scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Defective twin boundaries in nanotwinned metals

01 Aug 2013-Nature Materials (Nature Research)-Vol. 12, Iss: 8, pp 697-702
TL;DR: It is demonstrated that as-grown CTBs in nanotwinned copper are inherently defective with kink-like steps and curvature, and that these imperfections consist of incoherent segments and partial dislocations.
Abstract: Coherent twin boundaries (CTBs) are widely described, both theoretically and experimentally, as perfect interfaces that play a significant role in a variety of materials. Although the ability of CTBs in strengthening, maintaining the ductility and minimizing the electron scattering is well documented, most of our understanding of the origin of these properties relies on perfect-interface assumptions. Here we report experiments and simulations demonstrating that as-grown CTBs in nanotwinned copper are inherently defective with kink-like steps and curvature, and that these imperfections consist of incoherent segments and partial dislocations. We further show that these defects play a crucial role in the deformation mechanisms and mechanical behaviour of nanotwinned copper. Our findings offer a view of the structure of CTBs that is largely different from that in the literature, and underscore the significance of imperfections in nanotwin-strengthened materials.
Figures (5)

Content maybe subject to copyright    Report

LETTERS
PUBLISHED ONLINE: 19 MAY 2013 | DOI: 10.1038/NMAT3646
Defective twin boundaries in nanotwinned metals
Y. Morris Wang
1
*
, Frederic Sansoz
2
, Thomas LaGrange
1
, Ryan T. Ott
3
, Jaime Marian
1
,
Troy W. Barbee Jr
1
and Alex V. Hamza
1
Coherent twin boundaries (CTBs) are widely described, both
theoretically and experimentally, as perfect interfaces that
play a significant role in a variety of materials. Although
the ability of CTBs in strengthening, maintaining the ductility
and minimizing the electron scattering is well documented
1–3
,
most of our understanding of the origin of these properties
relies on perfect-interface assumptions. Here we report
experiments and simulations demonstrating that as-grown
CTBs in nanotwinned copper are inherently defective with
kink-like steps and curvature, and that these imperfections
consist of incoherent segments and partial dislocations.
We further show that these defects play a crucial role in
the deformation mechanisms and mechanical behaviour of
nanotwinned copper. Our findings offer a view of the structure
of CTBs that is largely different from that in the literature
2,4,5
,
and underscore the significance of imperfections in nanotwin-
strengthened materials.
CTBs formed during growth, deformation or annealing exist
broadly in many crystalline solids with low or medium stacking-
fault energies
1,5,6
. The strengthening behaviour and other attractive
properties of CTBs have been studied in nanotwinned metals (with
an average twin spacing <100 nm; refs 7–9). One prevalent view
is that CTB-strengthened materials have certain advantages over
nanocrystalline or ultrafine-grained materials; that is, materials
strengthened through traditional grain boundaries (GBs) that
are considered incoherent and defective
10
. GBs not only scatter
electrons, but can migrate and slide under shear stresses
11
, leading to
a maximum in strength in nanocrystalline materials
12,13
. In contrast,
such migration/sliding mechanisms may not be operative in CTBs
despite some reports of detwinning evidence
7,14,15
and the obser-
vation of a similar maximum strength in a nanotwinned copper
3
(nt-Cu). Existing models widely assume perfect CTBs and rational-
ize flow softening due to CTB migrations and detwinning as caused
by nucleation and motion of partial dislocations parallel to CTBs
(ref. 4). These mechanisms are informative as long as CTB lengths
are limited to the tens of nanometres typically used in molecular
dynamics simulations
4,16–18
. It still remains difficult through molec-
ular dynamics simulations to validate the migrations/detwinning of
the much longer CTBs seen in experiments (500 nm; ref. 3). There
could be alternative mechanisms that are intricately related to the
potential structures of CTBs and the characteristics of GBs, both of
which are not accounted for in the literature.
Recent studies of nanotwinned copper pillars without GBs
revealed strong deformation anisotropy and a brittle-to-ductile
transition behaviour (where CTBs are considered intrinsically
brittle)
2
, suggesting that CTBs alone are not sufficient for increased
plasticity despite their strong strengthening effect, and that a
reasonable mix of GBs is helpful to mediate the plasticity and
achieve high ductility. Experiments and simulations have frequently
1
Physical and Life Sciences Directorate, Lawrence Livermore National Laboratory, Livermore, California 94550, USA,
2
School of Engineering, The University
of Vermont, Burlington, Vermont 05405, USA,
3
Division of Materials Sciences and Engineering, Ames Laboratory (USDOE), Ames, Iowa 50011, USA.
*e-mail: ymwang@llnl.gov
quoted GBs as one primary source of dislocations in nanotwinned
polycrystalline materials, the nature of which is expected to
have impacts on the mechanical behaviour; that is, experimental
evidence hints at the relevance of GBs in nanotwinned metals.
Unfortunately, limited information is available regarding the nature
of GBs and the roles they are playing in controlling the plasticity.
We have examined the nanometre structural features of many
twin boundaries (TBs) and the characteristics of GBs using a re-
cently developed inverse pole figure orientation mapping (IPFOM)
in a field-emission transmission electron microscope (TEM). This
nanodiffraction-based technique has a spatial resolution of 1 nm
that is comparable to the conical-scanning dark-field imaging
approach
19
. In a standard IPFOM image, 250,000 frames of
diffraction patterns were collected, each of which was indexed by
EDAX orientation-image-mapping indexing software and applied
to determine the crystallographic orientations. The orientation-
resolving ability of IPFOM allows us to identify many previously
unobserved microstructural defect features on preconceived-to-be
completely perfect TBs. The materials used in our studies are nt-Cu
synthesized by magnetron sputtering (see Methods and Supple-
mentary Information). Shown in Fig. 1a is the cross-sectional mi-
crostructure of an as-deposited material, indicating a twin spacing
of 5–70 nm with grain column width spanning from 0.3 to
1.8 µm. Most TBs appear rather straight without obvious evidence
of defective structures. Similar materials have been synthesized in
parallel by other groups albeit with clear variations in twin density
and columnar grain size
14,20
. To obtain statistical misorientations
of TBs and GBs, we carried out detailed IPFOM studies in TEM
using the NanoMegas Astar system
21
(the pixellation effect of this
technique can be seen in Supplementary Figs S1 and S2). As shown
in Fig. 1b, under low resolution the as-synthesized nt-Cu contains
at least two types of TB; that is, 63{111} CTBs normal or inclined
to the growth direction that often transverse the columns, and a
very limited number of 63{112} incoherent TBs (ITBs) that tend to
bond CTBs terminated inside grains (marked with circles). Owing
to their low frequency, the influence of these ITBs on the mechanical
behaviour is considered inconsequential until the twins become
very thin (<2 nm; ref. 17). However, a zoomed-in high-resolution
IPFOM image shown in Fig. 1c indicates that originally thought
perfect 63{111} CTBs contain many kink-like steps along the
interfaces (a few are marked with arrows). These steps are inco-
herent 63{11
¯
2} segments, as evidenced by their crystallographic
orientations, and their heights (1–5 nm) are visibly smaller than
the ITBs mentioned above. The CTBs observed in IPFOM analysis
are imperfect and comprise numerous partial dislocations, as com-
pared with those with perfectly straight lines in two-dimensional
images. For simplicity, we hereinafter refer to these steps as kinks in
two-dimensional descriptions. Note that these nanometre features
are substantially larger than the Burgers vector length (0.147 nm) of
NATURE MATERIALS | VOL 12 | AUGUST 2013 | www.nature.com/naturematerials 697
© 2013 Macmillan Publishers Limited. All rights reserved

LETTERS
NATURE MATERIALS DOI: 10.1038/NMAT3646
ac d
350 nm
CTB
CTB
CTB
CTB
CTB
CTB
CTB
CTB
CTB
CTB
CTB
GD
ITB
ITB
ITB
001 011
112
111
b
350 nm
40 nm
5 µm
100 nm
Figure 1 | Microstructure of as-grown nanotwinned copper. a, A focused-ion-beam ion-channelling image of the as-grown nanotwinned copper. The
imaging condition is 30 keV, 9.7 pA. The orientations of representative TBs are marked with double-headed red arrows. b, An IPFOM image of growth
twins. The coherent and incoherent TBs are labelled as CTB and ITB (inside circles), respectively. The growth direction (GD) of the film is marked with an
arrow. c, An edge-on high-resolution IPFOM image of CTBs. Some selected kinks are marked with white arrows. The portion of CTBs marked with an X is
perfect TBs without defects. Note that the TEM sample for IPFOM is prepared by electropolishing. d, An IPFOM image of a long GB (from top to the
bottom), which shows mixed segments of low- and high-angle GBs. The GBs are characterized by a coincidence site lattice (CSL) model. The interface
colour scheme in bd is as follows: yellow, low-angle GBs with misorientation < 15
; black, large-angle GBs; red, 63 CSL boundary; blue, 69 CSL boundary.
The orientations of grains are coloured according to the inverse (001) pole figure shown in the inset of b.
a single Shockley partial previously reported on CTBs (ref. 3); but
the existence of these defective structures is not at odds with these
earlier high-resolution TEM studies. A close examination of Fig. 1c
suggests that the kink distributions are relatively non-uniform from
TB to TB but that their density is rather persistent; that is, kink
density per unit area increases with twin density. For the sample
shown in Fig. 1c, there exists an average of one kink per 10-nm-long
CTB (Supplementary Fig. S3). These findings strongly suggest that
as with conventional GBs, many CTBs have defective structures
(that is, kinks or incoherent components) that have not been
considered in previous studies. The formation mechanisms of these
kinks are not well understood; however, they seem to be the intrinsic
characteristics of growth twins. It is conceivable that these features
are related to processing or deposition conditions, and are expected
to impact mechanical properties.
Another IPFOM image shown in Fig. 1d reveals that GBs in
nt-Cu are more defective and full of partial-dislocation-like defects,
as manifested by abundant kinks along the GBs. Interestingly, the
statistical misorientation angle distribution maps obtained from
both cross-section (Supplementary Fig. S4) and plan-view indicate
an appreciable fraction of low-angle GBs (misorientation <15
),
leading to a GB network distribution that is far away from a
MacKenzie distribution
22
—the characteristic GB distribution of
a well-annealed coarse-grained copper. Note that the curvy and
defective nature of CTBs remains visible in this image. The GB dis-
tributions observed here strongly suggest that the classical Voronoi
construction approach for GB networks in past modelling studies
may not be applicable to nanotwinned materials reported here.
To help reveal the strain-dependent deformation mechanisms
and the deformation anisotropy (copper has an anisotropic factor
(A) of 3.21; A = 2C
44
/(C
11
C
12
), where C
11
,C
12
and C
44
are elastic
constants of copper)), we performed in situ synchrotron X-ray
diffraction (SXRD) experiments at two different temperatures to
track in real-time the elastic and plastic deformation of nt-Cu. In
addition to offering a better statistical approach for bulk materials
23
,
the in situ SXRD has the advantageous capability of resolving
anisotropic microstrains
24
. Six crystallographic indices (that is,
111, 200, 220, 311, 222, 400) were followed simultaneously in
these in situ SXRD experiments. To investigate the microplas-
ticity, we quantitatively calculate the lattice strain deviation as
hkl
= ε
hkl
σ /E
hkl
, where ε
hkl
= (d
hkl
d
hkl
o
)/d
hkl
o
(d
hkl
o
and d
hkl
are the inter-planar spacing before and during loading, respec-
tively), σ is the applied stress and E
hkl
is the measured average
elastic modulus for a specific plane hkl (see Supplementary Table
S1). Figure 2a,b shows the
hkl
behaviour parallel to the tensile
axis (that is, longitudinal direction) for four representative Miller
indices tested at room temperature and 180
C (see Supplementary
Fig. S6 for the heating profile), respectively; the
hkl
behaviour
for the transverse direction can be seen in Supplementary Fig. S7.
For clarity, the corresponding tensile stress–strain curves are in-
cluded, which show softening behaviour at both temperatures.
In coarse-grained face-centred-cubic copper, the deformation is
governed by full dislocation slips of 1/2h110i{111} type, leading
to preferential yielding and negative of 220 reflection at the
onset of macroscopic plasticity, whereas 200 exhibits the largest
positive (ref. 25). Interestingly, the trend in nt-Cu shows a
rather different behaviour, where 200 shows a pronounced negative
before 0.2% yield stress, and sharply curls back to the positive
territory on further loading. The 111 reflection exhibits similar
curling behaviour; but to a smaller degree. By comparison, the
of 220 remains positive during the deformation. The initial
negative deviation of for both 200 and 111 reflections in
nt-Cu is an indication of preferable slip activations next to h001i
and h111i orientations, suggestive of 1/6h112i{111} partial related
698 NATURE MATERIALS | VOL 12 | AUGUST 2013 | www.nature.com/naturematerials
© 2013 Macmillan Publishers Limited. All rights reserved

NATURE MATERIALS DOI: 10.1038/NMAT3646
LETTERS
0
50
100
150
200
250
300
350
¬600
¬400
¬200
0
200
400
600
0.00 0.01 0.02 0.03 0.04
111
200
220
311
Engineering stress (MPa)
Engineering strain
0.2%
0
100
200
300
400
500
600
¬1,000
¬500
0
500
1,000
Lattice strain deviation (×10
6
)
Lattice strain deviation (×10
6
)
0.000 0.005 0.01 0.015 0.02 0.025 0.03 0.035
111
200
220
311
Engineering stress (MPa)
Engineering strain
0.2%
a
b
σ
σ
Figure 2 | Lattice strain deviation behaviour in tension. a,b, Engineering
stress–engineering strain curves (black) and lattice strain deviation
behaviour (coloured) of nt-Cu along the longitudinal direction at room
temperature (a) and 180
C (b). The 0.2% macroscopic yield stress is
marked as σ
0.2%
. The strain rate is 6× 10
5
s
1
for both temperatures. The
plots show zoomed-in regions of the stress–strain curves corresponding to
plastic strains of 0.027 and 0.033 for the samples deformed at room
temperature and 180
C, respectively.
deformation mechanisms (see the inverse pole figure shown in
Fig. 1b, inset). The behaviour further indicates that the mi-
croplasticity initiates well before the macroscopic yielding occurs
in these materials, probably owing to some easy dislocation sources
(such as those observed from IPFOM). Moreover, we find that
the deformation of nt-Cu is sensitive to the temperature (Fig. 2b).
The yield stress drops by 43% at 180
C, the plastic deformation
of which is clearly more isotropic as the difference of among
various crystallographic orientations becomes substantially smaller
compared with room temperature.
The room-temperature peak broadening at different strain levels
plotted according to the Williamson–Hall approach (Fig. 3) sug-
gests that there is a significant amount of dislocation accumu-
lation and the development of strong anisotropy during loading
and after fracture, as evidenced by the increased Williamson–Hall
slope and the continuously broadening of all peaks (for example,
311) vis-à-vis the as-deposited state
26
. A substantial portion of
peak broadening remains permanent after fracture, in contrast
to the reversible peak broadening and strong plastic recovery in
nanocrystalline materials
23,27
. This behaviour generally agrees with
the evolution of the inhomogeneous (root-mean-square) strain
of nt-Cu and the texture development (Supplementary Fig. S8).
The increased dislocation density during deformation and after
0.01
0.02
0.03
0.04
0.05
0.06
20 30 40 50 60 70 80
As-deposited
0.01 strain
0.02 strain
0.03 strain
Fractured
FWHM (nm
¬1
)
Q (nm
¬1
)
111
200
220
311
222
400
Increased strain
Figure 3 | Full-width at half-maximum (FWHM) as a function of the
scattering vector (Q) at various loading strains and after fracture (that is,
Williamson–Hall plots). The scattering vector Q = 4πsinθ /λ, where 2θ is
the angle between incident and diffracted beams, and λ is the wavelength
(nm). These data are extracted from the profile fitting of 300 synchrotron
spectra recorded during the entire tensile test at room temperature. The
strain labelled in the figure includes macroscopic elastic and plastic
components (except for the fractured sample).
fracture contrasts with the lack of strain hardening in this material,
which may suggest other softening mechanisms that counterbalance
the dislocation hardening. Post-mortem TEM examinations of
the room-temperature-deformed sample shown in Fig. 4a indicate
strong accumulations of dislocations between the TBs and inside
the grains. TEM also reveals the disappearance of some growth
twins (that is, detwinning) in certain regions at room temperature.
At 180
C, most TBs are annihilated, concomitant with grain
coarsening (Fig. 4b). These TEM observations, together with in situ
SXRD tensile results, offer some clues that these defective TBs are
unstable against thermo-mechanical deformations.
Although the experimental observations reported above are
qualitatively self-consistent, we speculate that the real atomic pro-
cesses involved with these kinked TBs during deformation are quite
complex. Therefore, molecular dynamics simulations using models
built from kinked TBs and perfect TBs (Supplementary Fig. S9) are
performed to explore this question. Two types of GB, smooth or
defective, are also incorporated to tackle the effects of GB character-
istics on mechanical behaviour; that is, GBs with either high or low
yield stresses, respectively (Supplementary Fig. S10). Furthermore,
owing to the highly idealized models (for example, the tensile axis
is perfectly parallel to TBs when in tension and thus no critical
resolved shear stress), we simulate the deformation behaviour of
nt-Cu under both uniaxial tension and pure shear conditions, as the
most likely loading scenario in the experiments is the mix of both
due to the incline of TBs (Fig. 1a). From the simulations, we find
that the nt-Cu models with kinked TBs have rather rich deformation
processes that do not exist in the case of perfect TBs.
The stress–strain curves from different simulated microstruc-
tures deformed in tension are shown in Fig. 5a. The highest yield
point and ultimate strength are observed in the sample with smooth
GBs and perfect TBs, suggestive of difficult plastic deformation
processes in these near-perfect samples. Threading dislocations are
observed to emit from GBs and glide along slip directions parallel
to TBs (Supplementary Table S2 and Supplementary Movie S1).
Although this type of crystal slip has been considered as a hard mode
in nt-Cu, with potential slip resistance owing to the constraints of
small twin spacing on dislocation glide
28
, newly emitted threading
dislocations are found to transmit easily without blockage through
NATURE MATERIALS | VOL 12 | AUGUST 2013 | www.nature.com/naturematerials 699
© 2013 Macmillan Publishers Limited. All rights reserved

LETTERS
NATURE MATERIALS DOI: 10.1038/NMAT3646
ab
0.5 µm 0.5 µm
Figure 4 | Microstructure of post-mortem tensile samples. a,b, Bright-field TEM images of nt-Cu deformed at room temperature (a) and 180
C (b). Note
the loss of some growth twin structure in a (the region marked with asterisks), and the strong coarsening and complete disappearance of growth twins in b.
0% 1.7% tension 5% tension
P
P
P
M
L
L
M
M
MM
M
M
CTB
GBGB
cde
a
b
Pure tension
0123
Strain (%)
450123
Strain (%)
45
Density of new
dislocations (10
15
m
¬2
)
¬5
0
5
0
1
2
3
Stress (GPa)
Smooth GB + perfect TB
Defective GB + perfect TB
Smooth GB + kinked TB
Defective GB + kinked TB
Smooth GB + perfect TB
Defective GB + perfect TB
Smooth GB + kinked TB
Defective GB + kinked TB
Figure 5 | Molecular dynamics simulations of deformation mechanisms in nt-Cu subjected to uniaxial tension with a twin spacing of 5 nm.
a, Stress–strain curves for four idealized models containing either smooth or defective GBs, and perfect TBs or kinked TBs. Each arrow shows the onset of
yielding. b, Density of new dislocations as a function of applied strain. ce, Microstructure and deformation mechanisms in a columnar grain with both
defective GBs and kinked TBs at different stages of deformation. Vertical arrows indicate the positions of migrating kink imperfections. P, partial
dislocations emitted from GBs; M, kink migration; L, Lomer locks.
perfect TBs (ref. 29), and disappear in the opposite GB in the course
of deformation (Supplementary Fig. S11 and Supplementary Movie
S1). These mechanisms result in no increase, or even a decrease,
in dislocation density in all models with perfect TBs (Fig. 5b).
The inclusion of defective GBs (similar to those observed from
experiments) substantially decreases the initial yield stress with
no change in deformation mechanisms. The same effects of GB
structure dominate the yield point of nt-Cu containing kink defects
(Fig. 5a,c,d, Supplementary Table S2 and Supplementary Movie
S2a). Remarkably, the introduction of kinks results in significant
dislocation pinning giving rise to the formation of Lomer–Cottrell
locks at the intersection of threading dislocations and kinks
(Fig. 5c–e, Supplementary Discussion, Supplementary Fig. S12 and
Supplementary Movie S2b). This kink-dependent mechanism is
700 NATURE MATERIALS | VOL 12 | AUGUST 2013 | www.nature.com/naturematerials
© 2013 Macmillan Publishers Limited. All rights reserved

NATURE MATERIALS DOI: 10.1038/NMAT3646
LETTERS
consistent with the net increase in dislocation density observed after
initial yielding in our molecular dynamics simulations (Fig. 5b) and
in situ SXRD experiments (Fig. 3). The simulations further indicate
that these Lomer–Cottrell locks are short-lived and can be destroyed
on further deformation. Furthermore, marked softening is observed
at the stress peak (Fig. 5a and Supplementary Table S2) due to
partial dislocations nucleated at kink–GB junctions, gliding along
kink lines (Supplementary Fig. S13 and Supplementary Discussion).
Concurrently, the pinning of threading dislocations decreases, as
evidenced by the maximum in dislocation density reached in nt-Cu
with kinked TBs (Fig. 5b) because of kink migration, which also
contributes to the softening behaviour. Motion of kinks closest to
GBs is observable at the yield point, but becomes more important
to the plastic flow at larger strains (Fig. 5e) because of the low
resolved shear stress on kink steps in pure tension. We propose
that the competition between Lomer–Cottrell lock formation at
the initial yield point and kink-dependent softening mechanisms
at higher stresses may have contributed to the change in 200
lattice strain deviation observed in in situ SXRD experiments
(Supplementary Discussion).
Under pure shear stresses, the deformation of nt-Cu becomes
exclusively dominated by kink motion and migration of TBs (Sup-
plementary Table S2, Supplementary Fig. S14 and Supplementary
Movie S3). Interestingly, dislocations no longer nucleate from GBs
but rather from kink ledges (Supplementary Fig. S14d and Supple-
mentary Movie S3). The kinks are absorbed by GBs after migration,
which leads to detwinning. The shear stress for kink migration is
significantly smaller than that required to emit twinning partial dis-
locations from GBs in models with perfect TBs (Supplementary Fig.
S14a,b and Supplementary Movie S4). These results demonstrate
that shear stresses, such as those in the case of high-pressure torsion
from the experiments
30
, favour kink migration and detwinning—a
notion that bears surprising similarity to GB migrations
11
. The easy
motion and migration of kinks under pure shear qualitatively agree
with the early microscopic yielding of these materials and some
detwinning seen in in situ SXRD and TEMs.
Despite the remaining open questions and nearly idealized
approach of modelling, the combination of experimental and
molecular dynamics results here requires that CTBs be treated
as defective entities, in contrast to the traditional perfect-TB
assumption. The migration of kinks is an emphatic sign that
CTBs have intrinsic instability, notwithstanding experimental
observations that they tend to be more stable than conventional
GBs (refs 20,30). The migration of CTBs caused by kink motions
suggests that CTBs close to GBs or longer CTBs with smaller spacing
are more vulnerable to detwinning. This scenario is different from
the detwinning process through partial dislocation glide proposed
by modelling or from the detwinning caused by ITB motions
observed in some experiments
17,31
. Nevertheless, the kink-motion
proposition has strong implications to the detwinning phenomena
observed so far experimentally (for example, detwinning remains
active even at low temperatures and often occurs in certain
regions)
7,14,15
, as well as a maximum strength seen in nt-Cu (ref. 3).
The detwinning near GBs is likely to help unlock triple junctions
between GBs and TBs, facilitating GB migration and contributing
further to the strain softening. For short CTBs of a few nanometres,
a regime that is dominated by molecular dynamics simulations,
the likelihood of generating kinks decreases, and CTBs can be
considered quasi-perfect. On the other hand, the CTBs in the
materials experimentally studied and reported on here are in the
submicrometre range and observed to have inherent imperfections.
The presence of such defective CTBs must be considered in both
scientific research and practical utilities. We, in addition, believe
that the attention being given to potential defects in CTBs resulting
from annealing, deformation and various synthesis conditions is
desirable to fully understand CTB-strengthened materials.
Methods
The nanotwinned copper samples were synthesized by a layer-by-layer magnetron
sputtering technique. The details of sample conditions, twin density and geometry
are given in the Supplementary Information. The intrinsic nanoscale structures
of the as-synthesized CTB were characterized using an IPFOM technique in a
TEM. The details of the IPFOM technique, resolution, potential pixellation effect,
kink defect density, and additional high- and low-resolution IPFOM images are
available in the Supplementary Information. The tensile properties of nanotwinned
copper were characterized using an in situ SXRD approach performed at the
Advanced Photon Source of Argonne National Laboratory (USA) at two different
temperatures. Multiple nanotwinned samples were measured, and the in situ
deformation information was extracted from the lattice strain deviations, peak
profile and texture analysis (see more in the Supplementary Information).
In addition, the post-mortem deformation microstructure was examined by
conventional TEM. Multi-million atom molecular dynamics simulations were
carried out using models built from experimental observations, where both
perfect and defective twin boundaries were used, and the simulation results are
compared with each other and to experimental findings. Further simulation
details, results and movies are available in the Supplementary Information,
together with a Supplementary Discussion on lattice strain deviation behaviour,
additional deformation mechanisms, and the relevance of defective structures to
the mechanical behaviour of nanotwinned copper.
Received 7 November 2012; accepted 2 April 2013;
published online 19 May 2013
References
1. Zhu, Y. T., Liao, X. Z. & Wu, X. L. Deformation twinning in nanocrystalline
materials. Prog. Mater. Sci. 57, 1–62 (2012).
2. Jang, D. C., Li, X. Y., Gao, H. J. & Greer, J. R. Deformation mechanisms in
nanotwinned metal nanopillars. Nature Nanotech. 7, 594–601 (2012).
3. Lu, L., Chen, X., Huang, X. & Lu, K. Revealing the maximum strength in
nanotwinned copper. Science 323, 607–610 (2009).
4. Li, X. Y., Wei, Y. J., Lu, L., Lu, K. & Gao, H. J. Dislocation nucleation governed
softening and maximum strength in nano-twinned metals. Nature 464,
877–880 (2010).
5. Lu, K., Lu, L. & Suresh, S. Strengthening materials by engineering coherent
internal boundaries at the nanoscale. Science 324, 349–3652 (2009).
6. Yu, Q. et al. The nanostructured origin of deformation twinning. Nano Lett.
12, 887–892 (2012).
7. Shute, C. J. et al. Detwinning, damage and crack initiation during cyclic
loading of Cu samples containing aligned nanotwins. Acta Mater. 659,
4569–4577 (2011).
8. Chen, K. C., Wu, W. W., Liao, C. N., Chen, L. J. & Tu, K. N. Observation of
atomic diffusion at twin-modified grain boundaries in copper. Science 321,
1066–1069 (2008).
9. Lu, L., Shen, Y. F., Chen, X. H., Qian, L. H. & Lu, K. Ultrahigh strength and
high electrical conductivity in copper. Science 304, 422–426 (2004).
10. Ovid’ko, I. A. Materials science—deformation of nanostructures. Science 295,
2386–2386 (2002).
11. Rupert, T. J., Gianola, D. S., Gan, Y. & Hemker, K. J. Experimental observations
of stress-driven grain boundary migration. Science 326, 1686–1690 (2009).
12. Schiotz, J. & Jacobsen, K. W. A maximum in the strength of nanocrystalline
copper. Science 301, 1357–1359 (2003).
13. Szlufarska, I., Nakano, A. & Vashishta, P. A crossover in the mechanical
response of nanocrystalline ceramics. Science 309, 911–914 (2005).
14. You, Z. S., Lu, L. & Lu, K. Tensile behavior of columnar grained Cu with
preferentially oriented nanoscale twins. Acta Mater. 59, 6927–6937 (2011).
15. Wang, Y. D. et al. Low temperature deformation detwinning-a reverse mode of
twinning. Adv. Eng. Mater. 12, 906–911 (2010).
16. Wu, Z. X., Zhang, Y. W. & Srolovitz, D. J. Deformation mechanisms, length
scales and optimizing the mechanical properties of nanotwinned metals.
Acta Mater. 59, 6890–6900 (2011).
17. Wang, J. et al. Detwinning mechanisms for growth twins in face-centered cubic
metals. Acta Mater. 58, 2262–2270 (2010).
18. Stukowski, A., Albe, K. & Farkas, D. Nanotwinned fcc metals: Strengthening
versus softening mechanisms. Phys. Rev. B 82, 224103 (2010).
19. Liu, H. H. et al. Three-dimensional orientation mapping in the transmission
electron microscope. Science 332, 833–834 (2011).
20. Bezares, J. et al. Indentation of nanotwinned fcc metals: Implications for
nanotwin stability. Acta Mater. 60, 4623–4635 (2012).
21. Rauch, E. F., Rouvimov, S., Nicolopoulos, S. & Moeck, P. High throughput
automated crystal orientation and phase mapping of nanoparticles from
HREM–TEM images. Microsc. Microanal. 15, 756–757 (2009).
22. Mackenzie, J. K. 2nd paper on statistics associated with the random
disorientation of cubes. Biometrika 45, 229–240 (1958).
23. Budrovic, Z., Van Swygenhoven, H., Derlet, P. M., Van Petegem, S. &
Schmitt, B. Plastic deformation with reversible peak broadening in
nanocrystalline nickel. Science 304, 273–276 (2004).
NATURE MATERIALS | VOL 12 | AUGUST 2013 | www.nature.com/naturematerials 701
© 2013 Macmillan Publishers Limited. All rights reserved

Citations
More filters
01 May 1993
TL;DR: Comparing the results to the fastest reported vectorized Cray Y-MP and C90 algorithm shows that the current generation of parallel machines is competitive with conventional vector supercomputers even for small problems.
Abstract: Three parallel algorithms for classical molecular dynamics are presented. The first assigns each processor a fixed subset of atoms; the second assigns each a fixed subset of inter-atomic forces to compute; the third assigns each a fixed spatial region. The algorithms are suitable for molecular dynamics models which can be difficult to parallelize efficiently—those with short-range forces where the neighbors of each atom change rapidly. They can be implemented on any distributed-memory parallel machine which allows for message-passing of data between independently executing processors. The algorithms are tested on a standard Lennard-Jones benchmark problem for system sizes ranging from 500 to 100,000,000 atoms on several parallel supercomputers--the nCUBE 2, Intel iPSC/860 and Paragon, and Cray T3D. Comparing the results to the fastest reported vectorized Cray Y-MP and C90 algorithm shows that the current generation of parallel machines is competitive with conventional vector supercomputers even for small problems. For large problems, the spatial algorithm achieves parallel efficiencies of 90% and a 1840-node Intel Paragon performs up to 165 faster than a single Cray C9O processor. Trade-offs between the three algorithms and guidelines for adapting them to more complex molecular dynamics simulations are also discussed.

29,323 citations

Journal ArticleDOI
TL;DR: The potential of additive manufacturing to create alloys with unique microstructures and high performance for structural applications is demonstrated, with austenitic 316L stainless steels additively manufactured via a laser powder-bed-fusion technique exhibiting a combination of yield strength and tensile ductility that surpasses that of conventional 316L steels.
Abstract: Many traditional approaches for strengthening steels typically come at the expense of useful ductility, a dilemma known as strength-ductility trade-off. New metallurgical processing might offer the possibility of overcoming this. Here we report that austenitic 316L stainless steels additively manufactured via a laser powder-bed-fusion technique exhibit a combination of yield strength and tensile ductility that surpasses that of conventional 316L steels. High strength is attributed to solidification-enabled cellular structures, low-angle grain boundaries, and dislocations formed during manufacturing, while high uniform elongation correlates to a steady and progressive work-hardening mechanism regulated by a hierarchically heterogeneous microstructure, with length scales spanning nearly six orders of magnitude. In addition, solute segregation along cellular walls and low-angle grain boundaries can enhance dislocation pinning and promote twinning. This work demonstrates the potential of additive manufacturing to create alloys with unique microstructures and high performance for structural applications.

1,385 citations

Journal ArticleDOI
TL;DR: In this paper, the authors review recent advances in overcoming this tradeoff, by purposely deploying heterogeneous nanostructures in an otherwise single-phase metal, and advocate this broad vision to help guide future innovations towards a synergy between high strength and high ductility.

611 citations

Journal ArticleDOI
TL;DR: In this article, a review of recent basic research on two classes of twins: growth twins and deformation twins is presented, focusing primarily on studies that aim to understand, via experiments, modeling, or both, the causes and effects of twinning at a fundamental level.
Abstract: This article reviews recent basic research on two classes of twins: growth twins and deformation twins. We focus primarily on studies that aim to understand, via experiments, modeling, or both, the causes and effects of twinning at a fundamental level. We anticipate that, by providing a broad perspective on the latest advances in twinning, this review will help set the stage for designing new metallic materials with unprecedented combinations of mechanical and physical properties.

318 citations

Journal ArticleDOI
TL;DR: In this paper, the authors summarized and analyzed the current understandings on the influence of various types of internal defect sinks on reduction of radiation damage in primarily nanostructured metallic materials, and partially on nanoceramic materials.

288 citations

References
More filters
Journal ArticleDOI
TL;DR: In this article, three parallel algorithms for classical molecular dynamics are presented, which can be implemented on any distributed-memory parallel machine which allows for message-passing of data between independently executing processors.

32,670 citations

01 May 1993
TL;DR: Comparing the results to the fastest reported vectorized Cray Y-MP and C90 algorithm shows that the current generation of parallel machines is competitive with conventional vector supercomputers even for small problems.
Abstract: Three parallel algorithms for classical molecular dynamics are presented. The first assigns each processor a fixed subset of atoms; the second assigns each a fixed subset of inter-atomic forces to compute; the third assigns each a fixed spatial region. The algorithms are suitable for molecular dynamics models which can be difficult to parallelize efficiently—those with short-range forces where the neighbors of each atom change rapidly. They can be implemented on any distributed-memory parallel machine which allows for message-passing of data between independently executing processors. The algorithms are tested on a standard Lennard-Jones benchmark problem for system sizes ranging from 500 to 100,000,000 atoms on several parallel supercomputers--the nCUBE 2, Intel iPSC/860 and Paragon, and Cray T3D. Comparing the results to the fastest reported vectorized Cray Y-MP and C90 algorithm shows that the current generation of parallel machines is competitive with conventional vector supercomputers even for small problems. For large problems, the spatial algorithm achieves parallel efficiencies of 90% and a 1840-node Intel Paragon performs up to 165 faster than a single Cray C9O processor. Trade-offs between the three algorithms and guidelines for adapting them to more complex molecular dynamics simulations are also discussed.

29,323 citations

Book
01 Jan 1968
TL;DR: Dislocations in Isotropic Continua: Effects of Crystal Structure on Dislocations and Dislocation-Point-Defect Interactions at Finite temperatures.
Abstract: Dislocations in Isotropic Continua. Effects of Crystal Structure on Dislocations. Dislocation-Point-Defect Interactions at Finite Temperatures. Groups of Dislocations. Appendixes. Author and Subject Indexes.

10,220 citations

Journal ArticleDOI
TL;DR: The Open Visualization Tool (OVITO) as discussed by the authors is a 3D visualization software designed for post-processing atomistic data obtained from molecular dynamics or Monte Carlo simulations, which is written in object-oriented C++, controllable via Python scripts and easily extendable through a plug-in interface.
Abstract: The Open Visualization Tool (OVITO) is a new 3D visualization software designed for post-processing atomistic data obtained from molecular dynamics or Monte Carlo simulations. Unique analysis, editing and animations functions are integrated into its easy-to-use graphical user interface. The software is written in object-oriented C++, controllable via Python scripts and easily extendable through a plug-in interface. It is distributed as open-source software and can be downloaded from the website http://ovito.sourceforge.net/.

8,956 citations

Journal ArticleDOI
TL;DR: Calibration methods and software have been developed for single crystal diffraction experiments, using both approaches for calibrate, and apply corrections, to obtain accurate angle and intensity information.
Abstract: Detector systems introduce distortions into acquired data. To obtain accurate angle and intensity information, it is necessary to calibrate, and apply corrections. Intensity non-linearity, spatial distortion, and non-uniformity of intensity response, are the primary considerations. It is better to account for the distortions within scientific analysis software, but often it is more practical to correct the distortions to produce ‘idealised’ data. Calibration methods and software have been developed for single crystal diffraction experiments, using both approaches. For powder diffraction experiments the additional task of converting a two-dimensional image to a one-dimensional spectrum is used to allow Rietveld analysis. This task may be combined with distortion correction to produce intensity information and error estimates. High-pressure experiments can introduce additional complications and place new demands on software. Flexibility is needed to be able to integrate different angular regions se...

4,426 citations