You are on page 1of 16

Dislocation Dynamics

1 Introduction
Dislocation Dynamics (DD) is a computational tool used to study the dynamic collective
evolution of dislocations in a solid under different loading at the mesoscopic scale. It is a
modelling approach for the study of crystal plasticity in which individual dislocation lines are
discretized and their motion in the crystal is simulated. In this simulation technique, each
dislocation is represented as an elastic body, which follows certain rules for motion and
interaction with other dislocations, such as drag force upon gliding, cross-slip, junction
formation, etc. DD simulations were used to study the role of the dislocation microstructure
in strain hardening, low-strain fatigue and crack-tip plasticity. Discrete dislocation dynamics
average out the atomistic nature of material by lowering direct atomic interactions to linear
continuum elasticity. In this context, they calculate exact positions and velocities of all
dislocations segments at each instant.

2 Literature review
The elastic theory of dislocation dynamics is used to study the plastic properties of crystals
by direct simulation of the statics and dynamics of discrete dislocations appeared first during
the mid-1960s, later it is reconnected by several authors [1]. In the late-1980s, however, the
term discrete dislocation dynamics (DDD) was used in relation to two-dimensional computer
simulations aimed to give some idea on the physical origins and the condition for dislocation
pattern evolution [2]. Initial attempts were made to explain thin film plasticity using DD
assumed two-dimensional (2D) models. The concept of 3D discrete dislocation simulations
was introduced by L. Kubin, Y. Bréchet and G. Canova in the early 1990s [KUB 92], [DEV
92][3]. The first code Micromégas was a simple model in which dislocation lines of a FCC
single crystal are divided into sets of edge and screw dislocation segments embedded in a
continuum medium.

Plastic deformation in crystalline metals is a result of nucleation, multiplication and motion


of a large number of dislocations. Discrete dislocation dynamics (DDD) method (Van der
Giessen and Needleman, 1995; Kubin and Devincre, 1999; Ghoniem et al., 2000; Devincre et
al., 2001; Zbib and Diaz de la Rubia, 2002; Arsenlis et al., 2007; Liu et al., 2009) has been
developed to calculate the plastic deformation directly from the evolution of a large number
of dislocation segments [8-9]. The three dimensional DDD method was originally developed
based on the discretisation of dislocations of arbitrary shapes into a succession of screw and
edge piecewise segments. Zbib et al. (1998), stated that dislocations were treated as a
piecewise continuous array of straight line segments with mixed characters [24]. On the other
side, Ghoniem et al. (2000, 2002) developed a parametric dislocation dynamics (PDD) model
by denoting dislocations as spatial spline curves connected through dislocation nodes [11]. In
order to improve the computational efficiency in the study of large-scale plastic deformation,
Wang et al. (2006) presented a parallel algorithm for a 3D PDD [12]. Discrete dislocation
dynamics becomes a popular method to study the intrinsic mechanism of mechanical
responses of materials (Madec et al., 2003; Madec and Kubin, 2008)[13]. In addition, DDD
can predict the experimentally observed plastic anisotropy and tension–compression
asymmetry by introducing atomistic information (Wang et al., 2009; Wang and Beyerlein,
2010). Discrete DD models dislocation slip, multiplication, annihilation, jog, junction, cross
slip (Rhee et al., 1998; Madec et al., 2002; Wang et al., 2007)[14] and climb (Gao et al.,
2010) by a series of constitutive laws for materials, thus it can capture the interactions
between dislocations on different slip systems explicitly (Queyreau et al., 2009)[15], the
interactions between dislocations and internal microstructures as well as the formation of
different heterogeneous dislocation microstructure such as slip band (Kubin et al., 1992;
Kubin and Devincre, 1999;Wang et al., 2008) under various loads. The dislocation-internal
microstructure interaction and the formation of dislocation microstructures are important to
understanding the deformation mechanisms, DDD has become an important method to study
the mechanical behaviour of materials and its intrinsic dislocation mechanisms for both
monotonic (Motz et al., 2008)[16] and cyclic loadings (Déprés et al., 2004, 2006). For
monotonic loading, Groh et al. (2005) performed a three dimensional DDD simulation of the
stress–strain response of particle-reinforced metal matrix composites and found that the
constraint effect on dislocation mobility by the matrix channel width significantly influences
its mechanical response. For cyclic loading, Shin et al. (2007) studied the modelling of the
fatigue behaviour of shearable particle hardened materials by the DDD method. His results
showed that the particle size influences the manner of interaction between dislocations and
particles, further the cyclic behaviour. Recent 3-dimensional DD simulations have some
significant advantages over the aforementioned 2-D DD simulations; for example, the local
interactions between dislocations are accounted for, and the motion of dislocations, those that
interact with the free surfaces of the microcrystal, can be more accurately modelled. In initial
studies, the set of isolated Frank-Read sources (FRs) with ends fixed rigidly was considered
as the starting dislocation populations. Tang et al., using a fixed number of Frank-Read (FR)
sources as the initial condition, stated that dislocation escape through free surfaces plays an
important role in the size dependence of the plastic response of single-crystals. El-Awady et
al. demonstrated the impact of the weakest dislocation sources in samples and cross-slip
cause further strengthening and discontinuous on the stress-strain curves. The study done by
Tang differed that the initial source distribution is not FRs predefined and rather used
artificially generated jogged dislocations as starting dislocation populations for their
simulations while neglected the boundary conditions and cross-slip. He stated that sources
shut-down causes staircase flow stress at 0.2% plastic deformation scaled with specimen size
with an exponent between -0.6 and -0.9, depending on the initial structure and size regime.
There have been still debates regarding that that as most pillars have been made from well-
annealed single crystals or sputtered thin films which do not involve such high densities of
dislocation interactions. Despite these progresses and various studies, there are still many
questions that are unanswered regarding the plasticity at small scales, how the image stresses
induced by free surface and how confined geometries influence multiplication of dislocation
sources and the effect of crystal orientation (multi-slip versus single slip).
Dislocation dynamics simulations of plasticity in thin films

The use of dislocation dynamics (DD) simulations has shown to be powerful approach for
studying the mechanical properties of crystals, including dislocation interactions, dynamics
and microstructures. Initial efforts were made to explain thin film plasticity using DD
assumed two-dimensional (2D) models. Nicola and his co-worker’s conducted a series of 2D
simulations on polycrystalline thin films. They concluded that the yield strength of
freestanding thin films is nearly independent of film thickness and that the size effect results
from dislocation pile-ups at grain boundaries and passivation layers (Nicola et al., 2003,
2005, 2006)[6]. However, the 2D models cannot accurately capture the three-dimensional
nature of microstructural development in materials thus; they are unlikely to provide a good
description of thin film deformation. Hartmaier modelled polycrystalline films by
incorporating dislocation climb in their 2D simulations. It was found that the dislocation slip
mechanism is dominant in thicker films. The creep mechanism found in ultra-thin films with
thicknesses below 400 nm (Hartmaier, 2005)[17]. Han et al. investigated surface-induced size
effects using 2D simulations with results indicating that a free surface might act either as a
net dislocation source or as a dislocation sink, thus inducing harder and softer deformation
behaviours (Han et al., 2006)[18]. However, 2D models cannot accurately capture the three-
dimensional nature of microstructural development in materials and are unlikely to provide a
good description of thin film deformation. Three-dimensional DD simulations can, be used to
accurately model evolving microstructures in thin film mechanical behaviour. In 3
dimensional DD simulations, every dislocation configuration is divided into a succession of
elementary segments, which move under external forces in discrete steps and generate more
realistic dislocation structures.. Pant et al. (2003) employed 3 dimensional DD simulations to
study the interaction of threading dislocations in face-centered cubic (FCC) metal films [20].
It is found that different aspects of the dislocation interactions dominated the film behaviour
in wide ranges of film thickness and applied strain. Von Blanckenhagen (2004) studied the
plastic deformation of polycrystalline FCC metal thin films by doing simulation of the
dynamics of discrete dislocations in a representative columnar grain [21]. Their result showed
an inverse dependence of the flow stress on film thickness. The dependence of the hardening
rate on film thickness was produced by using an initial dislocation density of sources that was
independent of grain dimensions. Espinosa et al. (2005, 2006) assumed all dislocation
sources were situated at grain boundaries in their 3 dimensional DD simulation and proposed
a new interpretation of size scale plasticity of thin films based on the probability of activating
grain boundary dislocation sources [19]. Recently, Fertig Iii and Baker (2010) conducted 3D
DD simulations on single crystal thin films and demonstrated that weak dislocation
interactions can survive at high stress levels, owing to the inhomogeneity in the stress fields
in the film. The mean free path for dislocation motion was closely correlated to the
inhomogeneous stress distribution. None of previous, DD simulations on thin films
considered stress relaxation mechanisms, such as cross-slip of dislocations and dislocations
transmitting at grain boundaries, and therefore it is unclear how these dynamic behaviours of
dislocations will affect mechanical properties of polycrystalline thin films. To create simple,
accurate models that can be used to predict film behaviour, we must first identify the critical
features in the plastic deformation of polycrystalline thin films, which can be studied by full
3D DD simulations that include basic dislocation mechanisms. In this study, 3 dimensional
DD simulations that include both dislocations cross slip and stress relaxation at grain
boundaries have been used to investigate the size-dependent plasticity of polycrystalline thin
films. Based on the results, the plastic deformation of polycrystalline thin films are related to
quantities as dislocation density, grain size and thin film thickness, and finally develop a
model that describes well the plastic behaviour of thin films.

Basics of Dislocation Dynamics


In Discrete dislocation simulations, the dislocations are the simulated entities, can offer a way
to extend length scales beyond those of atomistic simulations. dislocation-based simulations
(i) represent the dislocation line in some way, ii) determine the forces or interaction energies
between dislocations, and (iii) calculate the structures as well as response of the dislocations
to external stresses. These simulations are useful for mapping out the underlying mechanisms
by providing data not available experimentally on, for ex., dislocation ordering, evolution of
large-scale dislocation structures (walls, cells, and pileups), dynamics (instabilities), etc. For
the micron-scale systems, recent DD simulations determine the size-affected mechanical
response.

In Dislocation Dynamics simulation technique, in order to simulate the motion and


interaction of dislocation lines, a number of rules, algorithms, and procedures have been
developed. In the methodology, we will see following aspects: how driving forces are exerted
on dislocations (3.2.1), how to determine dislocation velocities given these forces (3.2.2),
discretization and adaptive remeshing of the dislocation lines (3.2.3), time integration of the
equations of motion (3.2.4).

3 Methodologies:

3.1 Problem formulation:


The basic idea of Dislocation Dynamics is to include the physics of dislocations into a set of
governing equations which can be solved for the positions of a network of dislocation lines.
For given a dislocation configuration, loading conditions and boundary conditions, The
positions of the lines are described by the vector r(s,t), where s is a scalar parameter that tells
the location along the lines, and t denotes the time. We will need to discretize the system in
both space and time, and apply numerical methods to solve the governing equations. Fig. 1
shows an example of discretization in space:
Fig. 1 Position of a pair of dislocation loops at specific time t. (a) Continuous representation
described by position as function of parameter s. (b) Discrete representation using node-based
discretization. [23]

The forces acting on dislocation can be divided into two forces, i,e, drag forces, which resist
dislocation motion, and driving forces, which promote it. Dislocation lines are known to have
effective masses, which leads to inertial forces. In crystalline materials, drag forces intrinsic
to the crystal lattice are orders of magnitude larger than the inertial forces under a wide range
of conditions and making dislocation motion over-damped. This means that in the overall
equations of motion, we can neglect inertial terms together, and simply require that the total
driving force balance the total drag force, that is:

∑ Fdrag (v,s) + ∑ Fdrive(s) =0

Where v is the dislocation velocity

The equation 1 is explicitly solved for v and restated as above

V= M (Ftotdrive)

where Ftot drive[r(s), σext,...] = ∑Fdrive is the total driving force as a function of parameter
which depend upon the dislocation position r(s), the externally applied stress σext, and any
other features which exert driving forces. The function M(.), which gives the velocity for a
given total driving force, is called the mobility law. The final governing equation of motion
can be written as

Where g= M(Ftot drive[r(s), σext,...] )is an operator which calculates the velocity v(s) from a
given dislocation structure r(s) and loading condition.
3.2 Flowchart for Dislocation dynamics Simulation

Fig 2: Flowchart showing the basic steps for a dislocation dynamics code[23]

Now we will discuss each step in detail.

3.2.1 Driving forces:


These forces can be divided into two categories: forces arising from local stress fields (Peach-
Koehler forces) and forces due to the energy of the dislocation core.

The force acting on dislocation is given by

F(s) =( σ(s)·b* eta(s)) …. (Peach-Koehler forces)

The most common sources of stress in dislocation dynamics simulations are applied stresses
due to loading of the simulation cell and stresses from other dislocation. Other sources
include precipitates or inclusions, and free surfaces or secondary phase boundaries. The core
forces are calculated from core energy. The core forces reduce the length of dislocation and
will exert torque on dislocation line. The core energy is calculated as above

Where n is Poisson’s ratio, b is the Burgers vector.

3.2.2 Mobility Law


Mobility laws serve as constitutive equations in dislocation dynamics simulations, relating
the total driving force per unit length acting on a dislocation line to its velocity. It is modelled
by a phenomenological law which prescribes the dependence of the steady-state velocity of a
dislocation on local parameters, such as the stress, temperature, the line character (edge or
screw) and the slip system.

Where B(s) is drag coefficient tensor and is strongly material dependent.

3.2.3 Line discretization and Remeshing


To apply any numerical methods, we need to discretize the dislocation lines so that the
overall dislocation structure is characterized by a set of nodes (or segments) and a data
structure defining the connectivity between them. Discretization enables us to focus on a
finite number of degrees of freedom (DOF). Since the dislocation lines can change their
shape significantly during a simulation, and the total length of dislocation lines increases in
most cases so we need to implement remeshing in the algorithms to modify the discretization
when necessary. Dislocation lines are discretized in a number of ways. There are two types of
line discretization one is lattice-based discretization and other is node-based discretization.

In the lattice-based approach (used in the codes microMegas [23] and TRIDIS [102]), a grid
of computational points, i.e. a lattice, often with a simple cubic structure of spacing a, is
defined throughout the simulation cell. Based on the lattice structure, a finite set of
dislocation orientations is selected and their orientations are considered for the study. These
orientation states a unique set of straight line segments used to represent the dislocation lines.

Dislocation lines are discretized into a set of nodes and shape functions that connect the
nodes. The simplest case is being linear shape functions that result in straight line segments in
the node-based approach. In this approach, any dislocation orientation is allowed and
dislocation segments can move in any direction (consistent with their mobility law). In the
node-based approach the nodes are the fundamental degrees of freedom.
The segments are the fundamental degree of freedom in lattice-based models; we need to
calculate the total force acting on a segment with the line integral:

In Node-based codes, , require the total force acting on the nodes. This is determined in terms
of the line shape function Nji (s) which describes the contribution to node i from segment j as

The total force on node i is then the sum of the contributions from each of the segments it is
attached to

Fig. 3: Figure of (a,b) lattice-based and (c) node-based discretization. (a) The lattice grid
showing the segment directions ti and movement directions di. (b) Remeshing when a
segment more than twice the average length l, and. (c) Nodes are inserted when l > lmax or A
> Amax, and removed (unfilled circles) when l < lmin, or A < Amin with the area shrinking
(dA/dt < 0) [23]

3.2.4 Time integration


The dislocation line motion is governed by a partial differential equation (PDE) in time. We
can write this governing equation in terms of the motion of the nodes or segments and
convert the PDE into a coupled system of N ordinary differential equations (ODEs). In the
nodal representation we have

{ }
3.2.4.1 Subcycling
Generally, DD configurations arise are both stiff and nonlinear. Time step subcycling is used
to handle these modes by isolating them into a separate group for which small time step sizes
are used and the rest of the system continues to be time integrated with the larger time step
size. The basic approach to subcycling is explained in the following steps, The operator g(·)
includes variety of interactions, out of which some are nearly-linear and slowly varying and
others are nonlinear and rapidly varying in time. This operator can be “split” in any manner
of choosing into n + 1 operators such that

Group 0 is corresponding to the g0 operator, is the global group which contains the smooth
and slow modes. The other operators contain the nonlinear and rapid modes and are grouped
according to their severity. By splitting the operator in this way allows modes with
comparable time step sizes to be integrated together and maximizing efficiency. A solution is
then obtained by time integrating each of the operators independently and their solutions
coupled in the splitting scheme. The error incurred is known as the splitting error. When
designing a sub cycling algorithm, one has to identify how to split the system and what types
of integrators to use on the new operators. Consider the case in which the system is split into
two groups of nodes, with g1 contains nodes experiencing nonlinear interactions and g0 the
balance of the system. The implicit integrator yielded the best performance with this splitting
scheme. In force based subcycling algorithm, we will split g in terms of forces. Generally,
the short-ranged forces are most stiff and nonlinear that means they will require the smallest
time step sizes. Therefore, we can divide up these forces into groups based on the length scale
over which they act, which allows us to choose the time step size for each group as they are
time integrated separately. In this way, we can separate the modes that require small time step
sizes from the modes that allow large time step sizes.
Fig 4: Flow chart for force-based subcycling algorithm [27]

3.2.5 Boundary conditions


The boundary conditions (BCs) are needed to define in order to have a well-defined problem.
The form of the BCs is determined by the geometry we considered for the study. The types of
BCs used in DD simulations are divided into three groups as shown in Fig. 6: (a) infinite
BCs, (b) periodic BCs, and (c) heterogeneous BCs.

Fig 5: Different types of boundary conditions: a) Infinite BCs b) Boundary & heterogeneous
BCs c) free standing films with in plane periodic BCs d) A biomaterial interface in an infinite
medium [23]
Loads and boundary conditions

In solid mechanics area, generally in simulations (and experiments), two types of loadings in
DD are used and are as follows: stress-controlled and strain-controlled. Under stress control,
the stress state is specified and the dislocation lines respond to the resulting forces. The stress
state vary in time or it can be constant with time.

Under strain-control, usually a strain rate tensor, eij is specified and the resulting stress state
must be calculated as follows. The total strain at time t is

3.3 Types of simulations


DD simulations are considered as two types, one is small-scale which include the interactions
and behaviour of one or a few dislocation lines and other is large-scale simulations examining
the collective behaviour of many dislocations. Small-scale simulation includes the simulation
of intersecting dislocation lines, junction dissolution and junction formation, the interaction
of dislocations with precipitates and solutes and the interaction of dislocations with free
surfaces. Large scale simulation includes simulations of large-scale collective behaviour
involve simulating the stress-strain response of a material, which includes work hardening in
bulk metals as example.

3.4 Inputs parameters and Outputs of Dislocation dynamics Simulation


Inputs Output
Elastic properties Stress-strain curve
Cryatalographiic properties Total dislocation density
Dislocation Dislocation density per slip systems
Reaction between dislocations Dislocation reactions: junction, jogs, dipoles
Cross-slip properties Mobile dislocation
Dislocation sources Forest dislocation
Loading conditions

4.1 Case study 1: Creep test during compression of micropillars


Considering the case to simulate the collective dislocation behaviour in the presence of a
cluster of particles, to study the emergent interactions and larger-scale-patterning in BCC
iron, the simulations are performed within α-Fe single crystal micropillar. The size of the
simulation box is 2µm × 2µm × 4µm,in which the bottom surface fixed. The compression
load is applied on the top surface along the [001] direction. The four sides are free. The
dislocations can exit freely from all the surfaces except the bottom. Non-deformable particles
are randomly distributed inside of the box. The volume fraction of particles is considered as
0.02. The particles are mono-sized spheres with radius R = 80 nm. The misfit strain is given
as ε = 0.001. Three prismatic loops are stated as the initial dislocation configuration, which
are dispersed randomly inside of the simulation cell and assigned with different slip planes.
Two sides of each dislocation loop are pinned, and are indicated by the black segments
shown in figure 6b, so that each prismatic loop provides a pair of Frank Read source (FRs).
The initial FRs are located on [110] (blue segments), [101] (purple segments) and [011]
(blue-green segments) slip planes, respectively. The length of the FRs are given as L = 400 ±
50 nm. Dislocations are allowed to cross-slip on all planes of the {110} <111> slip system at
temperature T = 900 K. The α-Fe is approximated to be elastically isotropic. The dislocation
behaviour during the creep test under different compression stresses is simulated and studied.
In each simulation, the applied stress is increased from 0 to the specified stress level σ33
slowly, and then held constant, as given in figure 5b, to investigate dislocation.

Fig 6: Simulation results under a compressive stress of 180 MPa: (a) plastic strain–time curve and
dislocation density–time curve; (b) images showing typical dislocation configurations during the
evolution corresponding to the marked dots in (A). [22]

4.2 Case study 2: Bulk Plasticity Simulation


During plastic deformation the dislocation density tends to increase, because of that the
material strength also increases. This behaviour is called strain-hardening. The study of work
hardening is a key research area for DD simulations.. For this case, work hardening
simulations we use a 10 *10 * 10 µm simulation cell, applying periodic boundary conditions
in all directions. No cross-slip was allowed. The remesh parameter lmax (maxSeg) was 1.25
µm (5000b) (other remesh parameters were given to defaults). We start with 50 straight
dislocation lines with a 60 dergree character angle, random glide plane and random burger
vector with shown in figure.

Fig. 7: The work-hardening simulation performed on nickel at 300 K. (a) Initial


configuration, (b) after relaxation, (c) dislocation microstructure at 0.5% strain with [0 0 1]
uniaxial loading. [23].

We study the response of the system for the uniaxial tension with a constant strain rate of 103
s -1 applied in the [001] and [102] directions. A sub cycling-based time integrator was used,
with simulations run for 40 and 7.2 hours on a single CPU for [001] and [102] loading,
respectively. The resulting dislocation configuration after a total strain of 0.5% in the [001]
direction is shown in Fig. 7(c). Fig. 8 shows the evolution of stress and dislocation density
with respect to total strain. The initial yield strengths are same for both loading directions.
However for the [001] loading the crystal hardens with plastic strain as the dislocation
density increases. Dislocation density and flow stress remain relatively unchanged for the
[102] loading.

Fig. 8: Stress-strain and dislocation density-strain curves for the different loading directions.
[23]
5. Conclusion and Future scope
DD has become a powerful tool for computational modelling of dislocations and crystal
plasticity. It is used for some important applications. In this paper, we reviewed the the
progress of dislocation dynamics for plasticity in polycrystalline thin films. We also reviewed
the stepwise procedure for dislocation dynamics simulation. Two case studies are described
to study inputs parameters applied for simulation and also study the output of DD
simulations.

Some issues in this areas are further required to do research on it, for example dislocation
interactions cannot be calculated analytically in anisotropic elasticity, Most metals exhibit
anisotropy in their elastic behaviour, and but most DD codes use isotropic elasticity to
calculate the interactions between dislocation segments. The reason behind is the analytic
expressions for the stress fields of dislocations in anisotropic media are not known, and their
numerical calculation is very expensive [25]. An approximate method was developed recently
that uses spherical harmonics to estimate the interaction forces between dislocations [5]. With
this approach, the computational cost can be reduced according to the desired accuracy of the
approximation. Another aspect in which is grain boundary, Most DD codes are only able of
simulating single crystals, whereas most structural materials are polycrystalline. The grain
boundaries separating the individual grains of polycrystals can interact with dislocations in
complex ways. Grain boundaries can both absorb and emit dislocations. A grain is under
“misfit” stresses imposed by the surrounding grains during deformation, which exert forces
on dislocations. Dislocations can also be transmitted across grain boundaries. DD simulations
have been run with simplified grain and twin boundary models [26], but a robust DD model
for polycrystals still requires further development.
References
1. A.N. Gulluoglu, D.J. Srolovitz, R. LeSar, and P.S. Lomdahl, Scr. Metall. 23, 1347
(1989)
2. J. Lepinoux and L.P. Kubin, Scr. Metall. Mater. 21, 833 (1987).
3. Kubin, L.P., Canova, G., 1992. The modelling of dislocation patterns. Scripta
Metallurgica et Materialia 27, 957–962
4. N.M. Ghoniem and R. Amodeo, Solid State Phenom. 3&4, 377 (1988).
5. Nicola, L., Van der Giessen, E., Needleman, A., 2003. Discrete dislocation analysis of
size effects in thin films. Journal of Applied Physics 93, 5920–5928.
6. Nicola, L., Van der Giessen, E., Needleman, A., 2005a. Size effects in polycrystalline
thin films analyzed by discrete dislocation plasticity. Thin Solid Films 479, 329–338.
7. Nicola, L., Van der Giessen, E., Needleman, A., 2005b. Two hardening mechanisms
in single crystal thin films studied by discrete dislocation plasticity. Philosophical
Magazine 85, 1507–1518.
8. Liu, Z.L., Liu, X.M., Zhuang, Z., You, X.C., 2009. A multi-scale computational
model of crystal plasticity at submicron-to-nanometer scales. International Journal of
Plasticity 25, 1436–1455.
9. Devincre, B., Kubin, L.P., Lemarchand, C., Madec, R., 2001. Mesoscopic simulations
of plastic deformation. Materials Science and Engineering A 309–310, 211–219.
10. Ghoniem, N.M., Tong, S.H., Sun, L.Z., 2000. Parametric dislocation dynamics: a
thermodynamics-based approach to investigations of mesoscopic plastic deformation.
Physical Review B 61, 913.
11. Ghoniem, N.M., Huang, J., Wang, Z., 2002. Affine covariant–contravariant vector
forms for the elastic field of parametric dislocations in isotropic crystals.
Philosophical Magazine Letters 82, 55–63.
12. Wang, Z.Q., Ghoniem, N.M., Swaminarayan, S., LeSar, R., 2006. A parallel
algorithm for 3D dislocation dynamics. Journal of Computational Physics 219, 608–
621.
13. Madec, R., Kubin, L., 2008. Second-order junctions and strain hardening in bcc and
fcc crystals. Scripta Materialia 58, 767–770
14. Rhee, M., Zbib, H.M., Hirth, J.P., Huang, H., de La Rubia, T.D., 1998. Models for
long/short range interactions in 3D dislocation simulation. Modelling and Simulation
in Materials Science and Engineering 6, 467–492.
15. 15 Queyreau, S., Monnet, G., Devincre, B., 2009. Slip systems interactions in a-iron
determined by dislocation dynamics simulations. International Journal of Plasticity
25, 361–377.
16. Motz, C., Weygand, D., Senger, J., Gumbsch, P., 2008. Micro-bending tests: a
comparison between three-dimensional discrete dislocation dynamics simulations and
experiments. Acta Materialia 56, 1942–1955
17. Hartmaier, A., Buehler, M.J., Gao, H.J., 2005. Multiscale modeling of deformation in
polycrystalline thin metal films on substrates. Advanced Engineering Materials 7,
165–169.
18. Han, C.S., Hartmaier, A., Gao, H.J., Huang, Y.G., 2006. Discrete dislocation
dynamics simulations of surface induced size effects in plasticity. Materials Science
and Engineering A – Structural Materials Properties Microstructure and Processing
415, 225–233.
19. Espinosa, H.D., Panico, M., Berbenni, S., Schwarz, K.W., 2006. Discrete dislocation
dynamics simulations to interpret plasticity size and surface effects in freestanding
FCC thin films. International Journal of Plasticity 22, 2091–2117
20. Pant, P., Schwarz, K.W., Baker, S.P., 2003. Dislocation interactions in thin FCC
metal films. Acta Materialia 51, 3243–3258
21. Von Blanckenhagen, B., Arst, E., Gumbsch, P., 2004. Discrete dislocation simulation
of plastic deformation in metal thin films. Acta Materialia 52, 773–784.
22. F. X. Liu, A. C. F Cocks and E. Tarleton, Dislocation dynamics modelling of the
creep behaviour of particle-strengthened materials (2021).
23. Sills, R.B., Aghaei, A., Cai, W.: Advanced time integration algorithms for dislocation
dynamics simulations of work hardening. In preparation (2015)
24. Zbib, H.M., Rhee, M., Hirth, J.P., 1998. On plastic deformation and the dynamics of
3D dislocations. International Journal of Mechanical Sciences 40, 113–127.
25. Yin, J., Barnett, D.M., Cai, W.: Efficient computation of forces on dislocation
segments in anisotropic elasticity. Modell. Sim. Mat. Sci. Eng. 18, 045013 (2010)
26. Zhou, C., LeSar, R.: Dislocation dynamics simulations of plasticity in polycrystalline
thin films. Int. J. Plast. 30-31, 185-201 (2012)
27. Ryan B. Sills, Amin Aghaei , and Wei Cai (2015), Advanced Time Integration
Algorithms for Dislocation Dynamics Simulations of Work Hardening.

You might also like