You are on page 1of 13

Acta Materialia 51 (2003) 4135–4147

www.actamat-journals.com

Dislocation–dislocation and dislocation–twin reactions in


nanocrystalline Al by molecular dynamics simulation
V. Yamakov a, D. Wolf a,∗, S.R. Phillpot a, H. Gleiter b
a
Argonne National Laboratory, Materials Science Division, Argonne, IL 60439, USA
b
Forschungszentrum Karlsruhe, 76021 Karlsruhe, Germany

Received 28 January 2003; received in revised form 8 April 2003; accepted 8 April 2003

Abstract

We use massively parallel molecular dynamics simulations of polycrystal plasticity to elucidate the intricate dislo-
cation dynamics that evolves during the process of deformation of columnar nanocrystalline Al microstructures of grain
size between 30 and 100 nm. We analyze in detail the mechanisms of dislocation–dislocation and dislocation–twin
boundary reactions that take place under sufficiently high stress. These reactions are shown to lead to the formation
of complex twin networks, i.e. structures of coherent twin boundaries connected by stair-rod dislocations. Consistent
with recent experimental observations, these twin networks may cause dislocation pile-ups and thus give rise to
strain hardening.
 2003 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.

Keywords: Computer simulation; Nanocrystal; Fcc metals; Deformation twinning; Dislocations

1. Introduction experimental reports on deviations from the Hall–


Petch relation [7] suggest that the usual dislocation
The processes governing the deformation of pile-up mechanism, operating at the grain bound-
nanocrystalline materials (i.e. polycrystals with a aries (GBs) in coarse-grained materials, may not
grain size of less than 100 nm [1,2]) are still the operate in nanocrystalline materials [8–10].
subject of considerable debate [2–5]. While it is Instead, it appears that the relation between dislo-
commonly accepted that the intrinsic deformation cation and GB processes may be considerably
behavior of these materials must arise from the more complex.
dynamic interplay between dislocation and grain- The issue of exactly what role the dislocations
boundary processes [6], little is known about the play in the deformation of nanocrystalline
specific deformation mechanisms. For example, materials remains unresolved. On the one hand,
many experimental studies on nanocrystalline fcc
metals observe high dislocation densities which

Corresponding author. Tel.: +1-630-252-5205; fax: +1- suggest that dislocations actively participate in the
630-252-4798. deformation process. For example, studies in nan-
E-mail address: wolf@anl.gov (D. Wolf). ocrystalline samples, prepared by a variety of syn-

1359-6454/03/$30.00  2003 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
doi:10.1016/S1359-6454(03)00232-5
4136 V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147

thesis processes, [4,5,11–13] show that at grain to be considered, revealed complete development
sizes above 10 nm high dislocation densities of deformation twins in the high-SFE metal. Taken
(usually above 1015 m⫺2), comparable to the dislo- together, these simulations [17,20,21] suggest that
cation density in work-hardened coarse-grained deformation twinning may accompany the slip pro-
metals [14], are observed. On the other hand, no cesses in nanocrystalline fcc metals. This sugges-
dislocations were found in ball-milled Cu-Zr with tion has found experimental support in a recent
a grain size under 100 nm [15]. study by Wang et al. [24] who report that twinning
In addition to such experimental studies, recent is an important deformation mode in nanocrystal-
computer simulations have demonstrated that dis- line and ultrafine grains (⬍ 300 nm) of Cu.
locations may, indeed, play a role in the defor- The purpose of this paper is to study the inter-
mation of nanocrystalline fcc metals [16–21]. In play between deformation twinning and dislo-
particular, we have recently demonstrated a con- cation–slip processes in nanocrystalline fcc metals
nection between the grain size and the type and columnar microstructures of grain size between 30
structure of the existing dislocations in terms of a and 100 nm. In particular we focus on the reactions
length-scale competition between the grain diam- between slip dislocations and twin boundaries.
eter and the dislocation splitting distance (i.e. the Reactions between dislocations on different slip
size of the stacking fault (SF), that connects the systems, including the well-known process of
partial dislocations to form extended perfect cross-slip and the formation of Cottrell–Lomer
dislocations) [19]. Our conclusion was that when locks, are also observed and analyzed. Twin-
the grain size is less than the splitting distance, boundary networks, which are structures of coher-
only partial dislocations can be nucleated from the ent twin boundaries connected by stair-rod dislo-
GBs, i.e. single partials which produce SFs as they cations, are shown to form as an outcome of these
glide through the grain until they become incorpor- reactions in grains of 70 and 100 nm in size. It is
ated into the GBs on the opposite side [16–18]. suggested that these formations may lead to the
These SFs remain in the grains as planar defects. strain hardening seen in low-SFE coarse-grained
After the grain interior has been transected by a metals, thus offering an alternative explanation to
number of SFs, further dislocation propagation the experimentally observed Hall–Petch behavior.
cannot take place. The dislocation propagation pro- Molecular dynamics (MD) and molecular statics
cess can therefore operate only during the initial simulations of dislocations and twinning in fcc
stages of deformation and thus differs significantly metals performed to date have been limited to sin-
from the usual slip mechanism observed in coarse- gle crystals. For example, there have been studies
grained materials. The conclusion of these simula- of dislocation–dislocation reactions between differ-
tions was that for the smallest grain sizes the domi- ent slip systems, such as cross-slip in Ni [25] and
nant deformation mechanism must involve GB- Cu [26,27], annihilation of screw dislocations in
mediated processes, such as GB sliding [16–18] Cu [28] and Cottrell–Lomer lock formation and
and GB diffusion [22,23]. By contrast, when the destruction [29]. Dislocation–twin boundary inter-
grain size is larger than the dislocation splitting actions have also been studied recently by atom-
distance, complete nucleation of perfect dislo- istic simulations, but again only in single crystals
cations from the GBs can take place, thereby initi- [30]. These studies agree well with the theoretical
ating slip deformation [19]. analysis for dislocation reactions in a single crys-
In addition to dislocation activity, the very early tal [31,32].
stages of twinning, in terms of the emission of Our columnar model microstructures offer the
coupled partial dislocations from the GBs that pro- possibility to extend these simulation studies on
duce extrinsic SFs, have been observed in a low dislocation and twinning processes in single crys-
stacking-fault energy (SFE) metal, Cu, in some of tals to polycrystalline materials, while preserving
the earlier simulations [17]. Large-scale simula- most of the characteristics of fully three-dimen-
tions [20,21] on a columnar microstructure [19], sional (3D) models. The limitations of the colum-
allowing much larger grain sizes and plastic strains nar model come mainly from the inability, due to
V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147 4137

geometrical constraints, to produce curved dislo-


cation lines, i.e. all dislocation lines are constrained
to be parallel to the columnar axis. This eliminates
some size effects that usually come from the curva-
ture of the dislocation loops in fully 3D systems.
A more thorough discussion of the advantages and
limitations of our columnar model is given in Sec-
tion 2.
The paper is organized as follows. Our simul-
ation approach is briefly summarized in Section 2.
In Sections 3–5 we present the specific dislo-
cation–dislocation, dislocation–stacking fault, and
dislocation–twin reactions observed in our simula-
tions. A summary and conclusions are given in
Section 6.
Fig. 1. Orientations of the four grains relative to the edge of
the simulation cell (x direction, corresponding to the ⬍112⬎
2. Simulation background direction in grain 4; i.e. the misorientation angle of each grain
is given by the angle between its ⬍112⬎ direction and the x
Our study uses a recently developed model for axis). The orientations are mapped onto a [110]-projected snap-
molecular dynamics simulation of polycrystal plas- shot of the system with a grain size of d = 45 nm after its
thermal equilibration at T = 300 K but prior to application of
ticity in nanocrystalline fcc metals [19]. The model external stress. Also indicated are the orientations of the two
considers idealized, ⬍110⬎-textured microstruc- sets of {111} slip planes in each grain. Only atoms with nearest-
tures, consisting of four grains of identical size and neighbor coordination deviating from the fcc perfect-crystal
regular hexagonal shape (Fig. 1). The grains are value of 12 are shown. The continuous network of these misco-
oriented relative to the tensile axis (horizontal, X ordinated atoms tracing the GBs indicates that all these [110]
tilt GBs are high-angle GBs. The GBs remain perfectly straight
axis) of the simulation cell by rotations about [110] until the threshold stress for the nucleation of dislocations is
by the angles 0°, 30°, 60° and 90° (respectively reached.
for grains 4, 1, 3 and 2 in Fig. 1). The 12 high-
angle GBs in the system therefore consist of three
groups of identical asymmetric [110] tilt GBs with The minimum thickness of the simulation cell in
a tilt angle of 30°, 60° and 90°. The 30° and 90° the columnar (z) direction is determined by the cut-
GBs are high-energy GBs, showing a highly dis- off radius of the interatomic potential used for the
ordered atomic structure [33–35]. We note that, simulations. Our simulation cell thus contains only
probably due to its severe microstructural con- 10{220} periodically repeated spacings; their
finement in our nanocrystalline model material, the interplanar distance of (√2 / 4)ao results in an initial
structure of the 90° GB, seen in our simulations, z thickness of ~3.5ao (or ~1.42 nm for ao = 4.03
is considerably more disordered than the structure Å for the embedded atom method (EAM) potential
of the same GB, however in bicrystalline form, for Al [36] used here). By contrast with the z direc-
determined by Dahmen et al. [40]. The 60° GBs tion, the simulation-cell size in the x–y plane is
are vicinal to the special Σ = 3, 70.53° tilt mis- determined solely by the grain size; for grain sizes
orientation (i.e. {111} coherent twin boundary) and ranging between 45 and 100 nm, our simulated
consist of dislocations to accommodate the devi- systems thus contain between 450 000 and
ation from the coherent twin boundary. The 2 500 000 atoms.
⬍110⬎ column axis (z direction) ensures that, fol- Several techniques were used to identify the dif-
lowing their nucleation, dislocations can glide on ferent types of defects in our microstructure. As in
either of two {111} slip systems in each grain, our earlier simulations [19], stacking faults were
unimpeded by the 3D-periodic border conditions visualized by common-neighbor analysis [39]
imposed on the simulation cell [19]. which permits the distinction to be made between
4138 V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147

atoms in a local hcp environment and those in an limiting dynamical processes be compared with all
fcc environment. A single line of hcp atoms thus other relevant characteristic stresses and time
represents a twin boundary; two adjacent hcp lines scales in the system.
indicate an intrinsic stacking fault and two hcp Concerning the present simulations, in spite of
lines with an fcc line between them represents an the large tensile stresses (of 2.3 GPa) and strain
extrinsic stacking fault. The atoms that were found rates (~107 s⫺1) under which deformation is
to be neither in the fcc, nor in the hcp environment observed, typical dislocation-glide velocities (of
were marked as ‘disordered’ atoms. They are about 500 m/s) are well below the sound velocity
mostly seen in the GBs (see Fig. 1) and in the dis- (of 3664 m/s in the [100] direction for this poten-
location cores. Thus a single spot of disordered tial [28], compared to the experimental value of
atoms is most likely to be a core of a dislocation 3050 m/s). More importantly, however, the
or a vacancy. If this core is connected to a SF, then resolved shear stresses, of the order of 1 GPa on
we see a partial dislocation. In addition a lattice the (111) slip planes of our system, are well below
circuit analyses [31,32] on (110) plane (the x–y the theoretical shear strength, s th = 2.84 GPa esti-
plane in our columnar model in Fig. 1) was used mated by first-principles calculations for Al at 0
to identify the edge component of the dislocation’s K [41].
Burgers vector (due to the [110]-columnar The issue of the high strain rates in MD simula-
geometry of the model, the dislocation line is tions was also recently addressed in high-tempera-
always perpendicular to (110) plane as explained ture deformation simulations of nanocrystalline Pd
in the Section 3). To identify the screw component [23]. These simulations revealed that, in spite of
we use a cross-section of the dislocation’s glide the extremely high strain rates ( ⬎107 s⫺1), the
plane, as defined by the edge component and the Coble-creep equation [42] describing GB diffusion
dislocation line, where the atoms’ displacements at creep in coarse-grained materials was quantitat-
the dislocation core show the dislocation shift. An ively reproduced. Most importantly, the activation
example of such a cross-section to analyze an energy for the creep rate was found to agree com-
extended dislocation was given in Ref. [19]. pletely with that obtained from separate simula-
All of our simulations were carried out under tions of self-diffusion in bicrystalline GBs, i.e. in
constant tensile loads of 2.3–2.5 GPa applied along the absence of applied stress [33]. These simula-
the x-axis (see Fig. 1) and at a temperature of T tions demonstrate that the artificially high strain
= 300 K. (The melting point for this interatomic rates present in MD simulations, and the necessar-
potential was estimated [36] at 940 K, compared to ily high applied stresses, need not change the
the experimental value of 933 K.) The Parrinello– underlying GB physics. This is also supported by
Rahman constant stress technique [37] combined the experimental work of Pond and Glass [43] on
with the Nose–Hoover thermostat [38] and an inte- high-velocity deformation of Al single crystals
gration time step of 0.5×10⫺15 s was used to achi- showing that the deformation modes of Al at strain
eve the constant stress–constant temperature rates of up to 106 s⫺1 do not differ significantly
dynamics of the system. from those at slow-rate static testing.
It is important to be aware of the fundamental
limitations inherent to this model. Apart from
being limited to systems which are relatively small 3. Reactions between dislocations on different
in size (even in the columnar geometry considered slip systems
here), by their very nature MD simulations are
restricted to very high strain rates and, hence, To investigate the complex dislocaton dynamics
stress. For example, a strain of 1% occurring in 10 taking place in our nanocrystalline microstructures,
ns of simulation time corresponds to a strain rate we will start by analyzing several well-known dis-
of 106 s⫺1. It is therefore imperative that for each location reactions, such as cross-slip and formation
type of simulation, the characteristic stresses and of stair-rod locks as they are observed in our simul-
time scales associated with the deformation rate- ations. Although these are well-known processes,
V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147 4139

their mechanisms have not been directly seen A or B node of the Thompson triangle in Fig. 2.
before in simulations of polycrystalline materials. These 60° dislocations produce a shear defor-
More importantly, the crystallographic analysis of mation along the glide plane equal to the edge
these processes will help us to understand the more component of their total Burgers vector, which lies
complicated dislocation–twin reactions taking in the deformation plane (x–y plane). These dislo-
place in our system at larger strain. cations are commonly seen in our model because
they are needed to accommodate the resolved shear
3.1. Crystallographic analysis of the dislocations stress on the glide planes.
and cross-slip The second type of dislocation is a pure screw
BA. Being common to both glide planes, (111)
Two types of intrinsically dissociated dislo- and (111̄), it can dissociate on either one of them,
cations can exist on each of the two glide planes as Bd + dA in the (111) plane or as Bg + gA in
of each grain in our columnar microstructure. the (111̄) plane. Given these two possibilities of
Using the well-known Thompson crystallographic dissociation, this pure screw dislocation can pro-
notation [44] to represent the possible combi- duce cross-slip between the (111) and (111̄) planes.
nations of dislocations on these glide planes (Fig. Because its Burgers vector is along the columnar,
2), we identify these two types as following. z axis and, as defined in Section 2, perpendicular
The first type consists of the 60° CA and CB to the deformation plane, the BA screw dislocation
dislocations on the (111) plane and the 60° DA and does not contribute to the deformation and, thus,
DB dislocations on the (111̄) plane. They all dis- its nucleation is a rare event in our simulations.
sociate into a pure-edge partial, Cd or Dg combined Nevertheless, because of its dissociation into par-
with a 30° partial, dX or gX, where X is either the tials which do have a component in the defor-
mation plane, the BA dislocation can help in
accommodating a local shear stress (on the scale
of its splitting distance) in the [1̄1̄2] or [112] direc-
tions for the (111) and (111̄) planes, respectively.
Such an event is captured in the two snapshots
in Fig. 3. The figure shows a hexagonal grain of
45 nm size (grain 2 in Fig. 1) transected by a defor-
mation twin (the region between the two parallel
lines of hcp atoms starting from the triple junction
on the left and ending at the GB on the right side
of the grain). Parallel to this twin, but slightly
below it, an intrinsic SF (seen as two adjacent hcp
planes) originates at the GB on the right and ends
Fig. 2. Thompson notation in the unfolded Thompson’s tetra- with a single Shockley partial near the center of
hedron [44] for the dislocation–dissociation reactions possible the grain. Ahead of this SF, a moving extended
on the (111) and (111̄) glide planes of our columnar fcc micro- dislocation is marked by the circle. Whereas in Fig.
structure. Both the Thompson notation (such as CA for Burgers 3a this dislocation, being dissociated on the (111)
vector of a perfect dislocation, and Ad for Burgers vector of a plane (parallel to the twin) is gliding towards the
partial dislocation) and the corresponding set of indices for the
same directions are presented. The sense of the directions are left GB, in Fig. 3b taken only 2 ps later, this dislo-
indicated by [ ⬎. The dislocation lines are always restricted to cation has changed its plane of dissociation to
be along BA which is the columnar, z, axis in our microstruc- (111̄) and is now approaching the twin. As shown
ture. The deformation takes place in the x–y plane (see also in the inset in Fig. 3, this process is possible only
Fig. 1) which here, for clarity, is shown as a plane perpendicular because the total Burgers vector of the dissociated
to BA and crossing it in the middle. The dashed gd line con-
necting the middle points of the two Thompson triangles rep- dislocation BA is along the line common to the
resents the 1/6[110] stair-rod dislocation. The directions seen as two slip planes, i.e. the columnar direction. This
underneath the deformation plane are also given as dashed lines. allows for the dissociation to take place on both
4140 V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147

planes such that cross-slip can take place according


to the equation (see also Fig. 2)
BA(111)→Bd ⫹ dA→Bg ⫹ gA→BA(111̄), (1)
with the subscripts indicating the dislocation slip
plane before and after the cross-slip event.
Also in Fig. 3a,b one can see the emission of
two double Shockley partials from the bottom-right
GB, which in their upwards glide produce two
extrinsic SFs. The one on the right has been broad-
ened to a two-layer twin by the consecutive emis-
sion of another single Shockley partial—a process
discussed in detail previously [20,21].
3.2. Stair-rod dislocations
When two partial dislocations on different slip
planes meet, such as dA and Ag (see Fig. 2), they
can react to form a stair-rod dislocation, dg [31,32].
This stair-rod dislocation now connects the two
SFs at the intersection of the two slip planes. Simi-
lar reactions between different combinations of
partials from the two slip planes lead to various
types of stair-rod dislocations. These reactions
have been thoroughly studied in the literature [31].
Those that lead to a reduction of energy form
stable structures that are very resistant to climb or
glide in either of the two slip planes and, thus, pro-
vide barriers to other dislocation motion. In coarse-
grained fcc metals such reactions most commonly
take place through the interaction of two perfect
dislocations dissociated into partials on different
slip planes [31]. The approaching partials react,
producing a stair-rod dislocation which now con-
nects the two SFs from the two remaining partials
and form a triangular structure. A well-known
example of this is the Cottrell–Lomer lock [31,32].
Two formations of this type seen in our simula-
Fig. 3. Snapshots taken at 2 ps apart, revealing the process of tions are analyzed in Fig. 4a in which a number of
a cross-slip event performed by an extended dislocation in grain dissociated dislocations (such as the one marked
2 from Fig. 1 (marked by the circled region in both snapshots, as (1) in Fig. 4a have combined to form locks
see text) at about 10% strain. Here and throughout, common-
neighbor analysis was used to identify perfect-crystal atoms as
marked as (2) and (3). The lock marked as (2) was
being either in a local hcp (gray atoms) or fcc environment identified as a Cottrell–Lomer lock formed by the
(atoms not shown). Black atoms are ‘defected’, i.e. neither in reaction (see Fig. 4b).
an fcc nor hcp environment. The dissociation of the extended
1
dislocation in both slip planes is shown in the inset using the gA ⫹ Ad→gd ⫽ [110], (2)
unfolded Thompson’s tetrahedron representation. 6
where the partials gA and Ad result from the dis-
sociation of two perfect dislocations:
V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147 4141

1
dC ⫹ Dg→dD / Cg ⫽ [110]. (5)
3
The core of this dislocation is visualized by the
disordered atoms (shown as open circles) at the
connecting point of the two SFs in Fig. 4a, forma-
tion (3); this core is larger than the core of the
Cottrell–Lomer lock (formation (2)) because of the
twice larger Burgers vector of this dislocation,
leading to an about four times larger stored energy.
Formation (4) in Fig. 4a is the product of a more
complicated reaction between a partial dislocation
and a SF which will be analyzed in Section 5.

4. Interactions between a dislocation and a


stacking fault

Under the right conditions the emission of the


second partial dislocation in the nucleation process
of an extended dislocation (described in Section 2
and in Ref. [19]) can be prevented. This may hap-
pen because of a too high resolved shear stress on
the glide plane, which can produce SFs larger than
the grain size [19]. It can also happen because of
the particular GB atomic structure at the point of
nucleation, e.g. the presence of a GB dislocation
that cannot dissociate to give a second partial. This
Fig. 4. (a) A snapshot of a region in a 70 nm grain (with the can result in SFs of length of up to several tens of
orientation of grain 4 in Fig. 1) where a number of dissociated nms penetrating into the grain interior (see Fig.
dislocations, like the one marked as (1), have combined to form 3a,b). In most cases these SFs prevent the glide of
locks denoted as (2), (3) and (4). The type of the atoms is
marked as follows: (.)-fcc atoms; (왕)-hcp atoms and (䊊)-dis-
other dislocations on slip planes crossing the SF
ordered atoms. (b) Dislocation–reaction diagram for the locks plane. However, that this may not always be the
(2) and (3) between (111) and (111̄) planes on the unfolded case is shown in Fig. 5a–d, where a SF is cut by
Thompson’s tetrahedron as explained in the text. an approaching dislocation.
In Fig. 5a, a dissociated dislocation approaches
a SF of extrinsic type (two hcp lines separated by
DA→Dg ⫹ gA (3) an fcc line of atoms) produced by a double Shock-
ley partial nucleated from the GB in grain 1 (as
AC→Ad ⫹ dC. (4)
numbered in Fig. 1). As the dissociated dislocation
In this reaction the two SFs of the dissociated dis- comes closer to the SF (Fig. 5b) its elastic field,
locations meet at an acute angle of 70.53°, i.e. the apparently, initiates unfolding of one SF layer
angle between (111) and (111̄) planes. (interestingly, this is the layer at the other side of
In the case of the lock marked as (3), the two the SF) transforming the SF from extrinsic to
reacting dislocations meet at an obtuse angle of intrinsic type (two adjacent hcp lines of atoms).
109.47° and the reaction involves other partials, This produces two partial dislocations with
producing a different stair-rod dislocation (see opposite Burgers vectors, gliding apart from each
Fig. 4b): other along the SF (the cores of disordered atoms
4142 V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147

Fig. 5. Successive snapshots taken at 2 ps time intervals, illustrating the process of splitting of a stacking fault by an approaching
dislocation (marked by the circled region in both snapshots, see text). The process has taken place in grain 1 from Fig. 1 at about
5% strain. The atoms in (a) and (b) are marked as in Fig. 3a,b, and in the closer views (c) and (d) are marked as in Fig. 4a. The
distance between the incoming dislocation and the stacking fault is also given in (a) and (b).

at the points where the extrinsic SFs transform into employed here, have carefully estimated the
intrinsic ones). Fig. 5c further captures the moment resolved shear stress required to split a SF to be
when the incoming dislocation crosses the intrinsic 930 MPa. This value is consistent with our esti-
SF and starts to split it into two fragments. In the mate for the resolved shear stress on the (111)
next stage, revealed in Fig. 5d, the splitting of the planes (see Section 2). This confirms that the split-
intrinsic SF is completed, producing two more par- ting of the SF is not due to a dislocation reaction
tial dislocations gliding apart. Thus the perfect between the incoming dislocation and the SF (as
crystal lattice is restored, allowing the incident dis- Fig. 5c may suggest), but due to a long-range elas-
location to pass through. tic–field interaction. The fact that the incoming dis-
More careful analysis show that the process of location preserves its structure after the interaction,
splitting of a SF can take place even without the supports this conclusion.
help of an incident dislocation if the resolved shear
stress on the SF plane is large enough. Recently
Bulatov et al. [45], using an atomistic model of a
Frank–Read source and the same EAM potential
V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147 4143

5. Reactions between dislocations and


coherent twin boundaries

As discussed in Section 2, the dense network


of GBs in a nanocrystalline metal subject to low
temperature–high strain-rate deformation provides
a good environment for deformation twinning. In
particular, the initial stages of twinning in terms
of formation of extrinsic SFs have been seen in
simulations of nanocrystalline Cu of grain diam-
eters below 10 nm [17]. Also our previous simula-
tions of nanocrystalline Al, exploring grain sizes
up to 100 nm, albeit in columnar microstructures,
revealed extensive deformation twinning as a result
of dislocation–dislocation and dislocation–GB
reactions [20,21]. The deformation twins thus for-
med react with the pre-existing dislocations
nucleated from the GBs, leading to a number of
twinning and untwinning processes. As demon-
strated below, these processes are very similar to
the ones observed in coarse-grained low-SF energy
fcc metals or alloys [46–55] and are generally
based on well-known dislocation–twin reactions
[53–56].
When a dissociated dislocation glides into a
coherent twin boundary, a dislocation–twin reac-
tion takes place which may result in either a twin-
ning or an untwinning process [53–56]. Two types
of reactions, as distinguished by Rusakov and Lit-
vinov [55], are possible. The first type of reaction Fig. 6. Two snapshots in a grain of 70 nm showing (a) a pro-
occurs when the total Burgers vector of the dis- cess of untwinning of a deformation twin by the interaction
between an incident dislocation and a coherent twin boundary
sociated incident dislocation is not parallel to the
with the creation of a stair-rod lock at the boundary. The
approached twin plane. Then a stair-rod dislocation Thompson triangle for (111) plane on the left of the inset shows
is needed to account for the perpendicular compo- the partials constructing the incident extended dislocation
nent of the Burgers vector at the point of reaction. before the reaction with the twin boundary; the coupled tri-
By contrast, the second type of reaction occurs if angles for (111) and (111̄) planes on the unfolded Thompson’s
tetrahedron on the right show the resulting dislocations after the
the incident dislocation has a Burgers vector paral- reaction (see text). (b) a process (opposite to (a)) of broadening
lel to the twin plane. In this case cross-slip can of a deformation twin by an incident dislocation. The atoms are
take place at the twin boundary, transforming the marked as in Fig. 4a.
incident dissociated dislocation into a dissociated
dislocation in the twin plane.
Two examples of the first type of dislocation– twinning dislocation, Ag, and the stair-rod dislo-
twin reaction, resulting either in untwinning or cation, gd, which now links the SF of the incident
twinning processes, are shown in Fig. 6a,b. In Fig. dislocation to the twin boundary according to the
6a an incident dislocation AC in the (111) plane reaction:
(see the inset) has dissociated into Ad + dC par-
AC→Ad ⫹ dC→Ag ⫹ gd ⫹ dC. (6)
tials. The leading partial, Ad, has reacted at the
twin boundary on the (111̄) plane producing the Depending on its glide direction, Ag can either
4144 V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147

remove (Fig. 6a) or add (Fig. 6b) one lattice plane location, Ag, gliding on the SF plane (see Eq. (6)).
to the twinned region. As seen in Fig. 4a, in its glide Ag has started to
The combination of gd and the remaining partial transform the intrinsic SF into an extrinsic SF in
dC is sometimes considered as a dissociated Frank the same way as a twinning dislocation broadens
partial dislocation, gC [56]. Recent HRTEM obser- a twin (see Fig. 6b). Finally, the gliding partial has
vations by Medlin et al. [54] and Foiles and Medlin joined the partial ending the SF; together they have
[30] revealed that dissociated Frank dislocations at formed the double Shockly partial seen at the bot-
the coherent twin boundaries, as seen in Fig. 6a,b, tom end of formation (4).
are common in Al. Moreover, using atomistic This example illustrates that the reactions
simulations they showed that an initial compact between dislocations and a coherent twin boundary
Frank partial at the twin boundary is unstable as it may take place also between a dislocation and a
dissociates into a stair-rod and a Shockley partial SF, resulting again in an intrinsic–extrinsic trans-
dislocations separated by a SF from the twin formation of the SF or in cutting off an intrinsic
boundary. At elevated temperatures or in the pres- SF. However, this type of a cutting off of a SF
ence of vacancies this dissociated couple can climb (not shown here) is different from the one shown
along the twin boundary and grow the twin by a in Fig. 5, in that the incident dislocation will form
non-conservative, thermally activated process a stair-rod lock with one piece of the SF while the
[30,54]. other piece, together with the newly formed partial,
Our simulations show that dissociated Frank Ag, forms a perfect dislocation which glides away.
partials can appear from the interaction between a The type of the resulting stair-rod dislocation
gliding perfect dislocation and a twin boundary depends on the type of incident Shockley partial.
(see Eq. (6) and Fig. 6a,b) and actually activate the Eq. (6), which provides a gd dislocation, describes
twinning–untwinning processes by the conserva- only one possible reaction. Another possibility
tive glide of the emitted twinning dislocation—a arises if the leading partial was dC instead of Ad
possibility recently analyzed by Müllner and (i.e. a pure-edge instead of a 30° partial), leading
Solenthaler [53]. A subsequent climb of the to the reaction:
remaining dissociated Frank partial has not been
AC→Ad ⫹ dC→Ag ⫹ dD / Cg ⫹ gD. (7)
observed. However, this is not surprising in our
simulations because: (1) we do not have sources As discussed in connection with Eq. (4), the stair-
of vacancies that can facilitate the climb at low rod, dD/Cg, has a higher stored excess energy; i.e.
temperatures as in the TEM observations of Foiles such a reaction is energetically less favorable than
et al. [54], where they suggest that the 400 KeV the one in Eq. (6). In fact, our simulations indicate
beam is able to produce the necessary vacancies to that a pure-edge partial is usually repelled from the
assist the climb; (2) the time scale of 5 min per twin boundary and that, therefore the above reac-
micrograph during which climb by a few nm is tion (Eq. (7)) takes place only under high stress.
observed gives a climb rate too slow to be repro- An example of the second type of dislocation–
duced in an MD simulation where the time scale twin reaction, which results in a cross-slip event
is in the range of 1 ns. at the twin boundary without leaving a stair-rod
Returning to Fig. 4a, we can now explain forma- dislocation behind, is shown in Fig. 7a,b. In this
tion (4) in terms of a dissociated Frank partial as case the incident dislocation (see Fig. 7a) has a
expressed by Eq. (6). In this case the leading par- Burgers vector parallel to the twin plane. Cross-
tial, Ad of the approaching dissociated dislocation slip at the twin boundary transforms the incident
CA (one on the right in formation (4)) reacts with dissociated dislocation into a dissociated dislo-
the SF of the other dissociated dislocation, instead cation in the twin plane. This creates two twinning
with one of its partials (as in formations (2) and dislocations gliding in opposite directions (Fig.
(3) when a stair-rod lock is formed). Reacting with 7b). The result is a twinning or untwinning process
the SF, the incident partial, Ad, has been dis- (depending on the sense of the Burgers vectors) in
sociated into a stair-rod, gd, and another partial dis- which the incident dissociated dislocation is fully
V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147 4145

ning and cutting off of the twin, by sequential


untwinning of one atomic layer at a time. Such a
process of untwinning due to dislocation bombard-
ments from a nearby GB is seen in Fig. 8. A series
of partial dislocations nucleated from the GB
(upper right in Fig. 8) leaves several traces of
extrinsic SFs on their way down to the twin, con-
tinuously narrowing the twin (from right to left) by
untwinning one atomic layer at a time.
Further development of such processes can lead
to the formation of so-called T-twins [53], i.e.
cross-twinning events such as those seen in Fig.
8. An interesting analysis of how such events can
influence the twin density in a polycrystalline
metal was also given in [53]. Our simulations sug-
gest that these processes may also be relevant in a
nanocrystalline material under deformation con-
ditions when GBs become very active dislocation
sources. Moreover, our simulations for the largest
grain sizes (with a diameter of 70 and 100 nm)
suggest that the dislocation–twin reactions of the
type described above generally lead to the forma-
tion of complex networks of twins. For example,
Fig. 9 shows such a network consisting of twin
boundaries on both the (111) and (111̄) planes,
connected by stair-rod dislocations at the crossing
points. These networks can trap incoming dislo-
cations to form pile-ups, as seen in the right half
of the grain in Fig. 9. These deformation–twin

Fig. 7. Two snapshots in a grain of 70 nm, taken at 2 ps time


intervals, showing untwinning of a deformation twin by an
interaction between an incident dislocation and a coherent twin
boundary without the creation of a stair-rod lock at the bound-
ary. The process is similar to the cross-slip event as shown in
Fig. 3, but here it is helped by the presence of a twin boundary.
The atoms are marked as in Fig. 4a.

absorbed by the twin without the creation of a sess-


ile stair-rod dislocation. In this way, the twin
boundary helps to transfer the slip from one slip
system to another without the necessity of over-
coming the cross-slip energy barrier in the usual
cross-slip process (compare Fig. 7a,b with Fig. 3a,b).
As discussed by Müllner and Solenthaler [53], Fig. 8. A process of untwinning of a deformation twin, due
a series of Shockley partial dislocations incident to dislocation bombardment from a near by GB in a 70 nm
on a twin boundary can lead to complete untwin- grain at about 12% strain.
4146 V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147

cations still coming out of the GBs. Of course,


these reactions must be considered together with
the variety of dislocation–dislocation reactions
going on in such a dense dislocation environment.
The process of deformation twinning can have
a two-fold effect on the mechanical properties of
the material. On one hand, it can facilitate the
deformation by adding additional slip systems or
by facilitating the transfer between existing slip
systems through dislocation–twin reactions. On the
other hand, the twin boundaries can repel certain
types of gliding dislocations and give rise to pile-
ups with consequent strain hardening of the
material. Our simulations suggest that in an unde-
formed nanocrystalline metal the grains are free of
dislocations [19], i.e. statistically stored dislo-
cations, as defined by Ashby [57], are not present.
Fig. 9. A 100 nm grain at about 10% strain. The dislocation– Therefore, the dislocations nucleated from the GBs
twin reactions generally lead to the formation of complex twin early during the deformation can propagate freely
networks consisted of twin boundaries on both (111) and through the grain interiors. At this stage twinning
(111̄) planes, connected by stair-rod dislocations at the crossing can only facilitate the deformation. At a later stage,
points. These networks trap the dislocations in pile-ups, as can
be seen in the right part of the grain, which may lead to an
when the dislocation–dislocation and dislocation–
additional hardening of the nanocrystalline metal. twin reactions have formed a dense network of
obstacles for the gliding dislocations (Figs. 8 and
induced pile-ups may lead to strain-hardening of 9), the material should experience strain-hardening.
the nanocrystalline metal in the range of 6–12% This scenario is supported by recent experiments
strain, a phenomenon recently suggested to take by Wang et al. [24] in which deformation twinning
place in nanocrystalline Cu [24]. in nanocrystalline and ultrafine grains with a grain
size of less than 300 nm of Cu has been observed.
The twinning appears unexpectedly after 6% plas-
6. Discussion and conclusion tic strain in the regions of high stress at the GBs
in most of the larger grains. Consistent with the
In a series of papers [20,21] including this one, above simulations, and those presented in [20] and
we have reported the surprising possibility that [21], the authors [24] suggest that the initiation of
nanocrystalline fcc metals, even if they have a high twinning contributes to the strain-hardening of this
SF energy, can experience intensive twinning dur- material by forming dislocation pile-ups at the
ing deformation. In Refs. [20,21] we reported a twin boundaries.
number of such phenomena and analyzed several The propensity for twinning makes even a high-
mechanisms of twinning, each of them involving SFE nanocrystalline fcc metal to behave, to some
the GBs in one way or another, revealing that twin- degree, as if were a low-SFE coarse-grained metal.
ning arises from an intricate interplay between dis- As proposed in our previous work [19], analytically
location and GB processes. These studies have this correspondence can be expressed by comparing
shown that the ability of the GBs in a nanocrystal- the ratio between the grain size and the average size
line microstructure under high stress to nucleate of the SFs produced by the Shockley partial dislo-
partial dislocations at a high rate inevitably leads cations. When the grain size is very small, even in
to deformation twinning. Here we have gone high-SFE nanocrystalline metals, such as Al, this
further by analyzing the resulting reactions ratio is small enough and comparable to that in low-
between the already formed twins and the dislo- SFE metals or alloys. Of course, this similarity
V. Yamakov et al. / Acta Materialia 51 (2003) 4135–4147 4147

between coarse-grained and nanocrystalline [21] Yamakov V, Wolf D, Phillpot SR, Gleiter H. Acta
materials is expected to be only qualitative. In prac- Mater 2002;50:5005.
[22] Keblinski P, Wolf D, Gleiter H. Interface Science
tice, the yield stress of nanocrystalline metals is 1998;6:205.
much higher, following the Hall–Petch relation. [23] Yamakov V, Wolf D, Phillpot SR, Gleiter H. Acta
Mater 2002;50:61.
[24] Wang Y, Chen M, Zhou F, Ma E. Nature 2002;419:912.
Acknowledgements [25] Rao S, Parthasarathy TA, Woodward C. Phil Mag A
1999;79:1167.
VY, DW and SRP are supported by the US [26] Vegge T, Rasmussen T, Leffers T, Pedersen OB, Jacobsen
Department of Energy, BES-Materials Science KW. Phil Mag Lett 2001;81:137.
under contract W-31-109-Eng-38. VY is also [27] Zhou SJ, Preston DL, Lomdahl PS, Beazley DM.
grateful for support from the DOE/BES Compu- Science 1998;279:1525.
[28] Swaminarayan S, LeSar R, Lomdahl P, Beazley D. J
tational Materials Science Network (CMSN). We
Mater Res 1998;13:3478.
are grateful for grants of computer time on the [29] Bulatov V, Abraham FF, Kubin L, Devincre B, Yip S.
Cray-T3E at the John von Neumann Institute for Nature 1998;391:669.
Computing in Jülich, Germany and on the Chiba [30] Foiles SM, Medlin DL. Mat Sci Eng A 2001;319-321:102.
City Linux cluster at ANL. [31] Hirth JP, Lothe J. Theory of dislocations, 2nd edn. John
Wiley & Sons Inc, 1992.
[32] Weertman J, Weertman JR. Elementary dislocation theory.
References New York, Oxford: Oxford University Press, 1992.
[33] Keblinski P, Phillpot SR, Wolf D, Gleiter H. Phil Mag
[1] Birringer R, Gleiter H, Klein HP, Marquardt P. Phys Lett A 1999;79:2735.
A 1984;102:365. [34] Keblinski P, Phillpot SR, Wolf D, Gleiter H. Phys Rev
[2] Gleiter H. Acta Mater 2000;48:1. Lett 1996;77:2965.
[3] Karch J, Birringer R, Gleiter H. Nature 1987;330:556. [35] Wolf D. Grain boundaries: structure. In: Cahn R, editor.
[4] Nastasi M, Parkin DM, Gleiter H, editors. Mechanical The encyclopedia of materials: science and technology.
properties and deformation behavior of materials having Pergamon Press; 2001. p. 3597.
ultra-fine microstructures. Kluwer; 1993. [36] Ercolessi F, Adams JB. Europhys Lett 1994;26:583.
[5] Siegel RW. Mater Sci Forum 1997;851:235. [37] Parrinello M, Rahman A. J Appl Phys 1981;52:7182.
[6] Yip S. Nature 1998;391:532. [38] Melchionna S, Ciccotti G, Holian BL. Mol Phys
[7] Koch CC, Suryanarayana C. Nanocrystalline materials, 1993;78:533.
chapter 6. In: Li JCM, editor. Microstructure and proper- [39] Jonsson H, Andersen HC. Phys Rev Lett 1988;60:2295.
ties of materials, vol 2. Singapore: World Scientific Pub- [40] Dahmen U, Hetherington CJD, O’Keefe MA, Westmacott
lishing; 2000. p. 380–5. KH, Mills MJ, Daw MS, Vitek V. Phil Mag Lett
[8] Nieh TG, Wadsworth J. Scripta Metall 1991;25:955. 1990;62:327.
[9] Scattergood RO, Koch CC. Scripta Metall 1992;27:1195.
[41] Ogata S, Li J, Yip S. Science 2002;298:807.
[10] Wang N, Wang Z, Aust KT, Erb U. Acta Metall Mater
[42] Coble RL. J Appl Phys 1963;34:1679.
1995;43:519.
[43] Pond R, Glass CM. In: Shewnion PG, Zackay VF, editors.
[11] Wunderlich W, Ishida Y, Maurer R. Scripta Metall
Response of metals to high velocity deformation. New
Mater 1990;24:403.
York: Interscience Publishers; 1961. p. 145.
[12] Huang JY, Wu YK, Ye HQ. Acta Mater 1996;44:1211.
[44] Thompson N. Proc Phys Soc 1953;66B:481.
[13] Islamgaliev RK, Chmelik F, Kuzel R. Mat Sci Eng ; A
1997;A234-236:335. [45] de Koning M, Cai W, Bulatov VV. Phys Rev Lett, in press.
[14] Cottrell AH. Dislocations and plastic flow in crystals. [46] Mahajan S, Chin GY. Acta Met 1973;21:173.
Oxford: Clarendon Press, 1961. [47] Mahajan S, Chin GY. Acta Met 1974;22:1113.
[15] Morris DG, Morris MA. Acta Metall Mater 1991;39:1763. [48] Remy L. Scripta Met 1977;11:169.
[16] Schiotz J, Di Tolla FD, Jacobsen KW. Nature [49] Remy L. Metall Trans A 1981;12A:387.
1998;391:561. [50] Hartley CS, Blachon DLA. J Appl Phys 1978;49:4788.
[17] Schiotz J, Di Tolla FD, Jacobsen KW. Phys Rev B [51] Hartley CS. Mat Sci Eng 1986;81:543.
1999;60:11971. [52] Solenthaler C. Mat Sci Eng A 1990;125:57.
[18] Swygenhoven HV, Spaczer M, Caro A. Acta Mater [53] Mullner P, Solenthaler C. Mat Sci Eng A 1997;230:107.
1999;47:3117. [54] Medlin DL, Carter CB, Angelo JE, Mills MJ. Phil Mag
[19] Yamakov V, Wolf D, Salazar M, Phillpot SR, Gleiter H. A 1997;75:733.
Acta Mater 2001;49:2713. [55] Rusakov GM, Litvinov VS. Phys Met Metall 2000;89:445.
[20] Yamakov V, Wolf D, Phillpot SR, Mukherjee A, Gleiter [56] Frank FC. Proc Phys Soc 1949;62A:202.
H. Nat Mater 2002;1:45. [57] Ashby MF. Phil Mag 1970;21:399.

You might also like