Crystal Dislocation: Calculation of Cross-Slip Parameters in F.C.C. Crystals
Crystal Dislocation: Calculation of Cross-Slip Parameters in F.C.C. Crystals
While most crystal dislocations are total dislocations when viewed from sufficiently
far away, it is not uncommon to find them dissociated locally into a configuration
that can be described as two parallel partial dislocations connected by a planar defect
that is called a stacking fault in the crystal.
Related terms:
1 Introduction
In f.c.c. crystals dislocations can split into Shockley partials connected by a stacking
fault in a {111} glide plane. In order that cross-slip can take place a screw dislocation
must be constricted locally before it can dissociate in an intersecting {111} plane.
This process can occur by thermal activation aided by an external stress. Since
cross-slip events can control the overall plastic behaviour of crystals, knowledge of
the activation energy and its stress dependence is of considerable interest.
The first calculation of the cross-slip configuration by Schoeck and Seeger [1] was
on the energy level based on the Peierls [2] model taking account of anisotropy.
Although the energy of the critical transition configuration where a metastable
equilibrium exists was calculated correctly, it was assumed on the basis of entropy
arguments that this configuration is reached by a reaction path with higher energy.
The calculations were refined afterwards by Wolf [3] who also made allowance for
the narrowing of the stacking fault due to an applied stress.
Later Escaig [4], following a suggestion of Friedel, calculated the equilibrium con-
figuration on the force level using a method first proposed by Stroh [5] based on
the line tension approximation. This rests on questionable foundations since the
governing differential equation requires small slopes y of the partial dislocations
whereas the solution gives y infinity at the nodes. Further, the interaction energy
of the partials is calculated as if they consisted of parallel straight segments whereas
in the cross-slip nucleus they are strongly curved. The uncertainty of the calculation
is aggravated by the fact that no reliable procedure is given to determine the inner
cut-off radius r0 for the elastic solution or the critical distance for recombination rc
of the partials. They are arbitrarily chosen as r0=rc=b.
More recently a treatment was given by Saada [6] along the same Unes. He based
his arguments on the assumption that the driving force for cross-slip always results
from an applied stress 2 acting in the cross-slip plane. This overlooks completely
the fact that cross-slip can occur even without such a stress when the stacking fault
ribbon is narrowed in the primary plane and/or widened in the cross-slip plane. As
Escaig [4] has emphasized this narrowing-widening (NW) effect generally produces
a larger driving force than the action of 2. Saada [6] recalculates the constriction
energy and points out that without a knowledge of the cut-off radius r0 only the
order of magnitude of the activation energy can be obtained.
Another treatment was given by Duesbery et al. [7] who first calculated the equilibri-
um configuration by the relaxation method of Bacon [8] and Foreman [9]. However,
the resulting cut-off radius r0 also remains uncertain. Duesbery et al. choose r0 = 2b.
Hence values for constriction energies for aluminium (d ≈ 2b), nickel (d ≈ 3b), and
gold (d ≈ 4b) can hardly be viewed with much confidence.
In the usual treatment in the literature in linear elasticity theory, the cut-off radius
r0 is chosen so that the elastic energy outside r0 equals the total energy (i.e. elastic
energy plus core energy) but usually some ad hoc value of r0 is chosen. The ambiguity
of this procedure can be avoided when we choose an r0 in a region where the linear
elastic solution is still very reliable and include explicitly the (atomistic) core energy
in the calculations as done for instance by Piischl et al. [10]. The uncertainty in the
second kind of cut-off radius, the approach distance rc of the partials, can be removed
by demanding that the energy of dissociation in the elastic treatment equals that of
the Peierls model. In the following we give a treatment of the energy level taking
account of anisotropy which is based on this approach avoiding the ambiguity of
the previous treatments.
DISLOCATIONS
J.P. HIRTH, in Physical Metallurgy (Fourth Edition), 1996
In order that the deformation not produce a high energy fault on the cut surface, b
is usually a perfect lattice vector as illustrated for an edge dislocation in fig. 2. The
choice of the ± sense of is arbitrary, but once chosen, the ± sense of b is fixed by
the following convention: imagine a perfect reference crystal, select a vector in it,
and construct a closed circuit in it, right-handed relative to . Then construct the
same Burgers circuit in the real crystal, as shown for example in fig. 2. The vector SF
connecting the start of the circuit to the finish is the Burgers vector of a dislocation if
it is contained within the circuit. In this operation, the circuit must not pass through
the nonlinear core region within an atomic spacing or two of the dislocation line.
Fig. 2. An edge dislocation in a simple cubic crystal. A Burgers circuit is also shown,
projected toward the viewer for clarity so that it passes through the center of atoms
not shown. The sense vector points out of the page.
Some other properties follow directly from the above definitions. If is reversed,
the sense of b is also reversed as seen from fig. 2 since the circuit is also reversed
when is reversed. Since the dislocation line bounds a displaced area, the line
cannot end within otherwise perfect crystal but can only end at a free surface, a
grain boundary, a second-phase interface, or a dislocation node. A node is a point
where two or more dislocation lines join. Translation of a Burgers circuit along
without the circuit cutting through a dislocation core does not change the vector
SF or thus the total b (imagine such an operation for fig. 2); such circuits are called
equivalent Burgers circuits. Thus, if a dislocation denoted by its Burgers vector, b1,
splits into two dislocations b2 and b3, enclosed by equivalent circuits, fig. 5, an analog
of Kirchhoff 's law applies and b1=b2+b3 If the for dislocations 2 and 3 are reversed,
then the signs of b2 and b3 change by the earlier axiom and: Σibi = 0 for dislocations
meeting at a node if all sense vectors are selected to point toward the node.
Figure 9.24. (a) Projection of the atomic positions in two adjacent (110) planes in a
bicrystal formed by 111 {112} twinning in a body-centered cubic metal. The twin
habit plane is K1, the direction of the twinning shear is 1 and [110] is out of the
paper. (b) Dichromatic pattern associated with (a). b=(t −tμ) is the Burgers vector
of a twinning dislocation formed by joining steps t and tµ on the two crystals, as
indicated.
Thus, the atomic displacements that give rise to the macroscopic twinning shear
occur by the glide of these defects on successive planes, as illustrated schematically
in Fig. 9.25. An experimental observation of dislocations of this type is presented
in Fig. 9.26. Figure 9.26(a) illustrates the cross-section shape of the small twin seen
in the transmission electron microscope image in (b) and an explanation for the
contrast from the individual twinning dislocations is sketched in (c).
Figure 9.25. Schematic illustration of twinning in a body-centered cubic metal.
The projection is the same as that in Fig. 9.24 and the dislocations are the same
as the one defined there. (a) Untwinned crystal in [110] projection showing the
stacking sequence A, B, … of the planes. (b) Twinning dislocations with glide to
the right on successive planes under the applied shear stress indicated to produce
a twin-orientated region.
This structure has been verified experimentally, as demonstrated by the high resolu-
tion transmission electron microscopy image in Fig. 9.27(b). The black dots indicate
the match between the positions of atoms near the interface in this image and those
in Fig. 9.27(a). It can be seen from Fig. 9.27(a) that the atoms in the two atomic
planes labelled S that traverse the step have to shuffle as the dislocation glides along
the boundary, because atoms such as 1 and 2 are closer than 2 and 3 on the left
whereas 2 and 3 are closer than 1 and 2 on the right. The shuffles are short and
easily achieved in this particular case, and so the core of the dislocation spreads
along the interface. Computer simulation shows that this twinning dislocation
moves easily. For boundaries where complex shuffles are necessary, the glide of
twinning dislocations can require relatively high stress and the assistance of elevated
temperature.
Figure 9.27. (a) Atomic structure obtained by computer simulation of the structure
of a twinning dislocation in a twin boundary in titanium, an hexagonal-close-packed
metal. Unit cells are shown in outline and the position of the boundary is indicated
by a dashed line. The twinning dislocation, defined by the lattice vectors t and tμ,
has a very small Burgers vector, but requires shuffling of atoms in the layers labelled
S. (After Bacon and Serra (1994), Twinning in Advanced Materials, eds. M. H. Yoo and
M. Wuttig, p. 83. The Minerals, Metals and Materials Society (TMS).) (b) Experimental
HRTEM image of a boundary in titanium containing a twinning dislocation. The
dashed lines show the location of the interface and the dots indicate the positions
of some atoms near the interface.(From Braisaz, Nouet, Serra, Komninou, Kehagias
and Karakostas, Phil Mag. Letters 74, 331 (1996), with permission from Taylor and
Francis Ltd (http://www.tandf.co.uk/journals).)
Dislocations
V. Vitek, in Encyclopedia of Condensed Matter Physics, 2005
Dislocation Cores
It has already been mentioned above that every crystal dislocation possesses a core
region in which the linear elasticity does not apply and the structure and properties
of the core can only be fully understood when the atomic structure is adequately
accounted for. When a dislocation glides, its core undergoes changes that are the
source of an intrinsic lattice friction. This friction is periodic with the period of the
crystallographic direction in which the dislocation moves. The applied stress needed
to overcome this friction at 0 K temperature is called the Peierls stress and the
corresponding periodic energy barrier is called the Peierls barrier.
In general, the dislocation core structure can be described in terms of the relative
displacements of atoms in the core region. These displacements are usually not
distributed isotropically but are confined to certain crystallographic planes. Two
distinct types of dislocation cores have been found, depending on the mode of the
distribution of significant atomic displacements. When the core displacements
are confined to a single crystallographic plane, the core is planar (a more complex
planar core, called zonal, may be spread into several parallel crystallographic planes
of the same type). Dislocations with such cores usually glide easily in the planes of
the core spreading and their Peierls stress is commonly low. In contrast, if the core
displacements spread into several nonparallel planes of the zone of the dislocation
line, the core is nonplanar, extended spatially. The glide planes of dislocations with
such cores are often not defined uniquely, their Peierls stress is high, and their glide
well below the melting temperature is enabled by thermal activation over the Peierls
barriers.
where K are constants depending on the elastic moduli and orientation of the
dislocation line. On the atomic scale, the displacement across the plane of the core
spreading causes a disregistry that leads to an energy increase. The displacement
u produces locally a generalized stacking fault, and in the local approximation the
energy associated with the disregistry can be approximated as
where is the energy of the -surface for the displacement u. The continuous distri-
bution of dislocations describing the core structure is then found by the functional
minimization of the total energy with respect to the displacement u. When the
displacement vector is always parallel to the Burgers vector, the Euler equation
corresponding to the condition leads to the well-known Peierls equation
where u is the displacement in the direction of the Burgers vector and K is a constant
depending on the elastic moduli and orientation of the dislocation line; in the
isotropic elasticity K is given earlier. In the original Peierls–Nabarro model, ∂ /∂u
was taken as a sinusoidal function and the Peierls equation then has an analytical
solution
where 2 measures the width of the core (for , the disregistry is greater than half its
maximum value) and it is proportional to the separation of the crystal planes into
which the core spreads. The Peierls stress evaluated in the framework of this model
is
where μ is the shear modulus and a constant of the order of 1. This stress is
very small even when the core is very narrow; for example, if , . This is a general
characteristic of the planar cores and this is the reason why in f.c.c. metals, in which
dislocations possess planar cores, the Peierls stress is practically negligible.
The nonplanar cores can be divided into two classes: crosslip and climb cores. In the
former case, the core displacements lie in the planes of core spreading, whereas in
the latter case, they possess components perpendicular to the planes of spreading.
Climb cores are less common and are usually formed at high temperatures by a climb
process. The best-known example of the crosslip core is the core of the 1/2 1 1 1
screw dislocations in body-centered cubic (b.c.c.) metals that is spread into three
planes intersecting along the 1 1 1 direction. Figure 3 shows two alternate
structures of such core. The core shown in Figure 3a is spread asymmetrically into
the,, and planes that belong to the [1 1 1] zone and is not invariant with respect to the
diad and another energetically equivalent configuration, related by this symmetry,
operation exists. The core in Figure 3b is invariant with respect to the diad. These
core structures were found by atomistic calculations employing (1) central-force
many-body potentials and (2) a tight binding and/or density-functional theory based
approach, respectively. Atomistic calculations employing appropriate descriptions
of atomic interactions are the principal tool for investigations of such cores since
direct observations are often outside the limits of experimental techniques. The
example presented here shows that the structure of such cores is not determined
solely by the crystal structure but may vary from material to material even when
they crystallize in the same structure. The Peierls stress of dislocations with these
cores is commonly several orders of magnitude higher than that of dislocations with
planar cores. Furthermore, the movement of the dislocations is frequently affected
not just by the shear stress parallel to the Burgers vector in the slip plane but by
other stress components. Thus, the deformation behavior of materials with non-
planar dislocation cores may be very complex, often displaying unusual orientation
dependencies and breakdown of the Schmid law. The nonplanar dislocation cores are
the more common the more complex is the crystal structure, and thus these cores
are more prevalent than planar cores. In this respect, f.c.c. materials (and also some
hexagonal close-packed (h.c.p.) materials with basal slip) in which the dislocations
possess planar cores and, consequently, the Peierls stress is very low, are a special
case rather than a prototype for more complex structures.
Figure 3. Two alternate core structures of the 1/2 [1 1 1] screw dislocation depicted
using differential displacement maps. The atomic arrangement is shown in the
projection perpendicular to the direction of the dislocation line ([1 1 1]) and circles
represent atoms within one period. The [1 1 1] (screw) component of the relative
displacement of the neighboring atoms produced by the dislocation is represented
by an arrow between them. The length of the arrows is proportional to the magnitude
of these components. The arrows, which indicate out-of-plane displacements, are
always drawn along the line connecting neighboring atoms and their length is
normalized such that it is equal to the separation of these atoms in the projection
when the magnitude of their relative displacement is equal to |1/6 [1 1 1]|.
Using modern synchrotron X-ray sources from the late 1990s, it has been possible to
study complex dynamic behaviour of proteins using Laue techniques. The structures
of the intermediates and transient states of enzymes involved in their catalytic cycles
can be identified and the respective time scale 100 ps to 100 ms can be assigned. The
analysis of the bacterial photoreceptor photoactive yellow protein revealed that by
taking snapshots at every 150 ps throughout the photocycle of 10 ns to 100 ms, five
distinct intermediate structures could be identified and used to establish a reason-
able chemical kinetic mechanism.28 Consequently the chemical mechanism of the
process can be formulated.29 Recent advances made possible sampling picosecond
dynamics of the protein by 100–150 ps time resolution of measurement.30–33 New
frontiers include 5D time and temperature dependent studies34 as well as developing
new techniques to further shorten time scale of the experiments.35 Time-resolved
Laue studies require photoreaction-initiated systems for triggering the reaction
uniformly in the whole crystal volume rapidly with respect to the process under
study to explore dynamics of the sub-ns–ms range in these systems. For systems
presenting a significantly slower motion of s–h time scales initiation is not restricted
to fs photo reactions, but it can be carried out by for example substrate diffusion.
Intermediates of these systems can be freeze-trapped, so there is no need for using
Laue techniques.36,37
Figure 5.10 shows the microstructure of an iron-0.2% carbon alloy after tempering
at 700°C for 2 hours. spheroidized cementite particles can be resolved in the light
microscope, and coarse blocks of structure, differentiated by etching differences, still
in the parallel packet alignment of lath martensite, characterize the microstructure.
The fine martensitic crystals within the blocks have been eliminated, a change in
microstructure attributed to recovery mechanisms associated with the low angle
dislocation boundaries that separated parallel martensite crystals with the same
orientation. With increased tempering, the large angle parallel boundaries rearrange
to form equiaxed ferritic grains to minimize grain boundary energy. Thus the
equilibrium microstructure of spheroidized carbides and equiaxed ferrite grains in
highly tempered high purity iron-carbon alloys and low-alloy steels may be a result
of a sequence of recovery and grain growth mechanisms (Caron and Krauss, 1972;
Hobbs et al., 1972). Recovery mechanisms, involved not only in the elimination of the
low angle martensite interfaces but also in the reduction of intracrystal dislocation
densities, apparently lower strain energy to below the critical amount necessary for
recrystallization.
Mild steel will normally be covered with millscale which, over a period of time,
loosens and may fall away. It can contribute to corrosion and presents an unsound
basis for coating. Techniques for removal and preparation are well documented
[168]. Cast iron has a more adherent scale with some protective value; it corrodes at
a similar rate to mild steel, though the residue is less obviously coloured. Wrought
iron is generally similar to mild steel, though the corrosion rate may differ. Iron and
steel will frequently be coated with less corrodible non-ferrous metals such as zinc
or aluminium, which have other characteristics as described below.
Abstract
This paper reviews the current state of the art in the simulation of the mechanical
behavior of polycrystalline materials by means of computational homogenization.
The key ingredients of this modeling strategy are presented in detail starting with
the parameters needed to describe polycrystalline microstructures and the digital
representation of such microstructures in a suitable format to perform computation-
al homogenization. The different crystal plasticity frameworks that can describe the
physical mechanisms of deformation in single crystals (dislocation slip and twinning)
at the microscopic level are presented next. This is followed by the description of
computational homogenization methods based on mean-field approximations by
means of the viscoplastic self-consistent approach, or on the full-field simulation
of the mechanical response of a representative polycrystalline volume element by
means of the finite element method or the fast Fourier transform-based method.
Multiscale frameworks based on the combination of mean-field homogenization
and the finite element method are presented next to model the plastic deformation
of polycrystalline specimens of arbitrary geometry under complex mechanical load-
ing. Examples of application to predict the strength, fatigue life, damage, and texture
evolution under different conditions are presented to illustrate the capabilities of the
different models. Finally, current challenges and future research directions in this
field are summarized.
Crystal struc- Slip plane Slip direction Number of Slip direc- Number of slip
ture nonparallel tions per systems
planes plane
f.c.c. {1 1 1} 4 3 12=(4×3)
{1 1 0} 6 2 12=(6×2)
b.c.c. {1 1 2} 12 1 12=(12×1)
{1 2 3} 24 1 24=(24×1)
{0 0 0 1} 1 3 3=(1×3)
h.c.p. 3 1 3=(3×1)
6 1 6=(6×1)
h.c.p. Cd, Zn
Dislocations are crystal defects and, therefore, cause a distortion of the atomic
arrangement, which manifests itself into a long-range stress field. Via their stress
field, dislocations interact with each other. This constitutes a glide resistance, which
is macroscopically felt as a yield stress and eventually causes immobilization of
moving dislocations. The traveling distance (slip length) of a moving dislocation
is normally much smaller than the macroscopic dimension of the deformed body;
therefore, dislocations are stored in the crystal and the dislocation density grows
with progressing deformation which in turn increases dislocation interaction and
thus the flow stress. The relation between flow stress and dislocation density can
be represented by the Taylor equation
[1]
where the dislocation density (m−2) is defined as the total dislocation line length
per unit volume – typically 1010m−2 for annealed and 1016m−2 for heavily deformed
metals, b is the Burgers vector (slip vector), G is the shear modulus, and M is
the Taylor factor which relates the applied stress to the shear stress in the glide
system and depends on crystallographic texture (see below). Typically, M 3, and the
geometrical constant 0.5.
Any change in the glide resistance of the moving dislocations will affect the flow
stress. This is the reason for work hardening (increasing dislocation density), other
strengthening mechanisms (grain size, solute atoms, particles, precipitates, etc.),
and softening by recovery and recrystallization (decreasing dislocation density). It is
important to note that the macroscopic deformation behavior can be related to the
microscopic properties of the crystal dislocations. The force K on a dislocation due
to a stress applied to a crystal is given by the Peach–Koehler equation
[2a]
or the magnitude of the force K in the slip system due to the resolved shear stress
[2b]
The imposed strain rate is reflected by the dislocation velocity through the Orowan
equation
[3]
where m is the density of the mobile dislocations moving with an average velocity
. The velocity, of course, depends on the applied stress . One can imagine the
dislocations as rigid rods, moving with velocity due to a force K, and this process
is experienced macroscopically as a plastic deformation with strain rate at a flow
stress .
(3a)
n is the unit vector normal to P; ikl is the Levi-Civita symbol whose nonzero
components are
(3b)
(4)
Using the coordinate axes defined in Fig. 1, the components of a are written as
(5)
A detailed account of the calculation and use of SDDTs adapted to this specific
problem is found else where (see Planar Dislocation Arrays and Crystal Plasticity).