Vous êtes sur la page 1sur 5

Materials Chemistry and Physics 136 (2012) 75e79

Contents lists available at SciVerse ScienceDirect

Materials Chemistry and Physics


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

Multi-step shear banding for bulk metallic glasses at ambient and cryogenic
temperatures
J.W. Qiao a, b, *, H.L. Jia c, Y. Zhang b, P.K. Liaw c, L.F. Li d
a

College of Materials Science and Engineering, Taiyuan University of Technology, Taiyuan 030024, China
State Key Laboratory for Advanced Metals and Materials, University of Science and Technology Beijing, Beijing 100083, China
c
Department of Materials Science and Engineering, The University of Tennessee, Knoxville, TN 37996-2200, USA
d
Key Laboratory of Cryogenics, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, China
b

h i g h l i g h t s
< Multi-step shearing is reasonable to analyze energy conversion upon shear banding.
< At 77 K, the serrations disappear, and an N-step method is chosen to analyze the energy dissipation.
< The present investigation gives a proper method to reveal the shear banding for BMGs at different temperatures.

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 12 December 2011
Received in revised form
13 May 2012
Accepted 16 June 2012

Analysis of energy dissipation during shear-banding aids to understand plastic deformations of bulk
metallic glasses (BMGs). For Zr55Al10Ni5Cu30 BMGs at 298 K, multi-step shearing is proposed, and the
thermal energy during serrations cannot result in a temperature rise (DT) higher than Tm. At 77 K, N-step
shearing is supposed. When N is in the range of 24e47, it is reasonable to deduce the resulting
temperature from 0.8Tg to Tm, accompanied by a continuous plastic deformation.
Crown Copyright 2012 Published by Elsevier B.V. All rights reserved.

Keywords:
Amorphous materials
Mechanical test
Deformation
Fracture

1. Introduction
Bulk metallic glasses (BMGs) are a new kind of metallic alloys,
characterized by an amorphous structure, i.e., a lack of periodic
lattice arrangements. Due to the absence of defects, such as dislocations, BMGs exhibit super-high strengths compared to their
crystalline counterparts at the temperature far below the glasstransition temperature. However, very limited plasticity is accompanied upon loading, and localized shear bands are distributed near
the fracture surface, which restricts the actual structural applications of BMGs [1]. Shear banding is the inhomogeneous plastic
deformation feature of BMGs, and shear bands are about 10 nm in
thickness, which can accommodate the plastic strain [2]. The
understanding of the shear-banding behavior aids to reveal the
deformation nature and to design the ductile BMGs.

* Corresponding author. College of Materials Science and Engineering, Taiyuan


University of Technology, Taiyuan 030024, China. Tel.: 86 351 6018051.
E-mail address: qiaojunwei@gmail.com (J.W. Qiao).

Once yielding, shear banding is available for BMGs, and the


strain energy dissipates within the thin shear layer, which leads to
the adiabatic heating instantly and the formation of a viscous shear
layer [3]. At ambient temperature, shear banding is usually associated with the serration on the stressestrain curve after yielding
[4e7]. Regardless of the intermittent shearing along one primary
shear band or the propagation of newly-formed shear bands,
serrations dominate until the nal fracture. However, at cryogenic
temperature, the serration disappears even if BMGs are subjected
to shear banding, since the rapid heat conduction promptly cools
the shear layer below the w0.8Tg [3]. The very low shear-banding
velocities at cryogenic temperature do not favor dramatic temperature rises originating from the frictional sliding effects during
shear banding [8]. Thus, only if the strain energy is easily dissipated
within the shear layer, an improved plasticity could be gained [7,9],
i.e., cold shear banding results in a large plasticity [10]. Based on
different stress ow kinetics (serration and non-serration) after
yielding at ambient and cryogenic temperatures, the shear-banding
behaviors associated with the strain-energy dissipation will be
focused on in this study.

0254-0584/$ e see front matter Crown Copyright 2012 Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.matchemphys.2012.06.033

76

J.W. Qiao et al. / Materials Chemistry and Physics 136 (2012) 75e79

2. Experimental procedure
Ingots of nominal compositions (in atomic percent, at. %),
Zr55Al10Ni5Cu30, were prepared by arc-melting the mixture of Zr, Al,
Ni, and Cu with purity higher than 99.9% (weight percent) under
a Ti-gettered argon atmosphere. The Zr55Al10Ni5Cu30 liquid alloys
had a very high glass-forming ability [11] and were suctioned into
a copper mold with a cylinder diameter of 2 mm and a length of
about 50 mm. The glass-transition temperature, Tg, determined by
the differential-scanning calorimetry (DSC) with a constant heating
rate of 20 K min1, was to be 682 K. Cylindrical specimens, with an
aspect ratio (height/diameter) of about 2, were sliced from rods
and, subsequently, well polished for the two ends. The uniaxialcompressive tests at 298 K and 77 K were performed on the
cylindrical specimens using an MTS testing machine with a strain
rate of 2  104 s1. The microstructure and the fracture and lateral
surfaces of the deformed samples were investigated by scanningelectron microscopy (SEM) to identify the fracture mechanisms.
3. Results and discussion
Fig. 1 shows the SEM image of the microstructure of Zr55Al10Ni5Cu30 BMG. It can be seen that featureless contrast is available,
indicating an amorphous structure for the present alloy.
Fig. 2(a) shows the compressive engineering stressestrain curve
at 298 K. The yielding point is considered to be that when the rst
serration appears on the stressestrain curve. The yielding stress
and strain are to be 1630 MPa and 2.74%, respectively. Upon
yielding, the serrated ows prevail until the nal fracture, and the
maximum ow stress is 1900 MPa, accompanied by a total strain of
6.77%. The magnied serration with a strain range of w3.3ew6.7%
are shown in the inset of Fig. 2(a). Within this strain range, the total
number of serrations is 47, and the magnitude of stress serrations is
from 7.0 to 43.9 MPa. Below the strain of w3.3%, the magnitude is
less than 5 MPa, and serrations are not counted in the present
study, since the low-magnitude serrations may be caused by the
machine vibration [12]. As a contrast, the compressive engineering
stressestrain curve at 77 K is displayed in Fig. 2(b). Although the
curve is comparatively smooth, i.e., the obvious serrations disappear, the stress oscillation is still available, as magnied in the inset
of Fig. 2(b), marked by numbers. The rst distinguishing oscillation
is approximately taken as the yielding point in the stressestrain
curve, and the yielding stress and strain are to be 2115 MPa and
3.86%, respectively. The maximum stress is 2150 MPa, and the total
strain is 7.10%. The elastic stage is extended at 77 K compared to
298 K, which is ascribed to that the stress-driven generation of free
volumes at low temperature requires more elastic energy, and
similar phenomena have been reported elsewhere [13]. Note that

Fig. 1. The SEM image of the microstructure for Zr55Al10Ni5Cu30 BMG.

Fig. 2. The compressive engineering stressestrain curves at (a) 298 and (b) 77 K. The
insets in (a) and (b) corresponding to the magnied stressestrain curves after yielding,
respectively.

the yield stress and the total strain increase by lowing the
temperature from 298 to 77 K.
Fractographs are given to correlate with the mechanical
behaviors and indicate the deformation mechanisms. Fig. 3(a)
presents the lateral surface of the deformed samples near the crack
at 298 K. It can be seen that the multiplication of shear bands
prevails, indicating distinct plasticity, in accordance with the
macroscopic plasticity in Fig. 2(a). No barreling is observed for the
testing samples after compression since the plasticity is accommodated within viscous shear bands. It should be noted that here
the shear bands impenetrate the whole samples, and actually,
a shear layer forms. Comparatively, the lateral surface of the
deformed sample at 77 K is displayed in Fig. 3(b). Except the very
short shear bands with an average length of only w25 mm are
observed, no other trans-sample shear bands are in the vicinity. It is
reasonable to deduce that these immature shear bands originate
the elastic perturbation of the nal fracture along the crackpropagation direction, as marked by an arrow in Fig. 3(b). In this
process, huge strain energy is forced to release so that the BMGs
cannot support. As a consequence, the immature shear bands form
as carriers to absorb the energy. For the intrinsic brittle BMGs, such
as Fe- and Mg-based BMGs, the appearance of periodic nanoscale

J.W. Qiao et al. / Materials Chemistry and Physics 136 (2012) 75e79

77

angle between the shear direction and the loading direction with
an assumed value of 45 ).
Taking the shear bands as a source of zero thickness in an
innite medium, a detailed temperature-rise (DT) prole along the
width direction of shear bands, x, and with the time, t, can be
performed by solving the heat-diffusion equation:

DT

Fig. 3. The lateral surfaces of the deformed samples at (a) 298 and (b) 77 K. The inset
in (b) showing the fracture surface at 77 K.

steps on the fracture surface dissipates the energy upon dynamic


instability [14], this and present cases are different in the approach
but equally satisfactory in the result.
The fracture surface of the deformed sample at 77 K is shown in
the inset of Fig. 3(b). It is observed that vein patterns dominate,
typical of a fracture feature for BMGs, together with many liquid
droplets highlighted by arrows, which indicates that the thermal
energy is sufcient to cause the melt of the shear layer [15e17].
Upon yielding at 298 K, serration events include the accumulation and release of the elastic energy, characterized by the stressascending and the sharp-stress-drop portion, respectively, as
shown in Fig. 2(a). The elastic-energy density (Dd) in a process of
the accumulation is [7]

1
2

Dd DsD3

(1)

where Ds and D3 are the elastic stress and elastic strain in one
serration event, respectively. The storage elastic energy, DEs, in the
sample with a volume of V is DEs Dd$V and will be dissipated
during shear banding, corresponding to the stress-drop process in
the serrations. The energy per unit shear surface, H, is approximately calculated to be:

aDEs
S

(2)

where a z 0.9 is the ratio of the storage elastic energy converted to


heat [18], and S is the surface area of the shear layer and has a value
of S 1/4pd2/cos q (d is the diameter of the sample, and q is the

!
 2
H
1
x
p pexp
4a0 t
t
2rCp pa0

(3)

where r is the density with a value of 6.82 g cm3,


Cp z 0.33 J g1 K1 at 298 K [19], and a0 is the thermal diffusivity
with a value of 2.2  106 m2 s1 at 298 K [20]. For Eq. (3), the
maximum DT at x 0 is highly dependent on t, and the shorter of t,
the larger of DT. Presently, it is assumed that the shear-banding
speed is the speed of sound with a value of 4.5  103 m s1 [16]. It
is noted that the shear-banding speed is considerably higher than
the shear-slip speed with a typical value of thousands of micrometers [21] along the shear plane. Zhang et al. [22] calculated that the
minimal shear offset was 280 nm for Vit.1 alloys as well as a minimal
measured value of about 1 mm upon bending. Yang et al. [17]
obtained a shear offset of about 300 nm for Vit.105 alloys upon
tension. In practice, the measured shear offset is a result of multistep shearing [5], and each serration contributes to the nallyobserved shear offset [8]. Besides, the viscous layer easily move
forward along the shearing under a small energy driven when
a trans-sample shear band forms, resulting in an increased shear
offset. Namely, not all the shear offsets need the mechanical work
when the shearing occurs along the existed shear layer. For
instance, Maa et al. [8] obtained the shear offset of about 2 mm for
one-cycle reloading after the generation of a fully-developed shear
band, which is about ten times of the reported values [17,22]. Here,
we take a shear offset of 200 nm as an estimate for one serration,
and, thus, the corresponding shear time of t z 0.04 ns. In the
present cases, there are 47 serrations in total. The value of H is in the
range of 1.4e61.2 J m2. We choose H 29.4 J m2 within the range
of obtained H values, and then DT 250 K using Eq. (3), together
with a resulting temperature, T 250 298 548 K z 0.8Tg, as
shown in Fig. 3(a). Generally, a viscous shear layer forms above 0.8Tg
for BMGs [1]. There are 11 of 47 serrations, which can lead to the
temperature higher than 0.8Tg based on the calculation through Eq.
(3). The maximum H is in accordance with the maximum
DT 520 K, as depicted in Fig. 4(a), and the resulting temperature is
lower than the melting temperature, Tm z 1100 K [23]. At the nal
fracture, the released strain energy is largely higher than that
during serrations [24], which is sufcient to cause the melting of the
shear layer.
At 77 K, when the serrations disappear, it is impossible to
evaluate the storage elastic energy based on the serrations. Here,
we specify the shear banding as a continuous process at 77 K, and
the thermal energy can be computed as H bsyd, where b 0.3
[25], sy is the yielding strength at 77 K, and d is the shear offset. Due
to the fracture separation of two parts along the primary fracture
surface, d is not determined, and simply estimated to be 1 mm, since
most BMGs have a shear offset of several micrometers upon plastic
deformation [22]. From the inset in Fig. 2(b), there is at least 8-step
shearing. Provided that N 8, the average shear offset be 125 nm,
and the maximum temperature rise, DT, during each shearing be
higher than 3000 K, as predicted in Fig. 3(b) (Cp z 0.17 J g1 K1
[19] and a0 z 1.0  106 m2 s1 at 77 K [20]). If the temperature is
largely higher than Tm, the viscous shear layer cannot sustain the
loading and, consequently, results in premature failure. Thus, N 8,
d is overrated at 77 K. If it is supposed with N 47, i.e., 47-step
shearing at 77 K similar to 47 serrations at 298 K, it is very interesting to nd that DT just produces a resulting temperature higher

78

J.W. Qiao et al. / Materials Chemistry and Physics 136 (2012) 75e79

temperature near Tm. At the nal fracture, instantly, all of the strain
energy dissipates, and it is sufcient to cause many droplets, in
accordance with the inset in Fig. 3(b). Noted that it is assumed the
shear offset to be 1 mm at 77 K, and if the real shear offset is larger
than this assumption, N would vary accordingly. The present study
just holds a clue to understanding the multi-step shear banding.
Fig. 5 presents the distribution of temperature rises spatially and
temporally at 298 and 77 K. It should be noted that here H chosen
from Eq. (3) can cause a resulting temperature just higher than
0.8Tg, i.e., H 29.4 J m2 at 298 K, and N 47 at 77 K. Instant
shearing would produce a very high temperature rise, and a rapiddecreasing DT dominates with the time (t) and width direction of
shear bands (x). Only if the viscous shear layer could sustain the
loading, the plastic deformation continues.
4. Conclusion
In summary, the analyses of the energy dissipation of shear
banding are very important to understand the plastic deformation
of BMGs. Multi-step shearing is proposed in this study to analyze
the energy dissipation. For the Zr55Al10Ni5Cu30 BMG at 298 K, the
thermal energy converted from the storage elastic energy during
serrations cannot result in DT higher than Tm, even though there are
few DTs higher than 0.8Tg. Only if the nal fracture is available, DT is
greatly higher than Tm, which causes that the shear layer cannot
sustain the loading. At 77 K, the serrations disappear, and an N-step
method is chosen to analyze the energy dissipation. When N is in
the range of 24e47, it is reasonable to deduce that the viscous shear
layer accommodates the plasticity. The present investigation gives
a proper method to reveal the shear banding for BMGs at different
temperatures. The more detailed investigation on the multi-step
shear banding experimentally and theoretically needs to be performed, which exceeds the scope of present studies.
Acknowledgments

Fig. 4. The proles of the temperature rise along the distance at (a) 298 and (b) 77 K.

than 0.8Tg, as shown in Fig. 4(b), which makes that the viscous
shear layer easily accommodates the plasticity. Here, it is articially
determined that the upper limit of DT Tm  77, as depicted in
Fig. 4(b). As a consequence, N 24 can lead to a resulting

J.W.Q. would like to acknowledge the nancial support of National


Natural Science Foundation of China (No. 51101110) and the nancial
supports from Key Laboratory of Cryogenics, TIPC, CAS (Grant No.
CRYO201107) and State Key Lab of Advanced Metals and Materials
(Grant No. 2011-Z06). P.K.L. appreciates the supports of National
Science Foundations (DMR-0909037, CMMI-0900271, and CMMI1100080) with Drs. A. Ardell and C.V. Cooper as program directors.
References

Fig. 5. The prole of the distribution of temperature rises spatially (x) and temporally
(t) at 298 and 77 K.

[1] C.A. Schuh, T.C. Hufnagel, U. Ramamurty, Acta Mater. 55 (2007) 4067e4109.
[2] Y. Zhang, A.L. Greer, Appl. Phys. Lett. 89 (2006) 071907.
[3] J.W. Qiao, H.L. Jia, C.P. Chuang, E.W. Huang, G.Y. Wang, P.K. Liaw, Y. Ren,
Y. Zhang, Scr. Mater. 63 (2010) 871e874.
[4] W.H. Jiang, G.J. Fan, F.X. Liu, G.Y. Wang, H. Choo, P.K. Liaw, Int. J. Plasticity 24
(2008) 1e16.
[5] S.X. Song, H. Bei, J. Wadsworth, T.G. Nieh, Intermetallics 16 (2008) 813e818.
[6] H.M. Chen, J.C. Huang, S.X. Song, T.G. Nieh, J.S.C. Jang, Appl. Phys. Lett. 94
(2009) 141914.
[7] J.W. Qiao, F.Q. Yang, G.Y. Wang, P.K. Liaw, Y. Zhang, Scr. Mater. 63 (2010)
1081e1084.
[8] R. Maa, D. Klaumnzer, J.F. Lfer, Acta Mater. 59 (2011) 3205e3213.
[9] H. Li, C. Fan, K. Tao, H. Choo, P.K. Liaw, Adv. Mater. 18 (2006) 752e754.
[10] Y.Q. Cheng, Z. Han, Y. Li, E. Ma, Phys. Rev. B 80 (2009) 134115.
[11] A. Inoue, T. Zhang, Mater. Trans. JIM 37 (1996) 185e187.
[12] G. Wang, K.C. Chan, L. Xia, P. Yu, J. Shen, W.H. Wang, Acta Mater. 57 (2009)
6146e6155.
[13] J.W. Qiao, P.K. Liaw, Y. Zhang, Scr. Mater. 64 (2011) 462e465.
[14] Z.F. Zhang, F.F. Wu, W. Gao, J. Tan, Z.G. Wang, M. Stoica, J. Das, J. Eckert,
B.L. Shen, A. Inoue, Appl. Phys. Lett. 89 (2006) 251917.
[15] J.J. Lewandowski, A.L. Greer, Nat. Mater. 5 (2006) 15e18.
[16] K. Georgarakis, M. Aljerf, Y. Li, A. LeMoulec, F. Charlot, A.R. Yavari,
K. Chornokhvostenko, E. Tabachnikova, G.A. Evangelakis, D.B. Miracle,
A.L. Greer, T. Zhang, Appl. Phys. Lett. 93 (2008) 031907.

J.W. Qiao et al. / Materials Chemistry and Physics 136 (2012) 75e79
[17] B. Yang, C.T. Liu, T.G. Nieh, M.L. Morrison, P.K. Liaw, R.A. Buchanan, J. Mater.
Res. 21 (2006) 915e922.
[18] J. Rosler, H. Harders, M. Baker, Mechanical Behavior of Engineering Materials,
Springer Berlin Heidelberg, New York, 2007.
[19] D. Okai, T. Fukami, T. Yamasaki, T. Zhang, A. Inoue, Mater. Sci. Eng. A 375-377
(2004) 364e367.
[20] M. Yamasaki, S. Kagao, Y. Kawamura, Scr. Mater. 53 (2005) 63e67.
[21] S.X. Song, X.-L. Wang, T.G. Nieh, Scr. Mater. 62 (2010) 847e850.

79

[22] Y. Zhang, N.A. Stelmashenko, Z.H. Barber, W.H. Wang, J.J. Lewandowski,
A.L. Greer, J. Mater. Res. 22 (2007) 419e427.
[23] X. Zheng, P. Shen, X. Han, Q. Lin, F. Liu, Y. Zhang, Q. Jiang, Mater. Chem. Phys.
117 (2009) 377e383.
[24] F.H. Dalla Torre, A. Dubach, J. Schllibaum, J.F. Lfer, Acta Mater. 56 (2008)
4635e4646.
[25] D.T.A. Matthews, V. Ocelk, P.M. Bronsveld, J.Th.M. De Hosson, Acta Mater. 56
(2008) 1762e1773.

Vous aimerez peut-être aussi