ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web...

9
Ductile-to-brittle transition in spallation of metallic glasses X. Huang, Z. Ling, and L. H. Dai Citation: Journal of Applied Physics 116, 143503 (2014); doi: 10.1063/1.4897552 View online: http://dx.doi.org/10.1063/1.4897552 View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/116/14?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Temperature-induced ductile-to-brittle transition of bulk metallic glasses Appl. Phys. Lett. 102, 171901 (2013); 10.1063/1.4803170 Core/shell structural transformation and brittle-to-ductile transition in nanowires Appl. Phys. Lett. 100, 153116 (2012); 10.1063/1.3703303 Temperature-induced anomalous brittle-to-ductile transition of bulk metallic glasses Appl. Phys. Lett. 99, 241907 (2011); 10.1063/1.3669508 Ductile to brittle transition in dynamic fracture of brittle bulk metallic glass J. Appl. Phys. 103, 093520 (2008); 10.1063/1.2912491 Electromigration induced ductile-to-brittle transition in lead-free solder joints Appl. Phys. Lett. 89, 141914 (2006); 10.1063/1.2358113 [This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP: 159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Upload: others

Post on 05-Jul-2020

26 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

Ductile-to-brittle transition in spallation of metallic glassesX. Huang, Z. Ling, and L. H. Dai Citation: Journal of Applied Physics 116, 143503 (2014); doi: 10.1063/1.4897552 View online: http://dx.doi.org/10.1063/1.4897552 View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/116/14?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Temperature-induced ductile-to-brittle transition of bulk metallic glasses Appl. Phys. Lett. 102, 171901 (2013); 10.1063/1.4803170 Core/shell structural transformation and brittle-to-ductile transition in nanowires Appl. Phys. Lett. 100, 153116 (2012); 10.1063/1.3703303 Temperature-induced anomalous brittle-to-ductile transition of bulk metallic glasses Appl. Phys. Lett. 99, 241907 (2011); 10.1063/1.3669508 Ductile to brittle transition in dynamic fracture of brittle bulk metallic glass J. Appl. Phys. 103, 093520 (2008); 10.1063/1.2912491 Electromigration induced ductile-to-brittle transition in lead-free solder joints Appl. Phys. Lett. 89, 141914 (2006); 10.1063/1.2358113

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Page 2: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

Ductile-to-brittle transition in spallation of metallic glasses

X. Huang,1,2 Z. Ling,1 and L. H. Dai1,3,a)

1State Key Laboratory of Nonlinear Mechanics, Institute of Mechanics, Chinese Academy of Sciences,Beijing 100190, China2Institute of Systems Engineering, China Academy of Engineering Physics, Mianyang, Sichuan 621999, China3State Key Laboratory of Explosion Science and Technology, Beijing Institute of Technology, Beijing 10081,China

(Received 31 July 2014; accepted 29 September 2014; published online 9 October 2014)

In this paper, the spallation behavior of a binary metallic glass Cu50Zr50 is investigated with

molecular dynamics simulations. With increasing the impact velocity, micro-voids induced by ten-

sile pulses become smaller and more concentrated. The phenomenon suggests a ductile-to-brittle

transition during the spallation process. Further investigation indicates that the transition is con-

trolled by the interaction between void nucleation and growth, which can be regarded as a competi-

tion between tension transformation zones (TTZs) and shear transformation zones (STZs) at atomic

scale. As impact velocities become higher, the stress amplitude and temperature rise in the spall

region increase and micro-structures of the material become more unstable. Therefore, TTZs are

prone to activation in metallic glasses, leading to a brittle behavior during the spallation process.VC 2014 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4897552]

I. INTRODUCTION

Due to the unique disordered microstructures, metallic

glasses (MGs) have many excellent properties and receive

much attention in recent years.1–8 It is well known that me-

tallic glasses usually exhibit a brittle behavior like a glass at

macroscopic scale, but show different capability of plastic

deformation at microscopic scale.9–13 Thus, two distinct

morphologies are usually observed on the fracture surfaces

of MGs. For brittle fracture, the fracture surfaces are flat

with nano-scale periodic corrugations or dimple struc-

tures;13–15 but for ductile fracture (not globally), much

coarser patterns are found, such as river-like and cellular pat-

terns as well as honeycomb structures.16,17

The fracture behavior of MGs is sensitive to their com-

position, and Mg-based and Fe-based MGs are usually much

more brittle than Zr-based MGs.13,18,19 The fabrication pro-

cess is also important. The longer the annealing time is, the

more brittle MGs are.9 Besides, different loading conditions

may lead to various fracture behaviors. During plate-impact

experiments, Gupta and coworkers20,21 found that spallation

of a Zr-based MG exhibits a ductile-to-brittle transition.

With increasing the impact velocity, the pull-back velocity

slope increases monotonically, which indicates that the

loading-unloading response of the MG at macroscopic scale

is more brittle. Further examination show that the spalled

surfaces at microscopic scale agree with the macroscopic

phenomenon.21,22 Smoother morphologies are observed at a

higher impact velocity, while much coarser patterns are

observed at a lower impact velocity.

To answer the question of what controls the ductile-to-

brittle behavior in MGs, extensive works have been made

over the past decades. On one hand, some researchers tried

to find the macroscopic mechanical parameters that dominate

the ductile-to-brittle transition process. In 1975, Chen et al.23

found that Poisson’s ratio is closely correlated with plasticity

of MGs. Schroers and Johnson24 further proved that the

larger the Poisson’s ratio, the better is the plasticity of MGs.

Equivalent with Poisson’s ratio, another parameter l=jrevealed by Lewandowski et al.25 is a key parameter control-

ling the ductile-to-brittle transition of MGs, where l is the

shear modulus representing the resistance to plastic deforma-

tion, and j is the bulk modulus or the resistance to dilation.

A lower l=j or larger Poisson ratio implies more ductile

behavior. It is noted that MGs usually exhibit a significant

tension-compression plasticity asymmetry and shear-induced

dilation. Considering these intrinsic characters, Chen

et al.10,11 recently took the intrinsic strength of the material

into consideration, and proposed a shear-to-normal strength

ratio a and a strength-differential factor b to characterize the

ductile-to-brittle behavior in MGs. A smaller a implies

enhanced plasticity, while a larger b indicates brittle fracture

under tensile loading. On the other hand, researchers

intended to find the answer at atomic scale. Based on a over-

view of fracture patterns, Jiang et al.14 argued that the duc-

tile-to-brittle transition of MGs is controlled by competition

between shear transformation zones (STZs)26–28 and tension

transformation zones (TTZs)7,9,14,21,29 at microscopic scale.

In contrast to STZs that are corresponding to shape distor-

tions of atomic clusters under shear stresses, TTZs are

regarded as the fundamental carriers of bulk dilations under

negative pressures.7,14 When TTZs dominate, more brittle

facture behavior is expected. This view is supported by

recent impact toughness tests21 and spallation experiments.9

In these tests, typical brittle fracture patterns are observed

such as nm-sized vein patterns21 and nanosized corruga-

tions,9 and TTZs are thought to be the reason for the phe-

nomenon. More recently, Murali et al.30,31 studied the

fracture behavior of two typical MGs (FeP and CuZr) via

a)Author to whom correspondence should be addressed. Electronic mail:

[email protected]

0021-8979/2014/116(14)/143503/8/$30.00 VC 2014 AIP Publishing LLC116, 143503-1

JOURNAL OF APPLIED PHYSICS 116, 143503 (2014)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Page 3: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

atomistic simulations. It is revealed that even a brittle frac-

ture is dominated by nucleation and growth of voids in MGs,

and a higher degree of spatial fluctuation induces more brittle

behavior during the fracture process. Despite extensive

investigations, the atomistic scale mechanism that governs

the ductile-to-brittle transition in MGs is still unclear.

To reveal the ductile-to-brittle transition mechanism

during a spallation process, we present molecular-dynamic

(MD) simulations of a binary MG Zr50Cu50 in this paper. By

using a flyer-target configuration, the spallation behavior is

studied at different impact velocities from 600 m/s to

1800 m/s, with emphasis on the damage evolution process. It

is found that as the impact stress increases, a ductile-to-brit-

tle transition occurs, which agrees well with the available ex-

perimental results. Further investigation reveals that the

interaction between void nucleation and growth, which can

be interpreted as the competition between TTZs and STZs at

atomic scale, controls the ductile-to-brittle transition during

the spallation process.

II. MD SIMULATIONS OF SPALLATION

During the MD simulations, a simple binary MG

Zr50Cu50 is selected as the model material. To model the

atomic interactions in the Zr-Cu system, we adopt the

Finnis-Sinclair type interatomic potential with parameters

given by Mendelev et al.32 Calculations are carried out with

the open source code LAMMPS.33 Glass samples are pre-

pared via a melting-and-quenching process. The initial sys-

tem is a fcc lattice with the sites randomly occupied by Zr

and Cu atoms in accordance with the nominal composition.

It consists of �440 000 atoms arranged in a cubic shape, and

three-dimensional periodic boundary conditions with ambi-

ent pressure are applied. To obtain the Zr50Cu50 glass, simu-

lations are performed in the constant number of particles,

pressure, and temperature (NPT) ensembles with a time step

of 1 fs. Temperature gradually increases from 1 K to 2500 K,

equilibrates for 100 ps and cools down to 300 K, with the

same heating and cooling rate of 5 K/ps. After a further

relaxation for 100 ps, a glass sample is prepared with dimen-

sions of �20� 20� 20 nm3.

In simulations of spallation, we construct the traditional

flyer-target configurations.34,35 The flyer plate consists of

�2 200 000 atoms with dimensions of �100� 20� 20 nm3,

and the target has the same cross-section area (20� 20 nm2)

but its thickness is twice as that of flyer. To obtain such a

large system, the 400 000-atom glass (�20� 20� 20 nm3) is

replicated along the X direction, and equilibrates for another

100 ps to remove possible artifacts from the replication pro-

cess.34 In fact, we have also explored the flyer-target system

with a cross-section area of �10� 10 nm2 to examine the

size effect on spallation and the results are similar. In our

simulations, the loading direction is along the X axis, so the

nonimpact sides of flyer and target normal to the X axis are

free surfaces. But along the Y and Z axes, the periodic

boundary conditions are maintained to mimic one-

dimensional (1D) strain shock loading. Here, we denote the

desired impact velocity as V. The flyer plate and target are

assigned initial velocities of 2V=3 and �V=3 before

impacting, so that the flyer-target system has a center-of-

mass velocity of 0. Shock simulations adopt the constant

number of particles, volume, and energy (NVE) ensembles.

The time step for integrating the equations of motion is 1 fs,

and the run duration is 120 ps.

To obtain the physical properties of plates, the 1D bin-

ning analysis is used. The simulation cell is divided into fine

bins along the X axis by neglecting the heterogeneities in the

transverse directions, and we obtain the average physical

properties such as density (q), stresses (rx), particle velocity

(up), and temperature (T) profiles. To characterize the atomic

configuration, we use the Voronoi tessellation analysis.36

And the plastic deformation is identified by the nonaffine

displacement D2min proposed by Falk and Langer.27

III. RESULTS

During the shock simulations, the thickness of flyer

plates and targets are not changed. To achieve shock loading

with different amplitudes, we choose impact velocities V of

600, 900, 1200, 1500, and 1800 m/s, respectively. Figure 1

illustrates the free surface velocity histories on the target

side, similar to that measured by a velocity interferometer

system for any reflector (VISAR) in plate-impact experi-

ments.20,37,38 As shown in Fig. 1, typical “pull-back” waves,

which are signatures for spallation, are observed in all cases,

except for the case of V¼ 600 m/s. It indicates that spallation

occurs in the cases of V¼ 900, 1200, 1500, and 1800 m/s.

Besides, as the impact velocity increases, the pull-back ve-

locity slope also increases. It agrees well with the experi-

mental results,20 which indicates a ductile-to-brittle

transition behavior. To compare the spallation behaviors

under different loading amplitudes, the cases of V¼ 900 and

1500 m/s are further characterized.

Figure 2 shows the density evolution in a conventional

x - t diagram at impact velocities of 900 and 1500 m/s. With

color coding based on the local atomic number density, the

wave propagation and interaction process is illustrated,

which is related to the shock, release, tension, and spallation.

As shown in Fig. 2, the red color represents regions with a

higher density, while the blue color represents regions with a

lower density. A deeper blue color implies a larger amount

FIG. 1. Free surface velocity histories on the target side at different impact

velocities.

143503-2 Huang, Ling, and Dai J. Appl. Phys. 116, 143503 (2014)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Page 4: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

of spallation damage. According to the process of wave

propagation, the tensile stress duration needed for spallation

at the lower impact velocity is much longer than that at the

higher velocity. It is obvious that the distribution of spalla-

tion damage is different between these two impact velocities.

In the case of V¼ 900 m/s, damage is scattered over the spall

plane. But in the case of V¼ 1500 m/s, it is more

concentrated.

The corresponding stress profiles (rx) at different times

before and after spallation at V¼ 900 and 1500 m/s are

shown in Fig. 3. It is seen that the tensile region is formed

due to the interaction of two release waves reflected from the

free surfaces of flyer and target. As micro-damage nucleates

and grows, recompression waves are generated in the spalled

region and propagate toward the free surface. The recom-

pression wave is registered in the free surface velocity profile

as a “pull-back” wave. Compared with the case of

V¼ 1500 m/s where there is only one recompression wave,

two recompression waves are observed near the spall plane

at V¼ 900 m/s, which imply that a multi-spall occurs. The

result is in accordance with the scattered distribution of spal-

lation damage at a lower impact velocity, as shown in Fig. 2.

Next, we examine the damage evolution process in the

spalled region (where the recompression wave is generated)

at different impact velocities. Figures 4(a)–4(c) show the

spallation damage at the impact velocity of 900 m/s, and

Figs. 4(d)–4(f) show the damage at V¼ 1500 m/s. As the

impact velocity varies, the rate of damage evolution is differ-

ent. Thus, in the case of 900 m/s, a time spacing of 5 ps is

used to track the spall process, while 3 ps is adopted at

V¼ 1500 m/s. As shown in Fig. 4, spallation of Cu50Zr50

glass undergoes the process of nucleation, growth, and coa-

lescence of micro-voids. At a lower impact velocity, only a

few large voids (actually only one in the slice) dominate the

damage evolution process. In contrast, a large number of

voids can be observed at a higher impact velocity. The voids

are small and begin to coalesce. The damage characteristics

imply a smoother morphology on the fracture surfaces at a

higher impact velocity.

IV. DUCTILE-TO-BRITTLE TRANSITION MECHANISM

According to the results of plate-impact experiments,20–22

there are two typical characteristics at different impact veloc-

ities, which suggests a ductile-to-brittle transition during spalla-

tion of MGs: (1) at macroscopic scale, the pull-back velocity

slope increases with increasing the impact velocity; and (2) at

microscopic scale, it is frequent to observe a smoother morphol-

ogy on the fracture surfaces of the spalled samples at a higher

impact velocity, while much coarser patterns are observed at a

lower impact velocity. Our results generally agree with the ex-

perimental results,20–22 as shown in Figs. 1 and 3.

In the MD simulations, the most obvious difference

between the fracture phenomena at different impact veloc-

ities is the change of generated void numbers. With increas-

ing the impact velocity, there are much more voids observed

to nucleate and grow on the spall plane. The larger the void

number is, the smaller the void sizes are before coalescence.

Then the fracture surface is smoother, which is a typical

characteristic in brittle fracture. This interesting phenomenon

has also been observed in other works. For example, during

the MD simulations of the fracture behavior of two typical

MGs (FeP and CuZr),31 more smaller voids are observed in

brittle FeP MG, while one bigger void is found in ductile

CuZr MG. As the fracture behavior (brittle or ductile) is

determined by the plastic deformation at microscopic scale,

this phenomenon implies that the plastic deformation is

impeded when more voids are generated. Now the question

FIG. 2. The x – t diagram for shock

loading of Cu50Zr50: (a) V¼ 900 m/s;

and (b) V¼ 1500 m/s.

FIG. 3. The stress profiles at different

time: (a) V¼ 900 m/s; and (b)

V¼ 1500 m/s.

143503-3 Huang, Ling, and Dai J. Appl. Phys. 116, 143503 (2014)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Page 5: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

is, why the plastic deformation is impeded in the case with

more and smaller voids.

A. Competition of TTZs and STZs

In order to reveal the factors that influence the plastic

deformation during spallation, we explore the process of

void nucleation and growth. Figure 5 shows the nonaffine

displacement during the void nucleation and growth at

V¼ 900 m/s. Here, the critical size for void nucleation is

determined to be �1 nm in diameter.22,39 Thus, according to

the void size, Figs. 5(a)–5(c) illustrate the nucleation pro-

cess, and Figs. 5(d)–5(e) exhibit the growth process. As

shown in Figs. 5(a)–5(c), during the void nucleation process,

atoms with a larger nonaffine displacement are randomly dis-

tributed in the material. With increasing the time interval

(the reference configuration is the same at t¼ 81 ps in Fig.

5), the number of atoms with a larger nonaffine displacement

increases. There is no apparent difference observed between

the void nucleation location and other region. It implies that

the nonaffine displacement is induced by temperature (or

structural relaxation) instead of stresses. However, during

the void growth process, nonaffine displacement of atoms in

the region around the void is much larger than that away

from the void. It indicates that plastic deformation of the ma-

terial is mainly induced by void growth, there is nearly no

contribution from void nucleation.

FIG. 4. Damage evolution process at

different impact velocities: (a)–(c)

V¼ 900 m/s; and (d)–(f) V¼ 1500 m/s.

The colors indicate the normalized

local atomic number density.

FIG. 5. Snapshots of void nucleation

and growth at V¼ 900 m/s: (a)–(c)

Nucleation of voids; and (d)–(f)

Growth of voids. The colors represent

the value of D2min calculated from the

same reference configuration at t¼ 81

ps.

143503-4 Huang, Ling, and Dai J. Appl. Phys. 116, 143503 (2014)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Page 6: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

Further investigation reveals that the nucleation and

growth of voids is closely related to the fundamental unit-

processes of collective atomic motion in MGs. Figure 6

shows some close-up views of the atomic cluster motion

around the void. As shown in Fig. 6(a), during the nucleation

process, transformation of the atomic structure at the centre

of the void is similar to the picture of a TTZ.14 But during

the growth process as shown in Fig. 6(b), the motion of the

atomic cluster at the edge of the void is close to the picture

of a STZ.26 We know that TTZs are corresponding to bulk

dilations of atomic clusters, but STZs arouse shape distor-

tions (the accompanied dilations are very small). As STZs

are mainly activated during the void growth process, plastic

deformation induced by damage evolution in the material is

attributed to the void growth process.

Now the question is, as the impact velocities increase,

why does plastic deformation decrease? Because plastic de-

formation is closely related to the void growth process, void

growth at different impact velocities is examined. We com-

pare the diameter history of the biggest voids at V¼ 900 and

1500 m/s, as shown in Fig. 7. At a lower impact velocity, we

see that the void grows continuously with a gradually

increasing growth rate. But at a higher impact velocity, the

void grows fast at the initial stage, but the growth rate

decrease a lot after a short time of �4 ps. The difference can

be explained by the damage evolution process as shown in

Fig. 4. As there are more voids at the higher impact velocity,

a growing void quickly interacts with the surrounding voids,

leading to a coalescence process. This impedes the further

growth of voids. However, as there is only one void at the

lower impact velocity, it can grow continuously without con-

finements of other voids. Based on the above results, we

think that the plastic deformation during spallation of MGs

is controlled by competition of two rate processes at micro-

scopic scale. On one hand, the void growth process promotes

plastic deformation in the material. According to the conven-

tional void growth mechanism,40–42 the plastic zone around

the void is proportional to the void volume. Bigger voids

induce a larger region of the material to undergo plastic de-

formation. Thus, the larger the voids grow, the more exten-

sive plastic deformation the material undergoes. On the other

hand, the void nucleation process impedes plastic deforma-

tion in the material. As void growth is bounded by the spac-

ing between two nucleation sites, a higher nucleation rate

which decreases the spacing between voids impedes the

growth process. Therefore, plastic deformation in the mate-

rial is slight.

In fact, the interaction between nucleation and growth

can be interpreted as a competition between the fundamental

unit-processes of collective atomic motion in MGs. Since

void nucleation is related to the activation of TTZs, and

growth is induced by STZs around the voids, the damage

evolution process is intrinsically a competition between

FIG. 6. Motion of atomic clusters dur-

ing the damage evolution process at

V¼ 900 m/s: (a) nucleation; and (b)

growth.

FIG. 7. Diameter of the biggest void at different impact velocities.

143503-5 Huang, Ling, and Dai J. Appl. Phys. 116, 143503 (2014)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Page 7: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

TTZs and STZs. To characterize the competition process,

here we propose a non-dimensional number composed of

two time scales:

Ia ¼TSTZ

TTTZ; (1)

where TSTZ and TTTZ are the characteristic time scales for

activation of STZs and TTZs, respectively. As to a larger Ia,

TTZs are more dominant than STZs. To determine the two

time scales, we estimated the activation rates of TTZs and

STZs.

According to the STZ models,27,43 the activation rate of

a single potential STZ is written as

vSTZ ¼1

TSTZ¼ v1 exp �DF1 � s � c0 � X0

kh

� �; (2)

where vSTZ is the STZ activation rate, v1 is an attempt fre-

quency of order of the Debye frequency, s is the local shear

stress, c0 is the characteristic shear strain with the order of

�0.1, X0 is the STZ volume, k is the Boltzmann constant, his the temperature, and DF1 is the activation barrier.

For TTZs, they are similar in size to STZs, and are acti-

vated by high hydrostatic tensile pressure. In the same way,

we can estimate the activation rate of a single TTZ as

vTTZ ¼1

TTTZ¼ v2 exp �DF2 � p � ev � X0

kh

� �; (3)

where vTTZ is the TTZ activation rate, v2 is an attempt fre-

quency, p is the hydrostatic tensile pressure, ev is the charac-

teristic volumetric strain, and DF2 is the activation barrier,

which is mainly related to the dissipated energy forming new

surfaces.14 Thus,

Ia ¼v1

v2

expDF1 � s � c0 � X0

DF2 � p � ev � X0

� �; (4)

where v1, DF1, c0, X0 are material parameters according to

STZ models. If v2, DF2, and ev are also regarded as material

parameters, Ia are determined by local stress states.

Further analysis indicates that the local stress states

change before and after voids are nucleated. When there is no

void in the material during spallation, it is the 1D strain condi-

tion and the ratio of shear stress s to tensile pressure p is

sp¼ l

j; (5)

where l is the shear modulus and j is the bulk modulus.

For Cu50Zr50, l � 22 GPa and j � 123 GPa, therefore,

s=p � 0:18. Since the shear stress is much smaller than the

tensile pressure, TTZs may play a dominant role according

to Eq. (4). However, after voids are nucleated, the local

stress states are completely changed. Although the tensile

pressure p is nearly the same, the ratio of s to p around the

void increases to 0.75 (as a rough estimate, the asymmetry of

the loading and initial void shape is not taken into account).

Thus, Ia will decrease, and STZs may play a dominant role.

Note that only around the void’s surrounding where stress

concentration takes place, Ia is smaller. For the region that is

not influenced by the void, Ia is still relatively large and

TTZs is the dominant collective atomic motion.

B. Mechanism resulting in dominance of TTZs

If MGs undergo brittle spallation, it is obvious that

TTZs must dominate the fracture process. According to Eq.

(3), factors such as stresses, temperature, and the activation

barrier can influence the activation of TTZs in the material.

In order to find the reason that results in dominance of TTZs,

we further compare the evolution of the above factors at dif-

ferent impact velocities.

Figure 8 shows a comparison of the stress profiles at the

beginning of the damage evolution process. With increasing

the impact velocity, the stress amplitude near the spall plane

is slightly higher. According to Eq. (3), a higher tensile stress

can increase the work done by the system, and decreases the

energy barrier of TTZs, thus a higher activation rate is

expected. Besides, it should be noted that micro-inertia

might have influence on the competition between STZs and

TTZs. As the impact velocity increases, the loading rate is

higher and micro inertial effects on void growth become

more important. Activation of STZs around the voids may be

impeded by micro inertial effects, leading to a decrease of

void growth rate.

The history of material temperature near the spall plane

is illustrated in Fig. 9(a). The temperature keeps constant at

first, then increases sharply as the flyer impacts the target,

and finally decreases a little when the region of tension is

created. Compared with the case with V¼ 900 m/s, the mate-

rial temperature is apparently higher at V¼ 1500 m/s. As

higher temperature implies that atoms have a higher chance

of getting enough energy from thermal fluctuation to over-

come the free energy barrier, it contributes to a higher activa-

tion rate.

For the activation barrier, it is determined by local

atomic structures at the potential TTZ sites. Here, the degree

of local fivefold symmetry (LFFS) is used as a key factor to

characterize the local atomic structures.44 In the Voronoi tes-

sellation analysis, each atom is indexed with the Voronoi

indices hn3; n4; n5; n6; :::i, where n3, n4, n5, and n6 represent

FIG. 8. Stress profiles at the beginning of the damage evolution process: (a)

V¼ 900 m/s; and (b) V¼ 1500 m/s.

143503-6 Huang, Ling, and Dai J. Appl. Phys. 116, 143503 (2014)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Page 8: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

the number of triangles, tetragons, pentagons, and hexagons

on the Voronoi polyhedron, respectively, and the degree of

LFFS is defined as the fraction of the number of pentagons

(LFFS ¼ n5=P

ini). The average degree of LFFS in the

region near the spall plane is shown in Fig. 9(b). It is clear

that the average degree of LFFS increases under compres-

sion and decreases under tension at both impact velocities.

But at the time just before voids begin to nucleate, the aver-

age degree of LFFS can decrease to lower amplitude at the

higher impact velocity. Since a lower LFFS indicates that

the structural configuration of atoms is packed more loosely

and has higher potential energy, the local structure is more

unstable and it is easier for transformation of local atomic

clusters.

Further examination indicates that TTZs are prone to

activation from the region with a lower average LFFS. As

shown in Fig. 10, at the impact velocity of 900 m/s, the aver-

age degree of LFFS of the atomic cluster that void originates

from is 0.422, while that of the entire slice near the spall

plane is 0.472. As the impact velocity increases to 1500 m/s,

the average degree of LFFS of the atomic cluster that the

biggest void originates from (0.431) is also smaller than that

of the entire slice (0.464). It indicates that TTZs are indeed

easier to be activated in the region with a lower degree of

LFFS. As a previous work has shown that STZs prefer to be

initiated in regions with a lower degree of LFFS too, it is

obvious that a lower degree of LFFS means a lower

activation barrier for transformation of atomic clusters. As

the average degree of LFFS in the spall region is smaller at a

higher impact velocity, there are more potential sites for acti-

vation of TTZs. Thus, the activation rate is higher.

V. CONCLUSION

We have studied the ductile-to-brittle transition phe-

nomenon during spallation of a binary MG Zr50Cu50 with

MD simulations. Our results show that as the impact velocity

increases, the distribution of spallation damage becomes

more concentrated and the fracture patterns are smoother,

which agrees well with experimental observations in recent

works. The ductile-to-brittle transition in spallation is related

to extra fracture energy dissipation at a lower impact veloc-

ity and impedance of plastic deformation at a higher impact

velocity. Plastic deformation during the damage evolution

process is controlled by the interaction of two microscopic

rate processes (i.e., void nucleation and growth), which can

be interpreted as the competition of STZs and TTZs at

atomic scale. As the impact velocity increases, TTZs domi-

nates the fracture process and spallation exhibits a brittle

behavior. Further investigation shows that with increasing

the impact velocity, the tensile stress amplitude and material

temperature is higher in the spall region, and the atomic

structure is more unstable. All these reasons induce a larger

void nucleation rate or the dominance of TTZs.

FIG. 9. The history of material temper-

ature and average LFFS near the spall

plane at different impact velocities of

900 m/s and 1500 m/s: (a) temperature

and (b) LFFS.

FIG. 10. Atomic configuration before

void nucleation showing the LFFS of

atoms. The colors represent value of

LFFS at (a) V¼ 900 m/s and (b)

V¼ 1500 m/s.

143503-7 Huang, Ling, and Dai J. Appl. Phys. 116, 143503 (2014)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06

Page 9: Ductile-to-brittle transition in spallation of metallic ...faculty.imech.ac.cn/dailanhong/web files/paper/Huang-jap2014.pdf · Ductile-to-brittle transition in spallation of metallic

ACKNOWLEDGMENTS

Financial support was from the National Key Basic

Research Program of China (2012CB937500), the NSFC

(Grants Nos.: 11272328, 11472287, and 11402245), and the

CAS/SAFEA International Partnership Program for Creative

Research Teams.

1M. W. Chen, Annu. Rev. Mater. Res. 38, 445–469 (2008).2A. L. Greer and E. Ma, MRS Bull. 32(8), 611–615 (2007).3A. Inoue, Acta Mater. 48(1), 279–306 (2000).4H. Li, C. Fan, K. Tao, H. Choo, and P. K. Liaw, Adv. Mater. 18(6),

752–754 (2006).5W. H. Wang, Prog. Mater. Sci. 57(3), 487–656 (2012).6Z. P. Lu, C. Liu, J. Thompson, and W. Porter, Phys. Rev. Lett. 92(24), 245503

(2004).7M. M. Trexler and N. N. Thadhani, Prog. Mater. Sci. 55(8), 759–839

(2010).8X. J. Liu, Y. Xu, X. Hui, Z. P. Lu, F. Li, G. L. Chen, J. Lu, and C. T. Liu,

Phys. Rev. Lett. 105(15), 155501 (2010).9R. Raghavan, P. Murali, and U. Ramamurty, Acta Mater. 57(11),

3332–3340 (2009).10Y. Chen and L. H. Dai, Sci. Sin.-Phys. Mech. Astron. 42(6), 551 (2012).11Y. Chen, M. Q. Jiang, Y. J. Wei, and L. H. Dai, Philos. Mag. 91(36),

4536–4554 (2011).12M. Martin, L. Kecskesb, and N. N. Thadhani, Scr. Mater. 59, 688–691

(2008).13X. K. Xi, D. Q. Zhao, M. X. Pan, W. H. Wang, Y. Wu, and J. J.

Lewandowski, Phys. Rev. Lett. 94, 125510 (2005).14M. Q. Jiang, Z. Ling, J. X. Meng, and L. H. Dai, Philos. Mag. 88(3),

407–426 (2008).15J. X. Meng, Z. Ling, M. Q. Jiang, H. S. Zhang, and L. H. Dai, Appl. Phys.

Lett. 92, 171909 (2008).16E. Bouchaud, D. Boivin, J. L. Pouchou, D. Bonamy, B. Poon, and G.

Ravichandran, Europhys. Lett. 83(6), 66006 (2008).17C. T. Liu, L. Heatherly, D. S. Easton, C. A. Carmichael, J. H. Schneibel,

C. H. Chen, J. L. Wright, M. H. Yoo, J. A. Horton, and A. Inoue, Metall.

Mater. Trans. A 29(7), 1811–1820 (1998).18P. Lowhaphandu, L. A. Ludrosky, S. L. Montgomery, and J. J.

Lewandowski, Intermetallics 8, 487–492 (2000).19Y. Wu, H. X. Li, Z. B. Jiao, J. E. Gao, and Z. P. Lu, Philos. Mag. Lett.

90(6), 403–412 (2010).

20S. J. Turneaure, S. K. Dwivedi, and Y. M. Gupta, J. Appl. Phys. 101,

043514 (2007).21J. P. Escobedo and Y. M. Gupta, J. Appl. Phys. 107(12), 123502 (2010).22X. Huang, Z. Ling, H. S. Zhang, J. Ma, and L. H. Dai, J. Appl. Phys.

110(10), 103519 (2011).23H. S. Chen, J. T. Krause, and E. Coleman, J. Non-Cryst. Solids 18(2),

157–171 (1975).24J. Schroers and W. L. Johnson, Phys. Rev. Lett. 93(25), 255506 (2004).25J. J. Lewandowski, W. H. Wang, and A. L. Greer, Philos. Mag. Lett.

85(2), 77–87 (2005).26A. S. Argon, Acta Metall. 27(1), 47–58 (1979).27M. L. Falk and J. S. Langer, Phys. Rev. E 57(6), 7192–7205 (1998).28D. Pan, A. Inoue, T. Sakurai, and M. W. Chen, Proc. Natl. Acad. Sci.

U.S.A. 105(39), 14769–14772 (2008).29M. Q. Jiang, J. X. Meng, J. B. Gao, X. L. Wang, T. Rouxel, V. Keryvin, Z.

Ling, and L. H. Dai, Intermetallics 18(12), 2468–2471 (2010).30P. Murali, Y. W. Zhang, and H. J. Gao, Appl. Phys. Lett. 100(20), 201901

(2012).31P. Murali, T. Guo, Y. Zhang, R. Narasimhan, Y. Li, and H. Gao, Phys.

Rev. Lett. 107(21), 215501 (2011).32M. I. Mendelev, D. J. Sordelet, and M. J. Kramer, J. Appl. Phys. 102(4),

043501 (2007).33LAMMPS: MD simulation code available from Sandi National

Laboratories, USA.34B. Arman, S.-N. Luo, T. C. Germann, and T. Ca�gın, Phys. Rev. B 81(14),

144201 (2010).35S.-N. Luo, T. C. Germann, and D. L. Tonks, J. Appl. Phys. 107(5), 056102

(2010).36N. N. Medvedev, J. Comput. Phys. 67(1), 223–229 (1986).37S. A. Atroshenko, N. F. Morozov, W. Zheng, Y. J. Huang, Y. V.

Sudenkov, N. S. Naumova, and J. Shen, J. Alloys Compd. 505(2),

501–504 (2010).38S. M. Zhuang, J. Lu, and G. Ravichandran, Appl. Phys. Lett. 80(24),

4522–4524 (2002).39D. R. Curran, L. Seaman, and D. A. Shockey, Phys. Rep. 147(5–6),

253–388 (1987).40E. Bouchbinder, T. S. Lo, and I. Procaccia, Phys. Rev. E 77(2), 025101

(2008).41X. Y. Wu, K. T. Ramesh, and T. W. Wright, J. Mech. Phys. Solids 51(1),

1–26 (2003).42X. Huang, Z. Ling, and L. H. Dai, Int. J. Solids Struct. 50(9), 1364–1372

(2013).43E. R. Homer and C. A. Schuh, Acta Mater. 57(9), 2823–2833 (2009).44H. L. Peng, M. Z. Li, and W. H. Wang, Phys. Rev. Lett. 106(13), 135503

(2011).

143503-8 Huang, Ling, and Dai J. Appl. Phys. 116, 143503 (2014)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

159.226.199.222 On: Tue, 14 Oct 2014 06:36:06