You are on page 1of 5

ARTICLES

Shock deformation of face-centred-cubic


metals on subnanosecond timescales
E. M. BRINGA1 *, K. ROSOLANKOVA2 , R. E. RUDD1 , B. A. REMINGTON1 , J. S. WARK2 , M. DUCHAINEAU1 ,
D. H. KALANTAR1 , J. HAWRELIAK1 AND J. BELAK1
1
Lawrence Livermore National Laboratory, Livermore, California 94551, USA
2
Department of Physics, Clarendon Laboratory, University of Oxford, Parks Road, Oxford OX1 3PU, UK
* e-mail: ebringa@llnl.gov

Published online: 17 September 2006; doi:10.1038/nmat1735

Despite its fundamental importance for a broad range hock compression of condensed matter occurs in a wide

of applications, little is understood about the behaviour


of metals during the initial phase of shock compression.
S variety of situations, including high-speed automobile and
aircraft collisions1 , explosive welding2 , armour penetration3,4 ,
meteor impacts5 , interstellar dust–dust collisions6 and inertial
confinement fusion7,8 . Despite considerable research during the
Here, we present molecular dynamics (MD) simulations past half century9–11 , a detailed understanding of the three-
dimensional (3D) lattice relaxation processes during shock
of shock-wave propagation through a metal allowing compression beyond the elastic limit does not yet exist. Much of
what we know has been inferred from recovery experiments12,13
a detailed analysis of the dynamics of high strain-rate and analysis of wave profiles14 . The microscopic lattice response
plasticity. Previous MD simulations have not seen the and relaxation behind the shock front has been measured in situ
using time-resolved X-ray diffraction15–21 : shocked lithium fluoride
evolution of the strain from one- to three-dimensional was observed to relax to the hydrostatic state within tens of
nanoseconds18 , whereas shocked single-crystal copper relaxed in
compression that is observed in diffraction experiments. ∼1 ns or less20 .
A detailed picture of the homogeneous nucleation of
Our large-scale MD simulations of up to 352 million atoms
dislocations as a function of shock strength emerges from large-
resolve this important discrepancy through a detailed scale non-equilibrium MD simulations of shock-wave propagation
in perfect face-centred-cubic (f.c.c.) crystals22 . Surprisingly, when
understanding of dislocation flow at high strain rates. X-ray diffraction patterns were calculated from these perfect crystal
simulations23 , the predicted lattice compression was essentially
The stress relaxes to an approximately hydrostatic state uniaxial (1D), whereas the diffraction experiments unambiguously
and the dislocation velocity drops to nearly zero. The showed a prompt plastic relaxation to a hydrostatic (3D) state20 .
Here, we resolve this dilemma by showing that MD simulations
dislocation velocity drop leads to a steady state with no of shock compression of copper single crystals with pre-existing
defects can achieve a nearly fully relaxed 3D hydrostatic state
further relaxation of the lattice, as revealed by simulated provided that the simulations are large enough to access the spatial
and temporal scales of the experiment. This study represents the
X-ray diffraction.
first direct, one-to-one comparison and validation of strong-shock
MD simulations with dynamic experimental data at the same
spatial and temporal scales, both at the lattice level and globally.
The combination of the MD simulation detail with experimental
validation offers a powerful tool with which to explore new regimes
in the emerging field of materials science at extreme conditions.
We have carried out several very-large-scale MD simulations
to compare with experimental data from time-resolved diffraction
experiments of shocked single-crystal copper (2-μm-thick sample,

nature materials VOL 5 OCTOBER 2006 www.nature.com/naturematerials 805

©2006 Nature Publishing Group


ARTICLES

~3D 1D and there is no observable contribution to the final simulated


Shocked, microstructure due to the initial defects. For the ramped case,
relaxed Shocked Unshocked shown in Fig. 2b, there is no dislocation production during the
1

Intensity (a.u.)
az initial 40 ps period of the ramp due to the large activation threshold
ay
ax ax a0 a0 of the sources embedded in the simulation. Once that threshold
a'z a0 a0
Piston
az a0 is reached, however, the pre-existing defects lead to dislocation
Up 0.1 multiplication and partial stress relaxation. As a result, in the
‘mixed’ region of Fig. 2b, the homogeneous nucleation front is
0 3 6
Lattice strain (%) not creating dislocations as prodigiously, and the final ρd , shown
Shock, Us in Fig. 2e, is roughly a third of that in the ‘pure’ homogeneous
z
nucleation region.
Dislocations: Motion Nucleation Elastic
This sequence suggests that there are several relevant rates
inherent to the material: the rate of increase in dislocation
density (and the closely related hardening rate), the rate of
Figure 1 Dynamic lattice compression during shock loading. The upper red
plastic relaxation and the ramp rate. How they compare with
curve shows the particle velocity versus time behind a shock front moving to the
each other determines the behaviour of the system. The rate of
right (z direction) at speed Us . The response of an initially cubic unit cell of side a0 ,
plastic relaxation can be described by the Orowan equation25 ,
is illustrated schematically by the blue blocks. The initial strain is uniaxial
dεp /dt = ρm vd b, where dεp /dt , ρm , vd and b are the plastic strain
compression (1D), with a z < a0 and a x = a y = a0 and large internal shear stress.
rate, the mobile dislocation density, average dislocation velocity
The lattice responds by relaxing via volume-preserving dislocation flow to a more
and the Burgers vector, respectively. During shock loading, d εp /dt
3D-symmetric compression, az ∼ a x ∼ a y . The inset shows experimentally
can be inferred from in situ X-ray diffraction and the Burgers vector
measured diffraction data with a peak corresponding to the unshocked material at
is known from the lattice. ρm and vd during shock loading are
zero strain and a peak corresponding to prompt (subnanosecond) 3D relaxation
largely unknown and no theory exists to predict how they evolve
(a z ∼ a x ) for shocked copper at 3%–4% strain20 .
under the rapid rise of shear stress in shock waves. These quantities,
however, can be obtained from the MD simulations. We find
that the strain relaxation is governed by the dislocation evolution
∼0.4 ns shock transit time). The simulations covered up to in two distinct regimes: initial unobstructed rapid dislocation
∼1-μm-thick copper samples, using up to 352 million atoms. multiplication, followed by decreases in both the mobile dislocation
Our MD-simulated samples had pre-existing dislocation sources density and velocity as obstructions accumulate. The time history
in the form of prismatic dislocation loops. As illustrated in of the relevant variables for a material element, calculated from the
Fig. 1, one surface of the system, the piston, was driven inwards MD simulation, is shown in Fig. 3. The 50 ps ramp and the sources
with a velocity Up , leading to a shock wave with velocity give an extra ∼1% lateral strain relaxation before homogeneous
Us > Up . Here, we describe two simulations in detail, both nucleation of dislocations begins, compared with the lateral strain
with Up = 0.75 km s−1 and a peak pressure of ∼35 GPa, which in the zero rise-time simulation. The ramp drive with dislocation
is above the homogeneous dislocation nucleation threshold for sources achieves greater relaxation with a lower total dislocation
copper. For the first simulation, Up was taken instantaneously from density. A longer ramp is expected to give a larger contribution to
zero to a constant value, launching a shock with essentially zero the lateral relaxation.
rise time. Previous MD simulations of shocked samples almost A fully relaxed 3D state would be characterized by roughly zero
exclusively used a zero rise-time pulse because they spanned no shear stress. The shear stress decreases behind the shock front22 ,
more than ∼10 ps in their entirety. We note that recent simulations as shown in Fig. 4a for both simulations. More than half of the
of shocks in nanocrystalline systems conducted by some of the relaxation occurs 10 ps after shock passage, and corresponds to the
authors of this paper have a comparable number of atoms, but a dislocation transport regime (as opposed to the dislocation-loop
much shorter timescale24 . For the second simulation, the piston nucleation regime). This regime is captured in MD simulations
velocity was increased in a linear ramp over 50 ps, and then held here for the first time. The shear stress evolves to an asymptotic
constant at the same peak Up as the zero rise-time simulation. value of 0.43 GPa for the zero rise-time simulation and 0.34 GPa for
This ramped velocity simulates experimental loading pulses which the ramped loading simulation. This relaxation is comparable to
have finite rise times, ranging from several picoseconds to several that inferred from the diffraction experiments on copper13 . Using
nanoseconds. In the discussions to follow, compression is calculated a smaller cross-section and a sample ∼2 μm long we find that
as (1 − a/a0 ), where a0 is the equilibrium lattice parameter and the stress indeed saturates in less than 100 ps: nothing is gained
a is the value determined from a diffraction analysis of both our by continuing the simulations to longer times (see Supplementary
post-processed MD simulations and our in situ dynamic diffraction Information, Figs S2,S3). When the shear stress relaxes to its
experiments on copper (Fig. 1, inset). asymptotic value, the dislocation density also reaches an asymptotic
A large density of dislocations, ρd , is created by the shock value (Fig. 2d,e). It is clear that the sharpness of the front, which is
compression in both simulations, as shown in Fig. 2a,b with controlled by the ramp rate, determines the extent of relaxation. In
snapshots at ∼100 ps from both the fast-rise-time simulation and the homogeneous nucleation regime, the dislocation multiplication
the 50 ps rise-time simulation (see Supplementary Information, rate is high, so the system quickly hardens to a high flow stress.
Movies S1,S2). Whereas there is only homogeneous nucleation of Below the homogeneous nucleation threshold, the ramped wave
dislocations for the 0 ps rise-time case (Fig. 2a), there are three can activate the sources and relax the shear at a much lower
regions of dislocation activity for the 50 ps rise-time case (Fig. 2b hardening rate. The result of this change in the relative rates is that
and Supplementary Information, Fig S1), with Fig. 2c showing the the ultimate relaxation for the ramped wave is greater.
particle speed profile for Fig. 2b. The total dislocation density (ρd ) Given that significant stress relaxation has occurred, as shown
versus depth is plotted at several times in Fig. 2d, for the zero in Fig. 4a, the simulated X-ray diffraction would be expected to
rise-time case. ρd saturates at a steady-state value of ∼3 × 1013 cm−2 give a nearly uniform 3D compression, meaning ∼5% in each
after ∼70 ps. The value of ρd at the homogeneous nucleation front, orthogonal direction for 15% volume compression. Some care is
ρmax , is very large compared with the density of pre-existing sources, needed in connecting the stress and strain relaxations, however,

806 nature materials VOL 5 OCTOBER 2006 www.nature.com/naturematerials

©2006 Nature Publishing Group


ARTICLES

a d 100

ρd (1012 cm–2)
Time (ps)
10 48
98
138
158
b 1
0.1 μm
Shock
0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
z (μm)

Homogeneous Mixed Multiplication e 100

ρd (1012 cm–2)
10
Time (ps)
c 62
Vz (km s–1)

0.8 1 72
82
0.4 100
143
0 0.1
0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6
z (μm) z (μm)

Figure 2 Dislocation structure resulting from MD simulations of shocked [100] copper. a, Snapshot of a simulation with a shock drive of 0 ps rise time after t ∼ 100 ps,
showing only dislocation atoms. b, The same as a, except that the shock drive had a 50 ps linear-ramp rise time. In both cases, the copper crystal included pre-existing
dislocation sources. The colour in a,b is only to enhance the view of the dislocations. The three regions of dislocation activity—homogeneous, mixed and multiplication—are
marked. c, The particle velocity, V z , profile for the ramped shock, where the ramp extends over z = 0.29–0.43 μm. d,e, The resulting dislocation density profiles for the 0 ps
rise-time case (d) and the ramped case (e). The locations of the pre-existing sources are illustrated by red vertical arrows.

because the shear modulus is known to soften under uniaxial pristine single crystal (ρ0 ∼ 106−8 cm−2 ) in this manner would give
compression (along the Bain path)26 . The lattice compression the same result (in terms of relaxation time, dislocation density
inferred from X-ray diffraction (Fig. 4b) at times <10 ps in the generated, VISAR (velocity interferometer system for any reflector)
simulation indicates little lateral relaxation. It is only at times traces, X-ray diffraction and so on) as shocking a pre-shocked or
much greater than 10 ps that the simulation shows significant radiation damaged crystal with ρ0 ≤ 10l−2 due to the dominance
lateral relaxation, a result that was first observed in these very of homogeneous nucleation for steep rise-time shocks. For a
long time simulations. For the zero rise-time case, the maximum 35 GPa shock in copper, 10l−2 ∼ 1011 cm−2 . Finally, measuring
transverse compression is 4.3 ± 0.1%, for a compression of the time evolution of the crystal from uniaxial compression to
7.1 ± 0.16% along the shock direction, where the error bars indicate a 3D relaxed state would provide a measure of the product of
standard deviation. This has to be compared with the 5.2% needed the mobile dislocation density and average dislocation velocity,
for isotropic compression. For the 50 ps rise-time shock with ρm vd , a determining quantity that has so far eluded measurement
pre-existing sources, the average final lattice compression along the techniques at ultrahigh strain rates. This rate of 3D relaxation is
x and y directions was found to be 4.7 ± 0.1%, whereas the final important for understanding the dynamics of the lattice response
lattice compression along z , the direction of shock propagation, to high strain-rate deformation, in particular, due to a shock. The
was 6.25 ± 0.25%, indicating nearly complete plastic relaxation. large-scale MD simulations shown here comprise an important step
The remaining small difference with the experiment may be due to in elucidating these dynamics.
the large source activation threshold in this simulation (∼10 GPa) In conclusion, we have shown, with very large-scale MD
and the fact that most of the ramp lies below this threshold and simulations at the correct spatial and temporal scales, the ability
does not produce any dislocation activity, as indicated in Fig. 3. For to reproduce the experimentally observed global dynamics and
comparison, the shear stress corresponding to this compression is microscopic lattice response of shocked copper; we expect our
[(2/3)1/2 C  {ε zz − (ε xx + ε yy )/2}], where C  = 24 GPa at ambient results to also apply to other f.c.c. metals. We show three
pressure, giving 0.55 and 0.30 GPa for the final shear stress in the distinct regimes of lattice relaxation: dislocation nucleation (t1
zero rise time and ramped waves, respectively, in good agreement in Fig. 3a), transport (t2 ) and immobilization due to tangling
with the values in Fig. 4a. In the early stages of the relaxation, the and interactions (t > t2 ). The fraction of dislocations that remain
softened shear modulus would need to be used, leading to a more mobile, or more precisely, ρm vd , is the dominant factor in
precipitous drop in the stress than in the strain. determining the 3D relaxation rate. Given the detailed insights into
On the basis of our simulations, we make two predictions for the lattice relaxation dynamics provided by our MD simulations,
experiments with shock rise times of only a few picoseconds. (1) We it seems possible that experiments could be developed on future
predict that prompt 3D relaxation would not be observed in copper, facilities, with measurement timescales ∼10 ps, to actually observe
provided that diffraction measurements could be made at time these distinct regimes of relaxation. Of particular importance
intervals of a few picoseconds behind the shock front, which is now would be the ability to deduce both the total dislocation density and
feasible experimentally27–29 . (2) For strong shocks, above the limit the fraction that is mobile, as a function of time. Such experiments,
for homogeneous nucleation of dislocations, with steep rises the combined with very-large-scale MD simulations, such as those
effect of pre-existing defects with a density ρ0 would be negligible, described here, would provide an inroad into the new frontier of
provided that ρ0 ≤ ∼10l−2 , where l is the mean separation time-resolved materials science at extreme pressures (>100 GPa)
between dislocations at the shock front. Therefore, shocking a and strain rates (>106 s−1 ).

nature materials VOL 5 OCTOBER 2006 www.nature.com/naturematerials 807

©2006 Nature Publishing Group


ARTICLES

a t2 t1 t0 a
1.0
2.0
0 ps
0.8
50 ps
Δ t2 Δ t1
ρm /ρmax
0.6 1.6
0.4

σ shear (GPa)
0.2 1.2

0
80 60 40 20 0 0.8
Time (ps)

b
0.3 0.4

0.2
vd /c 0

0
120 90 60 30 0
0.1 Time after shock (ps)

0 b
80 60 40 20 0 0 ps
Time (ps) 10 1D
Z; X–Y
c 0.04 50 ps
(dε /dt)/(ρmaxc0b)

Z; X–Y
8
0.02

Lattice compression (%)


0 6
80 60 40 20 0
Time (ps) 3D
d
6 4
ε (%)

2 2

0
80 60 40 20 0
Time (ps)
0
0.6 0.5 0.4 0.3 0.2 0.1 0
z (μm)
Figure 3 The dislocation dynamics of 3D relaxation in MD for the ramped shock
case. a, The evolution of mobile dislocation density, ρ m (t ), normalized by total
dislocation density, ρ max . The times indicated correspond to when dislocation Figure 4 Response of the MD lattice during 3D plastic relaxation. a,b, The
multiplication starts (t = 0 or t0 ), saturates (t 1 ), when the dislocations become relaxation of the shear stress (a) and the corresponding lattice compression (b) from
entangled and immobilized (t2 ), the dislocation nucleation period (t1 ) and the simulated X-ray diffraction. Depth and time behind the shock front (bottom and top
dislocation transport period (t2 ). b, Average dislocation velocity, vd , normalized by scales) apply to both panels. The shock front is located at z = 0, with the piston
bulk sound speed, c0 . c, The plastic strain rate from the MD simulation, normalized originally at z ∼ 0.55 μm. The black horizontal dashed line indicates the value for
by ρ max c0 b. d, Accumulated plastic strain, namely, the time integral of the result full 3D compression. The black vertical dot–dashed line indicates the extent of our
shown in panel c. previous simulations23 . In both panels, the red curves correspond to the 0 ps
rise-time shock case, and the blue curves to the ramped case. The curve for the
0 ps case includes diffraction for different ‘blocks’ of material at each z location. The
METHODS ‘jitter’ of the curve indicates the magnitude of the error in our simulations, which is
∼0.2% well behind the shock front. We note that current experiments21 can
MD SIMULATIONS
distinguish compression differences of 0.3%.
The MD code MDCASK was run using the EAM potential of Mishin et al.30 to
represent copper, with the longitudinal stress threshold for homogeneous
nucleation of dislocations σ zz ∼ 30 GPa. Pre-existing dislocation loops, with a nanometres (subpicosecond rise time), with a volumetric compression of
calculated threshold stress for source activation of σ zz ∼ 10 GPa, were 7 nm ∼15%. Shear stress is taken as [6−1/2 {σ zz − 0.5(σ xx + σ yy )}], as determined in
long and separated by 145 nm along the shock direction, for a lateral sample the region behind the shock front where it reached a spatially constant value
size of 72 nm. We generated shock waves along the [001] direction, using (except for a rise close to the ‘piston’).
200 × 200 × 2,200 and 200 × 200 × 1,600 f.c.c. unit cells for the simulation Dislocations were identified using a centro-symmetry deviation (CSD)
shown in Fig. 2a and b respectively, with periodic boundary conditions filter31 . Atoms in a CSD range 0.075–0.25 were assumed to be part of
transverse to the shock-wave propagation. The source size is inversely dislocation lines, allowing the calculation of dislocation densities. This method
proportional to its activation threshold and such large lateral size is required to only gives an upper limit of the dislocation density, and close comparison with
avoid interaction among the periodic source images and to allow independent small volumes leads to dislocation densities that were generally lower by a
source activation and evolution. The large length of the system is required for a factor 2. ρd for a shock with a steep rise time can be estimated from the mean
long-time shock propagation and to follow the long-term dislocation evolution. separation between dislocation loops l ∼ a0 /ε v (ref. 22), where the stacking
The system was equilibrated at a temperature of 5 K before loading. The shock fault spacing is equated with dislocation-loop separation, a0 is the lattice
front in the zero rise-time simulation develops a natural width of a few constant and ε v = Up /Us is the volumetric strain. For our 35 GPa shock,

808 nature materials VOL 5 OCTOBER 2006 www.nature.com/naturematerials

©2006 Nature Publishing Group


ARTICLES

ρmax ∼ l−2 ∼ 2 × 1013 cm−2 , which is consistent with our MD CSD analysis 10. Meyers, M. A. Mechanism for dislocation generation in shock-wave deformation. Scripta Met. 12,
(Fig. 2d). The dislocation density is also related to the flow stress in forest 21–26 (1978).
1/2 11. Meyers, M. A. et al. Constitutive description of dynamic deformation: physically-based mechanisms.
hardening, τ = αGbρd , where the prefactor α = 0.3 for copper. G and b are Mater. Sci. Eng. A 322, 194–216 (2002).
the shear modulus and Burgers vector for the compressed lattice. Using 12. Gray, G. T., Hayes, D. B. & Hixson, R. S. Influence of the shock-induced alpha-epsilon transition in Fe
ρd = 0.8 × 1013 cm−2 for the ramped case, and elastic moduli C11 = 168 GPa on its post-shock substructure evolution and mechanical behaviour. J. Phys. IV 10, 755–760 (2000).
13. Meyers, M. A. et al. Laser-induced shock compression of monocrystalline copper: characterization
and C12 = 121 GPa (the ambient values), we find τ = 0.39 GPa, in reasonable and analysis. Acta Mater. 51, 1211–1228 (2003).
agreement with the value from Fig. 4a. ρd is still 102 times larger than that 14. Vorthman, J. E. & Duvall, G. E. Dislocations in shocked and recovered LiF. J. Appl. Phys. 53,
found in recovery experiments13 . This discrepancy may be attributed to several 3607–3615 (1982).
15. Johnson, Q., Mitchell, A. & Evans, L. X-ray diffraction evidence for crystalline order and isotropic
causes not present in the MD shock simulation: (1) dislocation depletion from compression during the shock-wave process. Nature 231, 310–311 (1971).
tension waves due to the release at the free surfaces; (2) reduced dislocation 16. Whitlock, R. R. & Wark, J. S. Orthogonal strains and onset of plasticity in shocked LiF crystals. Phys.
production due to rise-time, dislocation-source and strain-rate effects; and Rev. B 52, 8–11 (1995).
17. d’Almeida, T. & Gupta, Y. M. Real-time X-ray diffraction measurements of the phase transition in
(3) dislocation depletion due to long-time (μs–ms) thermal annealing of KCl shocked along [100]. Phys. Rev. Lett. 85, 330–333 (2000).
recovered samples. For (1), we find in our simulations that ρd decreases by a 18. Rigg, P. A. & Gupta, Y. M. Multiple X-ray diffraction to determine transverse and longitudinal lattice
factor ∼3–5 within a few picoseconds of the shock hitting the back free surface, deformation in shocked lithium flouride. Phys. Rev. B 63, 094112–094123 (2001).
19. Zaretsky, E. Multipeak pulse X-ray diffraction study of shocked single crystals. J. Appl. Phys. 93,
and it is expected to decrease more at longer times after full release of the 2496–2506 (2003).
compressed state. 20. Loveridge-Smith, A. et al. Anomalous elastic response of silicon to uniaxial shock compression on
Mobile dislocation densities and dislocation velocities were calculated as nanosecond time scales. Phys. Rev. Lett. 86, 2349–2352 (2001).
21. Kalantar, D. H. et al. Direct observation of the α- transition in shock-compressed iron via
follows. Given all atoms in dislocation lines in a certain region, a random atom nanosecond X-ray diffraction. Phys. Rev. Lett. 95, 075502 (2005).
is chosen at time t , and the ‘persistence’ of that atom is checked in follow-up 22. Holian, B. L. & Lomdahl, P. S. Plasticity induced by shock waves in nonequilibrium
snapshots of the dislocation lines. If the atom remains in the dislocation for a molecular-dynamics simulations. Science 280, 2085–2088 (1998).
23. Rosolankova, K. et al. in Shock Compression of Condensed Matter-2003 (eds Furnish, M. D.,
time t , either the dislocation is fixed or it is moving at a velocity lower than Gupta, Y. M. & Forbes, J. W.) 1195–1198 (AIP, Melville, New York, 2004).
∼b/t . Repeating this procedure for a large number of dislocation atoms in a 24. Bringa, E. M. et al. Ultrahigh strength in nanocrystalline materials under shock loading. Science 309,
given slice at different time steps, and using a sufficiently long t , a reasonable 1838–1841 (2005).
25. Hirth, J. P. & Lothe, J. Theory of Dislocations (Wiley, New York, 1982).
estimate for the mobile dislocation density can be obtained. From the 26. Swift, D. C. & Ackland, G. J. Quantum mechanical predictions of nonscalar equations of state and
distribution of ‘transit times’, an approximate velocity distribution for mobile nonmonotonic elastic stress-strain relations. Appl. Phys. Lett. 83, 1151–1153 (2003).
dislocations can also be inferred. 27. Gahagan, K. T. et al. Measurement of shock wave rise times in metal thin films. Phys. Rev. Lett. 85,
3205–3208 (2000).
Simulated X-ray diffraction calculation allowed possible lattice rotations.
28. Rose-Petruck, C. et al. Picosecond-milliångström lattice dynamics measured by ultrafast X-ray
Corrections owing to large dislocation densities32 were evaluated using the diffraction. Nature 398, 310–312 (1999).
(200) and (400) peaks. Supplementary Information, Fig. S4 shows the 29. von der Linde, D & Sokolowski-Tinten, K. X-ray diffraction experiments with femtosecond time
resolution. J. Mod. Opt. 50, 683–694 (2003).
diffraction that would be recorded on film for the final lattice compression
30. Mishin, Y. et al. Structural stability and lattice defects in copper: Ab initio, tight-binding, and
in Fig. 4b. embedded-atom calculations. Phys. Rev. B 63, 224106–224121 (2001).
31. Kelchner, C. L., Plimpton, S. J. & Hamilton, J. C. Dislocation nucleation and defect structure during
Received 18 June 2006; accepted 10 August 2006; published 17 September 2006. surface indentation. Phys. Rev. B 58, 11085–11088 (1998).
32. Warren, B. E. X-ray Diffraction (Addison-Wesley, Menlo Park, California, 1969).
References
1. Itoh, M., Katayama, M. & Rainsberger, R. Computer simulation of a Boeing 747 passenger jet Acknowledgements
crashing into a reinforced concrete wall. Mater. Sci. Forum 465/466, 73–78 (2004).
The authors would like to thank B. Sadigh for developing a faster version of MDCASK and for help
2. Acarer, M., Gulenc, B. & Findik, F. Investigation of explosive welding parameters and their effects on
with the method to obtain mobile dislocation densities, G. Gilmer for help developing the code,
microhardness and shear strength. Mater. Design 24, 659–664 (2003). W. Cai and V. Bulatov and especially M. de Koning for help developing the initial dislocation sources,
3. Cheng, W. L. & Itoh, S. High velocity impact of steel fragment on thick aluminum target. Mater. Sci. and the MCR and Thunder computer teams for constant help and support. We would also like to
Forum 465/466, 49–54 (2004). thank P. Erhart, M. Meyers, M. Schneider, J. McNaney, J. Colvin, J. Stölken, M. J. Caturla, B. D. Wirth
4. Kumar, K. S., Singh, D. & Bhat, T. B. Studies on aluminum armour plates impacted by deformable and M. Kumar for fruitful discussions and J. Aranibar, A. Caro, and W. J. Nellis for useful comments.
and non-deformable projectiles. Mater. Sci. Forum 465/466, 79–84 (2004). For the LLNL authors, this work was carried out under the auspices of the US Department of Energy
5. Turtle, E. P. & Pierazzo, E. Thickness of a European ice shell from impact crater simulations. Science by the University of California, Lawrence Livermore National Laboratory under contract
294, 1326–1328 (2002). No. W-7405-Eng-48.
6. Slavin, J. D., Jones, A. P. & Tielens, A. G. G. M. Shock processing of large grains in the interstellar Correspondence and requests for materials should be addressed to E.M.B.
medium. Astrophys. J. 614, 796–806 (2004). Supplementary Information accompanies this paper on www.nature.com/naturematerials.
7. Haan, S. W. et al. Design and simulations of indirect drive ignition targets for NIF. Nuclear Fusion 44,
S171–S176 (2004).
8. Tokheim, R. E. et al. Hypervelocity shrapnel damage assessment in NIF target chamber. Int. J. Impact Competing financial interests
Eng. 23, 933–944 (1999). The authors declare that they have no competing financial interests.
9. Smith, C. S. Metallographic studies of metals after explosive shock. Trans. Metal. Soc. AIME 212,
574–589 (1958). Reprints and permission information is available online at http://npg.nature.com/reprintsandpermissions/

nature materials VOL 5 OCTOBER 2006 www.nature.com/naturematerials 809

©2006 Nature Publishing Group

You might also like