examined at the
same tensile conditions. In addition to tensile testing in a furnace
under elevated temperatures, several tensile specimens were ex-
amined at 373 K in boiling water using the facility shown in Fig. 2.
The tensile tests in boiling water were conducted to examine the
possible positive or negative effects of hot water on mechanical
properties of the materials during hydroforming. Third, discs with
3 mm in diameter were punched from the HPT discs at 2 mm away
from the center. The 3 mm discs were ground mechanically to a
thickness of 0.15 mm and further thinned with a twin-jet electro-
chemical polisher using a solution of 5% HClO4, 25% C3H5(OH)3 and
70% C2H5OH. Transmission electron microscopy (TEM) was per-
formed at 200 kV for microstructural observation and for record-
ing selected-area electron diffraction (SAED) patterns.

3. Results and discussion

A TEM bright-field image, corresponding SAED pattern and


dark-field image are shown for the sample after HPT processing for
5 turns in Fig. 3(a), (b) and (c), respectively. It is apparent that the
microstructure consists of small grains with an average grain size
of  500 nm. The average grain size achieved in this Mg–Li alloy is
smaller than those reported in the HPT-processed pure Mg
( 1 μm [55,56]), due to the effect of solute atoms on the mobility
of dislocations, as discussed in an earlier paper [64]. Earlier works
reported that the grain size of Mg–Li alloys are reduced to the
submicrometer level after processing by HPT at room temperature
[58–60] and to a few micrometers after processing by ECAP at
373–573 K [61,62]. It should be noted that processing of Mg–Li
using ECAP is only feasible at elevated temperatures because of the
brittle behavior of the alloy at room temperature, whereas the
Fig. 1. Optical micrographs of samples (a) before extrusion and (b) after extrusion. alloy could be processed at room temperature in this study
H. Matsunoshita et al. / Materials Science & Engineering A 640 (2015) 443–448 445

Fig. 5. Nominal stress versus nominal strain curves obtained at different tem-
peratures with initial strain rate of 1  10  3 s  1 for samples processed by HPT for
5 turns.

extruded rod (48 Hv).


The stress–strain curves for the HPT-processed samples are
Fig. 3. TEM observation after HPT processing for 5 turns: (a) bright-field image,
(b) SAED pattern and (c) dark-field image.
delineated in Fig. 5 from tensile tests conducted at 298 K, 323 K,
343 K, 373 K, 423 K and 473 K at an initial strain rate of
because of high processing pressure. 1  10  3 s  1. The tensile strength decreases and the total elon-
The variations of hardness are plotted in Fig. 4 against the gation to failure increases with an increase in the temperature. The
samples pulled to failure at a temperature of 323 K or higher ex-
distance from the disc center after 5 turns. Microhardness in-
hibit elongations greater than 400%, indicating that superplasticity
creases with the distance from the disc center, but the hardness
is feasible in this material after HPT processing. It is emphasized
does not saturate to a steady state. The absence of a steady-state
that the superplasticity is attained at a low temperature as at
level indicates that the saturation level (which has been reported
323 K which is equivalent to 0.37 Tm (Tm: melting point of the
in many metals and alloys [4,56,64]) should appear at larger
alloy). The appearance of the tensile specimens before and after
numbers of turns in this two-phase alloy. Moreover, the strain
the tensile testing is shown in Fig. 6. Typical superplastic flows
softening that was reported at large strains in pure Mg [55–57] is
appear with smooth and uniform elongation throughout the gauge
not seen in this alloy because of the suppression of recrystalliza-
length. It is demonstrated clearly that processing by HPT leads to
tion due to solute atoms. The microhardness levels after the HPT
the advent of superplasticity with a maximum superplastic elon-
processing are 57–63 Hv, which are higher than that for the
gation of 1330% at 473 K. It should be noted that tensile strength at
room temperature deviates in the range of 110–160 MPa due to the
heterogeneity in the microstructure at the non-steady-state con-
dition of this study.
Since the mechanical properties deviate in a wide range in this
study, two to four tensile tests were conducted on the samples
examined at the same tensile testing conditions. Elongation to
failure is plotted in Fig. 7 against the tensile testing temperature at
an initial strain rate of 1  10  3 s  1. Although the values of elon-
gation to failure scatter, the samples pulled to failure at a tem-
perature of 373 K or higher always exhibit superplastic elongation
of 4400%. Because the temperature of 373 K is a critical tem-
perature for hydroforming in hot water at ambient pressure, Fig. 7
suggests that the HPT-processed Mg–Li alloy has good potential for
it.
The stress-strain curves for the HPT-processed sample are
shown in Fig. 8 from tensile testing conducted at 373 K in (a) air
and (b) hot water at initial strain rates of 1  10  3 s  1 to
5  10  2 s  1. The elongation to failure at an initial strain rate of
1  10  2 s  1 is as high as 330% and 350% for the samples pulled in
air and hot water, respectively. The samples pulled at lower strain
rates as 1  10  3 s  1 and 5  10  3 s  1 clearly exhibit superplastic
elongations of 4 400%. The superplasticity appears not only in air
Fig. 4. Vickers microhardness plotted against distance from disc center after HPT but also in hot water, suggesting that hydroforming should be
processing for 5 turns, including hardness level for extruded rod. feasible in the Mg–Li alloy when it is processed by HPT. It is noted
446 H. Matsunoshita et al. / Materials Science & Engineering A 640 (2015) 443–448

Fig. 6. Appearance of HPT-processed samples after pulling to failure at different


temperatures with initial strain rate of 1  10  3 s  1, including appearance of un-
deformed sample.

Fig. 8. Nominal stress versus nominal strain curves obtained at different strain
rates at 373 K in (a) air and (b) boiling water for samples processed by HPT for
5 turns.

are some slight differences in the morphology of the surface for


the samples pulled in air and in water. The XRD analyses suggest
that the possible formation of Li2O, MgO, LiOH and Mg(OH)2
compounds should be limited to very thin layer on the surface
Fig. 7. Plots of elongation to failure against temperature for HPT-processed sam-
because these compounds could not be detected within the sen-
ples after pulling to failure under strain rate of 1  10  3 s  1. sitivity limits of the XRD analyses. It is noted that the precise
compositions of thin oxides and hydroxides on the surface and
that serration is visible for the sample pulled in hot water. The subsurface should be examined using surface analysis methods
reason was carefully checked by pulling a rubber band in hot water such as X-ray photoelectron spectroscopy (XPS).
and the serration was detected in almost the same way with si- Superplastic flow of materials is generally described by the
milar frequency. It turned out that this periodic serration arose following equation [52]:
from the driving system in hot water and it is not due to the effect ADGb ⎛ b ⎞ ⎜⎛ σ ⎟⎞n
p

of solute atoms on dynamic strain hardening behavior and Por- ε̇ = ⎜ ⎟


kT ⎝ d ⎠ ⎝ G ⎠ (1)
tevin–LeChatelier instability. The appearance of the tensile speci-
mens before and after testing is shown in Fig. 9 for the samples where ε̇ is the strain rate, A is a constant, D is the diffusion coef-
pulled in (a) air and (b) hot water. The specimens become more ficient, G is the shear modulus, b is the Burgers vector, k is Boltz-
uniformly deformed and the elongation to failure improves by mann's constant, T is the absolute temperature, d is the grain size,
decreasing the strain rate. s is the flow stress, p is the grain size exponent and n is the stress
The oxidation of the samples in hot water was roughly ex- exponent of which reciprocal, m ¼1/n, is called the strain rate
amined using the scanning electron microscopy (SEM) and X-ray sensitivity. The ultimate tensile stress is plotted against the strain
diffraction (XRD) analyses. SEM micrographs indicate that there rate in Fig. 10 in a double-logarithmic scale for the HPT-processed
H. Matsunoshita et al. / Materials Science & Engineering A 640 (2015) 443–448 447

samples pulled to failure at 373 K in air and hot water. The slope of
the plot is now equivalent to m and it follows that m ¼  0.3.
Earlier works on ECAP processing and extrusion of this alloy also
reported a strain rate sensitivity of 0.2–0.4 at the testing tempera-
tures of 423 K and 473 K for strain rates in the range of 10  1 s  1 to
1 3 1
0 s [61,62]. The value of 0.3 is lower than m ¼0.5, which has
been generally accepted for superplastic deformation based on
grain boundary sliding. It is thus suggested that the deformation is
controlled not only by the contribution of grain boundary sliding
with low strain rate sensitivity but also by intragranular move-
ment of dislocations. The melting temperature (eutectic tempera-
ture) for the current Mg–Li alloy is 861 K [63] and thus the
temperatures of 323 K and 373 K correspond to homologous
temperatures of 0.37 and 0.43, respectively. The present study
demonstrates that low-temperature superplasticity has been well
attained in the Mg–8%Li alloy.

4. Conclusions

(1) An Mg–Li alloy with 8 wt% Li was successfully processed by


high-pressure torsion (HPT) at room temperature. Grain re-
finement to an average size of  500 nm with Vickers hardness
of 57–63 Hv and good tensile strength of 110–160 MPa were
achieved at room temperature.
(2) The HPT processing led to superplastic elongation of 4400%
with an initial strain rate of 1  10  3 s  1 at temperatures
higher than 323 K equivalent to 0.37Tm (Tm: melting tem-
perature). In this study, a maximum elongation of 1330% was
achieved with an initial strain rate of 1  10  3 s  1 at 473 K.
(3) Tensile testing in boiling water with the initial strain rates of
1  0  3 s  1 to 1  10  2 s  1 resulted in superplastic elonga-
tions of 350–480%, indicating the potential for hydroforming
in hot water.

Acknowledgment

The authors would like to thank Mr. Masaaki Kai for his kind
Fig. 9. Appearance of HPT-processed samples after pulling to failure with different
strain rates at 373 K in (a) air and (b) boiling water, including appearance of un-
assistance. This work was supported in part by the Light Metals
deformed samples. Educational Foundation of Japan, in part by a Grant-in-Aid, in In-
novative Areas “Bulk Nanostructured Metals” (No. 22102004) and
in part by a Grant-in-Aid and for Challenging Exploratory Research
from the MEXT, Japan (Nos. 26220909 and 15K14183). The HPT
process was carried out in the International Research Center on
Giant Straining for Advanced Materials (IRC-GSAM) at Kyushu
University.

References

[1] R.Z. Valiev, R.K. Islamgaliev, I.V. Alexandrov, Bulk nanostructured materials
from severe plastic deformation, Prog. Mater. Sci. 45 (2000) 103–189.
[2] R.Z. Valiev, Y. Estrin, Z. Horita, T.G. Langdon, M.J. Zehetbauer, Y.T. Zhu, Pro-
ducing bulk ultrafine-grained materials by severe plastic deformation, J. Mi-
ner. Met. Mater. Soc. 58 (4) (2006) 33–39.
[3] A.P. Zhilyaev, T.G. Langdon, Using high-pressure torsion for metal processing:
Fundamentals and applications, Prog. Mater. Sci. 53 (2008) 893–979.
[4] R. Pippan, S. Scheriau, A. Taylor, M. Hafok, A. Hohenwarter, A. Bachmaier,
Saturation of fragmentation during severe plastic deformation, Annu. Rev.
Mater. Res. 40 (2010) 319–343.
[5] H.S. Kim, Finite element analysis of high pressure torsion processing, J. Mater.
Process. Technol. 113 (2001) 617–621.
[6] D.J. Lee, E.Y. Yoon, D.H. Ahn, B.H. Park, H.W. Park, L.J. Park, Y. Estrin, H.S. Kim,
Dislocation density-based finite element analysis of large strain deformation
Fig. 10. Maximum flow stress plotted against initial strain rate for samples pro- behavior of copper under high-pressure torsion, Acta Mater. 76 (2014)
cessed by HPT for 5 turns and pulled to failure at 373 K in (a) air and (b) boiling 281–293.
water. Slope shows strain rate sensitivity of m¼ 0.3. [7] P.W. Bridgman, Effects of high shearing stress combined with high hydrostatic
448 H. Matsunoshita et al. / Materials Science & Engineering A 640 (2015) 443–448

pressure, Phys. Rev. 48 (1935) 825–847. and mechanical properties in a Zn–Al eutectoid alloy processed by high-
[8] R.Z. Valiev, O.A. Kaibyshev, R.I. Kuznetsov, R.S. Musalimov, N.K. Tsenev, Low- pressure torsion, Acat Mater. 72 (2014) 67–79.
temperature superplasticity of metallic materials, Dokl. Akad. Nauk. SSSR 301 [37] M. Kawasaki, T.G. Langdon, Microstructure development and superplasticity in
(1988) 864–866. a Zn–22%Al eutectoid alloy processed by severe plastic deformation, Mater.
[9] R.Z. Valiev, N.A. Krasiinikov, N.K. Tsenev, Plastic deformation of alloys with Sci. Forum 783–786 (2014) 2647–2652.
submicron-grained structure, Mater. Sci. Eng. A 137 (1991) 35–40. [38] C.T. Wang, T.G. Langdon, Strain weakening and superplasticity in a Bi–Sn eu-
[10] R.Z. Valiev, A.V. Korznikov, R.R. Mulyukov, Structure and properties of ultra- tectic alloy processed by high-pressure torsion, IOP Conf. Ser. Mater. Sci. Eng.
fine-grained materials produced by severe plastic deformation, Mater. Sci. Eng. 63 (2014) 012107.
A 168 (1993) 141–148. [39] N.A. Mara, A.V. Sergueeva, T.D. Mara, S.X. McFadden, A.K. Mukherjee, Super-
[11] S.X. McFadden, R.S. Mishra, R.Z. Valiev, A.P. Zhilyaev, A.K. Mukherjee, Low- plasticity and cooperative grain boundary sliding in nanocrystalline Ni3Al,
temperature superplasticity in nanostructured nickel and metal alloys, Nature Mater. Sci. Eng. A 463 (2007) 238–244.
398 (1999) 684–686. [40] S.X. McFadden, R.Z. Valiev, A.K. Mukherjee, Superplasticity in nanocrystalline
[12] S.X. McFadden, A.V. Sergueeva, R.S. Mishra, A.K. Mukherjee, High strain rate Ni3Al, Mater. Sci. Eng. A 319–321 (2001) 849–853.
superplasticity in microcrystalline and nanocrystalline materials, Mater. Sci. [41] R.S. Mishra, R.Z. Valiev, S.X. McFadden, A.K. Mukherjee, Tensile superplasticity
Technol. 16 (2000) 1340–1344. in a nanocrystalline nickel aluminide, Mater. Sci. Eng. A 252 (1998) 174–178.
[13] R.K. Islamgaliev, R.Z. Valiev, R.S. Mishra, A.K. Mukherjee, Enhanced super- [42] R.Z. Valiev, C. Song, S.X. McFadden, A.K. Mukherjee, R.S. Mishra, TEM/HREM
plastic properties in bulk metastable nanostructured alloys, Mater. Sci. Eng. A observations of nanostructured superplastic Ni3Al, Philos. Mag. A 81 (2001)
304–306 (2001) 206–210. 25–36.
[14] R.S. Mishra, R.Z. Valiev, S.X. McFadden, R.K. Islamgaliev, A.K. Mukherjee, High- [43] Y. Harai, K. Edalati, Z. Horita, T.G. Langdon, Using ring samples to evaluate the
strain-rate superplasticity from nanocrystalline Al alloy 1420 at low tem- processing characteristics in high-pressure torsion, Acta Mater. 57 (2009)
peratures, Philos. Mag. A 81 (2001) 37–48. 1147–1153.
[15] R.K. Islamgaliev, N.F. Yunusova, R.Z. Valiev, The influence of the SPD tem- [44] Z. Horita, T.G. Langdon, Achieving exceptional superplasticity in a bulk alu-
perature on superplasticity of aluminium alloys, Mater. Sci. Forum 503–504 minum alloy processed by high-pressure torsion, Scr. Mater. 58 (2008)
(2006) 585–590. 1029–1032.
[16] V.N. Perevezentsev, M.Y. Shcherban, M.Y. Murashkin, R.Z. Valiev, High-strain- [45] G. Sakai, K. Nakamura, Z. Horita, T.G. Langdon, Application of high pressure
rate superplasticity of nanocrystalline aluminum alloy 1570, Tech. Phys. Lett torsion to bulk samples, Mater. Sci. Forum 503–504 (2006) 391–398.
33 (2007) 648–650. [46] K. Edalati, Z. Horita, Processing sheets and wires by continuous high-pressure
[17] S.V. Dobatkin, E.N. Bastarache, G. Sakai, T. Fujita, Z. Horita, T.G. Langdon, Grain torsion, Rev. Adv. Mater. Sci. 31 (2012) 5–11.
refinement and superplastic flow in an aluminum alloy processed by high- [47] Y. Ma, M. Furukawa, Z. Horita, M. Nemoto, R.Z. Valiev, T.G. Langdon, Sig-
pressure torsion, Mater. Sci. Eng. A 408 (2005) 141–146. nificance of microstructural control for superplastic deformation and forming,
[18] C. Xu, S.V. Dobatkin, Z. Horitac, Terence G. Langdon, Superplastic flow in a Mater. Trans. 37 (1996) 336–339.
nanostructured aluminum alloy produced using high-pressure torsion, Mater. [48] F.A. Mohamed, Y. Li, Creep and superplasticity in nanocrystalline materials:
Sci. Eng. A 500 (2009) 170–175. current understanding and future prospects, Mater. Sci. Eng. A 298 (2001)
[19] J.M. Garcia-Infanta, A.P. Zhilyaev, A. Sharafutdinov, O.A. Ruano, F. Carreno, An 1–15.
evidence of high strain rate superplasticity at intermediate homologous [49] R.Z. Valiev, R.K. Islamgaliev, I.P. Semenova, Superplasticity in nanostructured
temperatures in an Al-Zn-Mg-Cu alloy processed by high-pressure torsion, J. materials: new challenges, Mater. Sci. Eng. A 463 (2007) 2–7.
Alloy. Compd. 473 (2009) 163–166. [50] T.G. Langdon, Seventy-five years of superplasticity: historic developments and
[20] S. Sabbaghianrad, M. Kawasaki, T.G. Langdon, Microstructural evolution and new opportunities, J. Mater. Sci. 44 (2009) 5998–6010.
the mechanical properties of an aluminum alloy processed by high-pressure [51] M. Kawasaki, T.G. Langdon, Review: achieving superplasticity in metals pro-
torsion, J. Mater. Sci. 47 (2012) 7789–7795. cessed by high-pressure torsion, J. Mater. Sci. 49 (2014) 6487–6496.
[21] G. Sakai, Z. Horita, T.G. Langdon, Grain refinement and superplasticity in an [52] T.G. Langdon, A unified approach to grain boundary sliding in creep and su-
aluminum alloy processed by high-pressure torsion, Mater. Sci. Eng. A 393 perplasticity, Acta Metall. Mater. 42 (1994) 2437–2443.
(2005) 344–351. [53] T.C. Chang, J.Y. Wang, C.L. Chu, S. Lee, Mechanical properties and micro-
[22] A. Alhamidi, Z. Horita, Application of high-pressure torsion to Al–6%Cu–0.4%Zr structures of various Mg–Li alloys, Mater. Lett. 60 (2006) 3272–3276.
alloy for ultrafine-grain refinement and superplasticity, J. Mater. Sci. 49 (2014) [54] R.B. Figueiredo, M.T.P. Aguilar, P.R. Cetlin, T.G. Langdon, Processing magnesium
6689–6695. alloys by severe plastic deformation, IOP Conf. Se. Mater. Sci. Eng. 63 (2014)
[23] S.A. Torbati-Sarraf, T.G. Langdon, Mechanical properties of ZK60 magnesium 012171.
alloy processed by high-pressure torsion, Adv. Mater. Res. 922 (2014) 767–772. [55] K. Edalati, A. Yamamoto, Z. Horita, T. Ishihara, High-pressure torsion of pure
[24] S.A. Torbati-Sarraf, T.G. Langdon, Properties of a ZK60 magnesium alloy pro- magnesium: Evolution of mechanical properties, microstructures and hydro-
cessed by high-pressure torsion, J. Alloy. Compd. 613 (2014) 357–363. gen storage capacity with equivalent strain, Scr. Mater. 64 (2011) 880–883.
[25] Y. Harai, M. Kai, K. Kaneko, Z. Horita, T.G. Langdon, Microstructural and me- [56] K. Edalati, J.M. Cubero-Sesin, A. Alhamidi, I.F. Mohamed, Z. Horita, Influence of
chanical characteristics of AZ61 magnesium alloy processed by high-pressure severe plastic deformation at cryogenic temperature on grain refinement and
torsion, Mater. Trans. 49 (2008) 76–83. softening of pure metals: investigation using high-pressure torsion, Mater. Sci.
[26] M. Kai, Z. Horita, T.G. Langdon, Developing grain refinement and super- Eng. A 613 (2014) 103–110.
plasticity in a magnesium alloy processed by high-pressure torsion, Mater. Sci. [57] M. Joshi, Y. Fukuta, S. Gao, N. Park, D. Terada, N. Tsuji, Fabrication of fine re-
Eng. A 488 (2008) 117–124. crystallized grains and their mechanical property in HPT processed pure
[27] O.B. Kulyasova, R.K. Islamgaliev, A.R. Kilmametov, R.Z. Valiev, Superplastic magnesium, IOP Conf. Ser. Mater. Sci. Eng. 63 (2014) 012074.
behavior of magnesium-based Mg–10 wt% Gd alloy after severe plastic de- [58] Z. Ranachowski, P. Ranachowski, F. Rejmund, A. Pawelek, A. Piatkowski,
formation by torsion, Phys. Met. Metallogr. 101 (2006) 585–590. Z. Jasienski, The study of influence of high pressure torsion process on acoustic
[28] R.S. Mishra, V.V. Stolyarov, C. Echer, R.Z. Valiev, A.K. Mukherjee, Mechanical emission activity of compressed Mg–Li alloys, Arch. Acoust. 33 (2008)
behavior and superplasticity of a severe plastic deformation processed na- 123–128.
nocrystalline Ti–6Al–4V alloy, Mater. Sci. Eng. A 298 (2001) 44–50. [59] S. Kudela, A. Pawelek, Z. Ranachowski, A. Piatkowski, S. Kudela Jr.,
[29] A.V. Sergueeva, V.V. Stolyarov, R.Z. Valiev, A.K. Mukherjee, Superplastic be- P. Ranachowski, Effect of Al alloying on the Hall–Petch strengthening and AE
haviour of ultrafine-grained Ti–6A1–4V alloys, Mater. Sci. Eng. A 323 (2002) in compressed Mg–Li–Al alloys before and after HPT processing, Kovove Ma-
318–325. ter. 49 (2011) 271–277.
[30] A.V. Sergueeva, V.V. Stolyarov, R.Z. Valiev, A.K. Mukherjee, Enhanced super- [60] B. Srinivasarao, A.P. Zhilyaev, I. Gutierrez-Urrutia, M.T. Perez-Prado, Stabili-
plasticity in a Ti–6Al–4V alloy processed by severe plastic deformation, Scr. zation of metastable phases in Mg–Li alloys by high-pressure torsion, Scr.
Mater. 43 (2000) 819–824. Mater. 68 (2013) 583–586.
[31] M. Kawasaki, J. Foissey, T.G. Langdon, Development of hardness homogeneity [61] M. Furui, H. Kitamura, H. Anada, T.G. Langdon, Influence of preliminary ex-
and superplastic behavior in an aluminum–copper eutectic alloy processed by trusion conditions on the superplastic properties of a magnesium alloy pro-
high-pressure torsion, Mater. Sci. Eng. A 561 (2013) 118–125. cessed by ECAP, Acta Mater. 55 (2007) 1083–1091.
[32] J. Wongsa-Ngam, M. Kawasaki, Y. Zhao, T.G. Langdon, Microstructural evolu- [62] M. Furui, C. Xu, T. Aida, M. Inoue, H. Anada, T.G. Langdon, Improving the su-
tion and mechanical properties of a Cu–Zr alloy processed by high-pressure perplastic properties of a two-phase Mg–8%Li alloy through processing by
torsion, Mater. Sci. Eng. A 528 (2011) 7715–7722. ECAP, Mater. Sci. Eng. A 410–411 (2005) 439–442.
[33] R.S. Mishra, R.Z. Valiev, A.K. Mukherjee, The observation of tensile super- [63] A.A. Nayeb-Hashemi, J.B. Clark, A.D. Pelton, The Li–Mg (lithium–magnesium)
plasticity in nanocrystalline materials, Nanostruct. Mater. 9 (1997) 473–476. system, Bull. Alloy. Phase Diag. 5 (1984) 365–374.
[34] M. Furukawa, Y. Ma, Z. Horita, M. Nemoto, R.Z. Valiev, T.G. Langdon, Micro- [64] K. Edalati, D. Akama, A. Nishio, S. Lee, Y. Yonenaga, J.M. Cubero-sesin, Z. Horita,
structural characteristics and superplastic ductility in a Zn–22%Al alloy with Influence of dislocation-solute atom interactions and stacking fault energy on
submicrometer grain size, Mater. Sci. Eng. A 241 (1998) 122–128. grain size of single-phase alloys after severe plastic deformation using high-
[35] M. Kawasaki, T.G. Langdon, Developing superplasticity and a deformation pressure torsion, Acta Mater. 69 (2014) 68–77.
mechanism map for the Zn–Al eutectoid alloy processed by high-pressure
torsion, Mater. Sci. Eng. A 528 (2011) 6140–6145.
[36] T.S. Cho, H.J. Lee, B. Ahn, M. Kawasaki, T.G. Langdon, Microstructural evolution

You might also like