You are on page 1of 5

week ending

PRL 117, 155502 (2016) PHYSICAL REVIEW LETTERS 7 OCTOBER 2016

Controlling Strain Bursts and Avalanches at the Nano- to Micrometer Scale


Yinan Cui,* Giacomo Po, and Nasr Ghoniem
Mechanical and Aerospace Engineering Department, University of California, Los Angeles,
420 Westwood Plaza, Los Angeles, California 90095, USA
(Received 24 August 2016; revised manuscript received 15 September 2016; published 7 October 2016)
We demonstrate, through three-dimensional discrete dislocation dynamics simulations, that the complex
dynamical response of nano- and microcrystals to external constraints can be tuned. Under load rate
control, strain bursts are shown to exhibit scale-free avalanche statistics, similar to critical phenomena in
many physical systems. For the other extreme of displacement rate control, strain burst response transitions
to quasiperiodic oscillations, similar to stick-slip earthquakes. External load mode control is shown to
enable a qualitative transition in the complex collective dynamics of dislocations from self-organized
criticality to quasiperiodic oscillations.

DOI: 10.1103/PhysRevLett.117.155502

αE σ_
Power-law scaling of avalanche phenomena is widely σ_ ¼ ð_ε0 − ε_ p Þ þ 0 ; ð1Þ
observed in many nonequilibrium natural systems. 1þα 1þα
Examples are found in geologic earthquakes, snow ava- where α ¼ K p =K is the relative stiffness ratio, K ¼ EA=H
lanches, sandpile slides, and strain bursts during plastic is the pillar stiffness, E, A, and H are the Young module,
flow [1,2]. The realization that such vastly diverse the cross section area, and the height of the pillar,
physical systems display common features implies scale respectively. ε_ p is the plastic strain rate resulting from
invariance and compels a search into universal funda- all internal dislocation dynamical activities. Once the
mental laws. The common scaling raises the possibility stiffness ratio α is infinitely large—or σ_ 0 and ε_ 0 are very
that the intricate system behavior can be described by low—σ_ becomes very sensitive to ε_ p , implying that the
simple local rules, despite the complexity of the under- driving force changing rate (σ) _ is dominated by and
lying internal dynamics. One concept that is widely used comparable to its internal relaxation rate (_εp ). This
to interpret this universality is self-organized criticality indicates that the corresponding slip statistics are expected
(SOC) [3]. In a SOC system, the dynamics has an attractor to violate SOC.
characterized by infinite correlation time and length,
hence displaying scale-free scaling. A key hypothesis
(a) Open-loop Controller (b) σ. ε.
behind this abstraction is that the driving force varying 0 0

rate is much slower than the internal relaxation rate [3,4] Closed-loop Controller
Kp=αK
of a system undergoing SOC. Nevertheless, since this Proportional-Integral-
F0 (t ) Ff (t) Derivative (PID)
condition may not always hold, one wonders whether the .
ΔU e (t ) ε
qualitative aspects of a system’s dynamical behavior
change when the driving force changing rate is compa- U (t ) -
Actual +
rable to its internal relaxation rate. Our objective here is to displacement H K
investigate the relationship between the external driving H K
force and relaxation dynamics associated with strain Target
displacement
bursts during the nano- and microscale plastic deforma- U 0 (t)
tion of crystals.
Fixed
At the smallest of physical scales (e.g., nano- to micro-
scale), the release of plastic strain by intermittent “bursts”
FIG. 1. Simplified sketch of pillar compression. (a) Experimen-
has been found to belong to this power-law scaling tal setup with an open-loop (directly applying a force F0 ) and a
behavior [2,5–8]. One additionally unique aspect of plas- closed-loop control (to realize displacement control). (b) Simu-
ticity is that the driving force varying rate can be exper- lation setup, a proportional dominated closed-loop control is
imentally tailored. Considering a simple but illustrative considered here with Ff ¼ K p ðU 0 − UÞ, which is simplified as a
case, a pillar is subjected to uniaxial compression in Fig. 1. spring with a finite machine stiffness K p . The external stress rate
The force actuator, typically a voice coil, can exert an open- σ_ 0 ¼ F_ 0 =A, the target strain rate ε_ 0 ¼ U
_ 0 =H, and the actual strain
loop stress rate σ_ 0 and/or be controlled to impose a strain _
rate ε_ ¼ U=H, where A and H are the cross section area and the
rate ε_ 0 . For a proportional controller with the stiffness K p , height of the pillar, respectively. One typical dislocation con-
the internal stress rate in the pillar is [9] figuration in a pillar with d ¼ 3000b is shown as an example.

0031-9007=16=117(15)=155502(5) 155502-1 © 2016 American Physical Society


week ending
PRL 117, 155502 (2016) PHYSICAL REVIEW LETTERS 7 OCTOBER 2016

(a)
However, it is generally believed that the machine 10
0

stiffness K p only contributes to the cutoff of the power-


law scaling [6,8,15]. The present investigation demon-

CCDF C( U)
strates that, if the machine stiffness is extremely high, 10
-1

dislocation avalanche dynamics (and hence strain bursts) = ,d=1000b


= ,d=3000b
undergo a transition from scale-free critical behavior to =0,d=1000b
-2
quasiperiodic oscillations. Interestingly, this is consistent 10 =0,d=3000b
with recent findings on the role of very slow loading rates (a U-0.5

low σ_ 0 and ε_ 0 ) [16,17], as suggested by Eq. (1). The 10


-1
10
0
10
1

Burst displacement U (nm)


underlying microstructure mechanism for this dynamical (b)
5
× 10 × 10
4

regime transition is disclosed. Considering that the dynami- = , original data


smoothed results

Smoothed plastic strain rate


cal behaviors under soft and hard machine stiffness con- 2 2
ditions are vastly different, the corresponding intermittent

Plastic strain rate


plasticities will henceforth be described as avalanche and
burst, respectively. Moreover, a dislocation-based branch- 1 1
ing model is proposed, giving a clear and precise physical
picture of the avalanche dynamical behavior.
The vast majority of existing submicron mechanical 0 0
testing experiments can only cover a narrow range of 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Strain (%)
machine stiffness. In addition, the time necessary for (c)
5
× 10 × 10
5

dislocations to travel through a 1 μm sample is estimated =0, original data

Smoothed plastic strain rate


smoothed results
to be about 1 ns [18]. In state-of-the-art experiments, the 6
2

feedback loop frequency is ≈78 kHz (with a time constant

Plastic strain rate


Depinning
≈13 μs) [8], which means that the current experimental 4
transition

controller response rate is much slower than the sample 1


plastic relaxation rate by 4 orders of magnitude. Namely, 2
the driving force changing rate is much slower than the
internal relaxation rate. Therefore, most previous exper- 0 0
imental conditions correspond to the regime where SOC 0.2 0.4 0.6 0.8
Strain (%)
1 1.2 1.4

is observed. Discrete dislocation dynamics (DDD) studies,


as a computer simulation tool, make it possible to supple- FIG. 2. (a) Statistical properties of burst displacement under
ment experimental testing and explore regimes that are pure strain and stress control modes for pillar with diameters
currently difficult to access experimentally [6,19]. The d ¼ 1000b and 3000b. (b),(c) Typical evolution of plastic strain
current research presents the first systematic 3D-DDD rate and its averaged value in 0.24 μs windows, showing
investigation on the slip statistics at the submicron scale, (b) quasiperiodic strain bursts under pure strain control, and
accounting for the effects of the interaction of an external (c) a depinning transition dislocation avalanche under pure stress
control.
loading mode [20–22]. Compared to most of the existing
two-dimensional (2D) DDD studies [2,23], the key Figure 2(a) presents the results of statistical analysis of
approximations inherent in 2D techniques are resolved. the burst displacement magnitude ΔU. To obtain maximum
Specifically, dislocation junction formation and destruc- resolution of the limited simulation data set, the comple-
tion and the occurrence of cross slip are all accounted for mentary cumulative distribution function (CCDF) is used.
with minimal ad hoc assumptions. Figure 2(a) clearly illustrates that ΔU, under pure stress
The simulation setup is schematically shown in Fig. 1(b). control, exhibits a well-defined power-law distribution
We conducted simulations of compression tests on Cu spanning several orders of magnitude. The power-law
pillars of different diameters, ranging from 1000b to 3000b exponent for the corresponding probability density is found
(≈300 nm − 1 μm), where b is the Burgers vector magni- to be 1.5, agreeing well with the generally accepted range
tude. The aspect ratio, H=d, is 3. Two extreme machine of 1.35–1.67 [5,6,24–26]. In addition, the power-law
stiffness cases are first considered, corresponding to pure distribution is consistent across system size, implying
strain control (α ¼ þ∞) and pure stress control (α ¼ 0). the existence of scale-free universality. In contrast, the
Here, under pure strain control, the applied strain rate is CCDF of ΔU under pure strain control seems not to exhibit
ε_ 0 ¼ 960 s−1 . Correspondingly, under pure stress control, power-law scaling behavior for both small and large system
the actual loading rate σ_ 0 is E_ε0 . We carry out 50 and 20 sizes. Meanwhile, most of the data concentrate within one
separate simulations with different initial dislocation con- order of magnitude. An analogous breakdown of the
figurations under each loading mode, for d ¼ 1000b and power-law scaling under pure strain control is also
d ¼ 3000b, respectively. observed for the statistics of the burst duration [9].

155502-2
week ending
PRL 117, 155502 (2016) PHYSICAL REVIEW LETTERS 7 OCTOBER 2016

(a) 350
Then, how do we describe the strain burst statistics under
pure strain control? When discussing the temporal statistics 300

Stress (MPa)
of earthquakes, distinct dynamical behaviors are distin- 250
guished by the coefficient of variation C ¼ sx =x̄ [27], 200
where sx and x̄ are the standard deviation and the mean 150
value, respectively. For the cases of C > 1 and C < 1, the 100 =
distribution is referred to as “clustered” and “quasiperi- 50
=0
odic,” respectively; otherwise, if C ¼ 1, it is a random =0.5
Poisson distribution [27]. Taking the results of ΔU here, C 0 0.2 0.4 0.6 0.8
Strain (%)
is calculated as 1.9 and 0.9 under pure stress and pure strain
control, respectively. This suggests that the dynamical (b) = (c) =0.5 (d) =0
behaviors under pure strain control become quasiperiodic.
Similar to previous studies [16,27], quasiperiodicity here is
found to be stochastic due to the intrinsic scatter induced by
random cross slip or different dislocation configurations.
Quasiperiodic strain bursts under pure strain control are
manifested through the smoothed plastic strain rate, as FIG. 3. Typical simulation results under different loading
clearly shown in Fig. 2(b). Here, the time series of ε_ p is modes for a pillar with d ¼ 1000b. (a) Stress-strain curves.
smoothed over a fixed time window of 0.24 μs. For (b)–(d) Snapshots of dislocation configurations (from the top
comparison, the smoothed plastic strain rate under pure view) at a strain value of 0.4%. Arrows indicate the bowing out
stress control, also shown in Fig. 2(c), corresponds to a directions of activated sources.
depinning phase transition.
Close examination of dislocation configuration evolution
reveals that the mechanisms that control avalanche versus stress level keeps almost constant during each avalanche
quasiperiodic burst behavior are significantly different, and event [see Fig. 3(a)]. If one activated source leads to the
they are highly dependent on the external constraint. First, formation of a weaker one, the weaker source can be
let us consider pure strain control. In the submicron regime immediately activated. Thus, distinctly different from the
(e.g., d ¼ 1000b), each strain burst is found to be domi- strain control case discussed above, multiple sources can
nated by sequential activation and deactivation of single operate in a correlated fashion [see Fig. 3(d)]. All
arm dislocation sources. Once a source is activated, the correlated sources contribute then to an increasing mag-
accompanying plastic strain leads to a decrease in the stress nitude of the strain burst, turning it into an avalanche. Such
level [see Eq. (1), α ¼ þ∞]. Even if a weaker source is highly correlated dynamical behavior suggests a close-to-
formed during one burst event, sometimes it also cannot criticality nonequilibrium state [3].
operate due to the lower prevailing stress after relaxation. Since it is difficult to experimentally achieve such
This makes it difficult to trigger the simultaneous operation extreme machine stiffness, it is then interesting to examine
of multiple dislocation sources [see Fig. 3(b)], especially dislocation dynamics with finite machine stiffness. All of
for small samples with limited volume. We have recently the results in Fig. 3(a) correspond to the same size and
shown that dislocation sources themselves are transient initial dislocation configuration. The calculated stress-
because they generally result from the formation of dipolar strain curve with finite machine stiffness (α ¼ 0.5,
loops by cross slip [7]. This rapid stress drop prevents the σ_ 0 ¼ 0) in Fig. 3(a) displays a very similar behavior to
strain burst from continuously growing into a full-fledged past experimental results [8,26], and it exhibits a serrated
avalanche. Consequently, large-scale cooperative inter- yield character with longer decaying stages compared to
actions between dislocations that can lead to SOC cannot pure strain control. The observation of simultaneous
be realized under pure strain control. Note that this operation of multiple sources in Fig. 3(c) suggests that a
discussion applies to a sample size ranging from several finite machine stiffness actually promotes correlated dis-
nanometers to about 1 μm. For smaller pillars, surface location motion, compared to pure strain control.
nucleation of dislocations becomes dominant [28], and the To further elucidate the statistical difference between
rapid stress drop may inhibit correlated surface nucleation, avalanche versus quasiperiodic dynamics, a simple
while, for larger pillar size, Taylor-type interaction mech- dislocation-based branching model is proposed. It is
anisms prevail [29,30], and the rapid stress drop may inspired by the present 3D-DDD simulations and motivated
suppress cooperative dislocation interactions. by Zapperi’s sandpile branching model [31], in which we
By contrast, a dislocation avalanche under pure stress translate the branching idea into dislocation language. The
control is clearly associated with the correlated dislocation discrete plastic deformation is assumed to mainly proceed
motion. According to Eq. (1), when α ¼ 0, the stress rate through the intermittent activation of dislocation sources
cannot sense the internal dislocation activity. Thus, the [32,33]. One activated source may lead to the stochastic

155502-3
week ending
PRL 117, 155502 (2016) PHYSICAL REVIEW LETTERS 7 OCTOBER 2016
(a) Second branch Kth branch
generation or activation of other sources, similar to the All sources are deactivated,
First branch Burst event stops.
branching process shown in Fig. 4(a). If na=1 If na=2
New deactivated source,
The detailed algorithm proceeds as follows. Assuming a stored as potential source.
If na=2 This branch stops.
Current New activated
pillar initially with ns dislocation sources, we can randomly activated source If na=0
give each source a specific length λ according to a given source This branch stops.

source length probability distribution. The fate of each (b) 400 (c) 90
d=1000b, =
80
source (active or not) is determined by checking to see 350
70
DDD results

Probabiliy density (%)


Branching model
whether the instantaneous applied stress σ k can reach the
300
60

Stress (MPa)
250
50
source operation stress, 200
40
150 30
pffiffiffi 100 = ,Pure strain control 20
σ k M ≥ τ0 þ α1 μb ρ þ α2 μb=λ; ð2Þ 50
=0.5,finite machine stiffness
=0, pure stress control
10
0
0 0.2 0.4 0.6 0.8 1 2 3 4 5 6 7
Strain (%) Activated source number during each burst
where M is the Schmid factor, the three terms on the (d)
=0
(e)
=0,d=1000b
right-hand side are the lattice friction stress, the elastic =0.1

Probability density P( U)
=0,d=3000b

Probability density P( U)
=0.5 -2 U-1.5
interaction stress described by the Taylor relation, and the -2
10 =
10
U-1.5
source strength, respectively. α1 and α2 are dimensionless
constants, set to 0.5 and 1 [33], respectively. ρ is the
-4
-4
10
10 Pure
instantaneous dislocation density, estimated by dividing the strain
control
-6
10
total source length by the pillar volume. 10
0
10
1
10
2
10
0
10
1
10
2

Burst displacement U (nm) Burst displacement U (nm)


Once the weakest source is activated during deformation,
a strain burst begins [33,34]. After each source is activated, FIG. 4. (a) Schematic showing the random branching disloca-
the burst strain Sk increases by a specific value dεp . tion source generation and activation process. na is the number of
Considering that ε_ p is much higher than the applied strain newly generated dislocation sources, and the green filled circles
rate ε_ 0 during a strain burst, according to Eq. (1), σ k drops indicate that a new source is activated. Only an activated source
by Edεp α=ðα þ 1Þ, and the total strain increases by may trigger a further branching process. (b)–(d) Typical predicted
dεp =ðα þ 1Þ. It is assumed that the activated source is results for a pillar with d ¼ 1000b. (b) Stress-strain curve.
broken (ceases to operate) after it sweeps the entire slip (c) Comparison of the activated source number during each
plane once. However, this can randomly trigger the gen- burst under pure strain control. (d) Probability density function of
burst displacement for different machine stiffness. (e) Probability
eration of additional na sources. If the newly generated
density function of burst displacement for different sample sizes.
source can be activated according to Eq. (2), it triggers a
subsequent generation of na sources. Otherwise, the new
source is stored for possible dislocation generation, which as the nearest integer of 2 rand, where rand represents a
may activate during subsequent deformation stages. This random value from 0 to 1. Accordingly, the probabilities of
branching source generation process repeats itself until all na being 0, 1, and 2 are 25%, 50%, and 25%, respectively.
dislocation sources cannot be activated under the combined This is different from a previous sandpile branching model
effect of the instantaneous applied stress and the resistance [31], where the new activated site number was taken as a
stress, given by the right side of Eq. (2) [see Fig. 4(a)]. At constant value of 2. na ¼ 0 means that the source is
that instance, this strain burst event stops and the stress destroyed after operation once, na ¼ 1, 2 indicate that
continues to increase till it triggers another strain burst other sources are generated due to interactions with other
event. dislocations, cross slip, forming superjogs, or forming
In the following, we investigate the slip statistics dipolar loops [7]. Note that more deactivated sources
using this abstract branching model and compare it to may be left in the sample if na ¼ 2, leading to a slight
the more fundamental DDD simulations discussed above. increase in the dislocation density ρ after each branching
Compression tests are also modeled for Cu pillars with process. This results in an increase in the elastic interaction
diameters d ¼ 1000b and 3000b. Similar to the DDD resistance stress. Similar to 2D-DDD simulations [36], the
simulations, surface nucleation is not considered. If the source length is assumed to follow a Gaussian distribution,
stress is higher than the surface nucleation stress (about with a mean value λ̄ ¼ d=2, determined according to the
1.2 GPa for Cu [35]) or if the strain is higher than 0.5, yield stress of our DDD results. Its standard deviation is set
events are not recorded. If there is only one activated to 20%λ̄, so that the predicted activated source number for
source, each burst strain corresponds to the generated each strain burst event is statistically equivalent to those
plastic strain when the dislocation sweeps the entire slip obtained by our DDD results under pure strain control
plane once. Therefore, dεp is set to bM=H= cos β [33], [see Fig. 4(c)].
where β is the angle between the normal direction of the Figure 4(b) presents predicted typical stress-strain curves
slip plane and the loading orientation. Through examina- under different loading modes, which agree well with our
tion of the dislocation configuration evolution, na is taken simulation results in Fig. 3(a), including the stress level and

155502-4
week ending
PRL 117, 155502 (2016) PHYSICAL REVIEW LETTERS 7 OCTOBER 2016

the stepped or serrated burst features. In addition, the power- [9] See Supplemental Material at http://link.aps.org/
law scaling of burst displacement ΔU is also well repro- supplemental/10.1103/PhysRevLett.117.155502, which in-
duced under pure stress control for different pillar sizes in cludes Refs. [10–14], for machine stiffness simulation
Fig. 4(e). The power-law exponent of the probability method, and burst event identification and analysis.
[10] M. D. Uchic and D. M. Dimiduk, Mater. Sci. Eng. A
distribution of ΔU agrees with that obtained by the present
400–401, 268 (2005).
3D-DDD. Figure 4(d) clearly indicates that, as the machine [11] J. Weiss, T. Richeton, F. Louchet, F. Chmelik, P. Dobron, D.
stiffness increases, the power-law tails gradually become too Entemeyer, M. Lebyodkin, T. Lebedkina, C. Fressengeas,
wide to recognize proper scale-free power-law statistics. and R. J. McDonald, Phys. Rev. B 76, 224110 (2007).
The excellent agreement between the abstract branching [12] C. Fressengeas, A. J. Beaudoin, D. Entemeyer, T. Lebedkina,
model prediction and the fundamental 3D-DDD simula- M. Lebyodkin, and V. Taupin, Phys. Rev. B 79, 014108
tions further verify that hard machine stiffness leads to a (2009).
deviation from scale-free SOC because the rapid stress [13] D. M. Dimiduk, M. D. Uchic, S. I. Rao, C. Woodward, and
relaxation disturbs correlated dislocation motion. The T. A. Parthasarathy, Model. Simul. Mater. Sci. Eng. 15, 135
current finding offers a new pathway towards controlling (2007).
the correlated extent of dislocation dynamics and the [14] N. Friedman, A. T. Jennings, G. Tsekenis, J.-Y. Kim, M.
Tao, J. T. Uhl, J. R. Greer, and K. A. Dahmen, Phys. Rev.
intermittent statistics by tuning the machine stiffness. It
Lett. 109, 095507 (2012).
opens up new possibilities for novel experiments with a [15] M. Zaiser and N. Nikitas, J. Stat. Mech. (2007) P04013.
faster response rate that can reveal the quasiperiodic [16] S. Papanikolaou, D. M. Dimiduk, W. Choi, J. P. Sethna,
oscillation dynamics of dislocation systems. The impor- M. D. Uchic, C. F. Woodward, and S. Zapperi, Nature
tance of often-neglected interaction with the external (London) 490, 517 (2012).
loading system on intermittent plastic flow has been [17] K. A. Dahmen, Y. Ben-Zion, and J. T. Uhl, Nat. Phys. 7, 554
demonstrated. The complex dynamics of collective dis- (2011).
locations producing strain bursts is shown to be controlled [18] D. Dimiduk, E. M. Nadgorny, C. Woodward, M. D. Uchic,
through a simple tuning of the relative value of the driving and P. A. Shade, Philos. Mag. 90, 3621 (2010).
force rate to the internal relaxation rate. [19] J. A. El-Awady, Nat. Commun. 6, 5926 (2015).
[20] N. M. Ghoniem, S. H. Tong, and L. Z. Sun, Phys. Rev. B 61,
This material is based upon work supported by the 913 (2000).
U.S. Department of Energy, Office of Science, Office of [21] G. Po and N. Ghoniem, https://bitbucket.org/model/model/
Fusion Energy Sciences, under Award No. DE-FG02- wiki/home (2015).
03ER54708, and the U.S. Air Force Office of Scientific [22] G. Po, M. Lazar, D. Seif, and N. Ghoniem, J. Mech. Phys.
Research (AFOSR), under Grant No. FA9550-16-1-0444. Solids 68, 161 (2014).
[23] M. Zaiser, Adv. Phys. 55, 185 (2006).
We would like to thank Professor Michael Zaiser
[24] K. Ng and A. Ngan, Acta Mater. 56, 1712 (2008).
(Friedrich-Alexander University Erlangen-Nuremberg) [25] X. Zhang, B. Pan, and F. Shang, Europhys. Lett. 100, 16005
and Professor Stephanos Papanikolaou (Johns Hopkins (2012).
University) for the inspiring comments and discussions. [26] S. Brinckmann, J.-Y. Kim, and J. R. Greer, Phys. Rev. Lett.
100, 155502 (2008).
[27] Y. Ben-Zion, Rev. Geophys. 46 (2008).
*
cuiyinan@ucla.edu [28] J. R. Greer and W. D. Nix, Phys. Rev. B 73, 245410
[1] J. T. Uhl et al., Sci. Rep. 5, 16493 (2015). (2006).
[2] M. C. Miguel, A. Vespignani, S. Zapperi, J. Weiss, and [29] R. Gu and A. Ngan, Acta Mater. 60, 6102 (2012).
J.-R. Grasso, Nature (London) 410, 667 (2001). [30] R. Gu and A. Ngan, J. Mech. Phys. Solids 61, 1531 (2013).
[3] P. Bak, How Nature Works: The Science of Self-Organized [31] S. Zapperi, K. B. Lauritsen, and H. E. Stanley, Phys. Rev.
Criticality (Copernicus, New York, 1996). Lett. 75, 4071 (1995).
[4] P. Sammonds, Nat. Mater. 4, 425 (2005). [32] D. Kiener and A. Minor, Acta Mater. 59, 1328 (2011).
[5] D. M. Dimiduk, C. Woodward, R. LeSar, and M. D. Uchic, [33] Y. Cui, P. Lin, Z. Liu, and Z. Zhuang, Int. J. Plast. 55, 279
Science 312, 1188 (2006). (2014).
[6] F. F. Csikor, C. Motz, D. Weygand, M. Zaiser, and S. [34] J. A. El-Awady, M. Wen, and N. M. Ghoniem, J. Mech.
Zapperi, Science 318, 251 (2007). Phys. Solids 57, 32 (2009).
[7] T. Crosby, G. Po, C. Erel, and N. Ghoniem, Acta Mater. 89, [35] S.-W. Lee, A. T. Jennings, and J. R. Greer, Acta Mater. 61,
123 (2015). 1872 (2013).
[8] R. Maass, M. Wraith, J. T. Uhl, J. R. Greer, and K. A. [36] S. Papanikolaou, H. Song, and E. Van der Giessen (to be
Dahmen, Phys. Rev. E 91, 042403 (2015). published).

155502-5

You might also like