ArticlePDF Available

Abstract

A stochastic discrete slip approach is proposed to model plastic deformation in submicron domains. The model is applied to the study of submicron pillar (D <= 1 um) compression experiments on tungsten (W), a prototypical metal for applications under extreme conditions. Slip events are geometrically resolved in the specimen and considered as eigenstrain fields producing a displacement jump across a slip plane. This novel method includes several aspects of utmost importance to small-scale plasticity, i.e. source truncation effects, surface nucleation effects, starvation effects, slip localization and an inherently stochastic response. Implementation on an FFT-spectral solver results in an efficient computational 3-D framework. Simulations of submicron W pillars (D <= 1 um) under compression show that the method is capable of capturing salient features of sub-micron scale plasticity. These include the natural competition between pre-existing dislocations and surface nucleation of new dislocations. Our results predict distinctive flow stress power-law dependence exponents as well as a size-dependence of the strain-rate sensitivity exponent. The results are thoroughly compared with experimental literature.
A stochastic discrete slip approach to microplasticity:
Application to submicron W pillars
Carlos J. Ruestesa,b,, Javier Seguradoc,a
aIMDEA Materials Institute, C/Eric Kandel 2, Getafe, 28906, Madrid, Spain
bInstituto Interdisciplinario de Ciencias asicas (ICB), Universidad Nacional de Cuyo
UNCuyo-CONICET, Facultad de Ciencias Exactas y Naturales, Padre Contreras
1300, Mendoza, 5500, Mendoza, Argentina
cUniversidad Polit´ecnica de Madrid, Department of Materials Science, E.T.S.I.
Caminos, Madrid, 28040, Madrid, Spain
Abstract
A stochastic discrete slip approach is proposed to model plastic deformation
in submicron domains. The model is applied to the study of submicron pil-
lar (D1µm) compression experiments on tungsten (W), a prototypical
metal for applications under extreme conditions. Slip events are geometri-
cally resolved in the specimen and considered as eigenstrain fields producing
a displacement jump across a slip plane. This novel method includes several
aspects of utmost importance to small-scale plasticity, i.e. source truncation
effects, surface nucleation effects, starvation effects, slip localization and an
inherently stochastic response. Implementation on an FFT-spectral solver re-
sults in an efficient computational 3-D framework. Simulations of submicron
W pillars (D1µm) under compression show that the method is capable
of capturing salient features of sub-micron scale plasticity. These include the
natural competition between pre-existing dislocations and surface nucleation
of new dislocations. Our results predict distinctive flow stress power-law de-
pendence exponents as well as a size-dependence of the strain-rate sensitivity
exponent. The results are thoroughly compared with experimental literature.
Keywords: plasticity, simulation, slip localization, stochasticity, tungsten
Corresponding author: Carlos J. Ruestes e-mail:carlos.ruestes@imdea.org
Preprint submitted to International Journal of Plasticity April 17, 2024
arXiv:2404.10430v1 [cond-mat.mtrl-sci] 16 Apr 2024
1. Introduction
Progress towards fulfilling the promise of nanoscale solutions to engineer-
ing problems hinges on the development of computational and character-
ization techniques for an adequate assessment of the mechanical response
of the resulting materials and parts. For instance, micropillar compression
testing (Uchic et al., 2004; Uchic and Dimiduk, 2005) can provide critical
data for the development of a number of technologies at different engineering
scales, from the micro/nanoscale (e.g micro/nano electromechanical systems
-MEMS / NEMS-, microsensors and actuators and microscale energy harvest-
ing and storage systems) to the macroscale. The technique is instrumental
in establishing correlations between the microstructure and the mechanical
properties of materials (Greer and De Hosson, 2011; Dehm et al., 2018). The
analysis and interpretation of the experimental data obtained often relies on
advanced computational simulations at different length scales, which aim at
determining the underlying deformation micromechanisms responsible of the
observed behavior.
At the lowest scale, molecular dynamics simulations (MD) were used to
study the interaction of dislocations with pillar surfaces, leading to the pro-
posal of a surface-induced cross-slip mechanism (Weinberger and Cai, 2008).
In addition, several studies have been published with different focuses, such
as unraveling the tension-compression asymmetry observed experimentally
(Healy and Ackland, 2014) or analyzing the origin of size effects observed
experimentally (Yang et al., 2021; Yaghoobi and Voyiadjis, 2016), to name
a few. Nevertheless, the detailed nature of atomistic simulations, together
with integration timesteps of the order of 1-3 fs, enforces the use of ultra-high
strain rates of 107/s and above and very small pillar sizes (D < 100 nm).
The comparison between MD simulations and micropillar compression ex-
periments is thus limited by the inherent drawbacks associated with the sim-
ulation technique. Still, MD has made fundamental contributions to the un-
derstanding of source nucleation in micropillars and nanowires (Weinberger
and Cai, 2008; Zhu et al., 2008), even allowing to estimate nucleation stresses
on pristine fcc nanowires within 20% of the experimental values (Jennings
et al., 2013).
At the mesoscale, discrete dislocation dynamics simulations (DDD) have
been used to provide insights into the mechanisms responsible for the ex-
perimentally observed staircase stress-strain behavior (Tang et al., 2008) as
well as athermal size-dependent strengthening (Rao et al., 2008). The tech-
2
nique has also been used to study the effect of specimen size on plastic flow
and work-hardening (El-Awady et al., 2009) and also to study the charac-
teristics of source truncation controlled flow behavior for submicron FCC
single crystal micropillars (Cui et al., 2014). Moreover, in order to capture
surface-related effects, MD simulation-derived models of surface nucleation
(Ryu et al., 2011; Zhu et al., 2008) and surface-induced cross-slip (Wein-
berger and Cai, 2008) have been incorporated into DDD frameworks (Ryu
et al., 2013, 2015; Hu et al., 2019), allowing for the exploration of pillar sizes
in the range of 100 nm to a few microns. The technique has also been used to
develop stochastic models for the onset of plasticity in micro- and nano-scale
structures (Shao et al., 2014). In spite of their advantages, DDD simulations
still suffer from timescale limitations (few ns to few µs) and typical strain
rates used are in the range of 103/s 106/s, though strain rates as small
as 0.1/s are possible depending on the computational resources (Fan et al.,
2021). Both techniques, MD and DDD, typically rely on an implementation
in the form of massively parallel solvers to obtain results in a reasonable
time, which in turn impose important hardware requirements.
At the continuum scale, several groups had opted for finite element simu-
lations (FE) with different material model approaches. FE simulations using
simple isotropic plasticity have proven useful in providing recommendations
for experiments design and reliable testing (Zhang et al., 2006; Kiener et al.,
2009). Simulations using crystal plasticity constitutive equations (CPFE)
simulations have been successfully used to study the effects of initial crystal-
lographic orientation, diameter-to-length ratio and friction on the response of
Cu single crystalline pillars (Raabe et al., 2007). CPFE has also been used
for the study of polycrystalline Ni-based superalloy micropillars (Cruzado
et al., 2015). The use of CPFE models has also extended to size effects
in Al pillar compression, focusing on a continuum description of starvation
(J´erusalem et al., 2012). CPFE relies on homogenization of the evolution
of large dislocation ensembles and on scale-separation, hypotheses that can
be perfectly assumed when studying the deformation of large crystal grains
or specimens of several microns. However, in small-scale testing the number
of dislocations in the specimen can be very small (just a few dislocations
in the full specimen), such that their effect on the deformation is localized
in specific planes and show an stochastic nature. These aspects cannot be
homogenized in a continuum-scale law.
In order to fill gaps between DDD and CPFE models several attempts
have been made to extend standard CPFE models to account for some char-
3
acteristic features of small-size tests (i.e. the stochastic nature of the response
and slip-band localization). Regarding the stochastic response, Ng and Ngan
(2008) proposed a Monte Carlo model based on the survival probability and
the burst size vs. stress distributions but with no further connection with the
material structure. Later on, Konstantinidis et al. (2014) proposed a cellular
automaton, based on the gradient plasticity framework of Zhang and Aifan-
tis (2011), to model stochastic effects in pillar compression. In addition, Lin
et al. (2015) presented a crystal plasticity model in which stochasticity was
introduced by defining the plastic strain as composed of a series of strain
bursts whose sizes follow a power-law distribution function and whose rates
are determined by a constitutive equation. More recently, a general stochas-
tic CP formulation based on kinetic Monte Carlo, was proposed in which
the active slip systems were selected based on the strain rates dictated by a
mobility law depending on resolved stress and temperature (Yu et al., 2021).
However, the homogenized nature of the formulation prevents its direct use
in applications where slip localization is expected.
The slip-band localization observed in experiments, resulting from the
activation of discrete slip events in a particular plane, cannot be naturally
captured by CPFE models. Some attempts have been made to emulate this
localization in CP frameworks. Lin et al. (2016) developed a dislocation-
based crystal plasticity model in which, in order to trigger localization in a
single slip band, strain-softening was added to the slip resistance evolution
law and a void was introduced in the center of the pillar to trigger localiza-
tion. More recently, Wijnen et al. (2021) presented a CPFE model adapted
for simulating heterogeneous plastic deformation in single crystals. In this
model, only some regions of the domain under study are modeled using CP
keeping the others as elastic material. These slip areas are sampled from
a probability distribution based on characteristics of the dislocation sources
and plastic localization is accomplished by considering different slip bands
and the weakest link principle (Norfleet et al., 2008). These extended CPFE
models allow to describe the strain localization in pillars but maintain a de-
terministic nature and do not explicitly consider plastic events by individual
dislocations.
Summarizing, the mechanical response of metals and alloys under micro
compression testing is typically characterized by certain features: i) samples
usually have a low dislocation density and exhibit source truncation harden-
ing; ii) Plastic deformation is usually localized and in the form of slip bands;
iii) For diameters below 150 nm, or starved samples such as those produced
4
by mechanical annealing, surface nucleation of dislocations is likely to occur.
As a consequence, plasticity takes place by events of discrete nature and the
mechanical response of such structures is typically stochastic. Despite all the
aforementioned computational advances, capturing all these features into a
single and efficient computationally-cheap 3-D framework is still a challenge.
The purpose of this work is two fold. Firstly, we present a multiscale com-
putational framework capable of capturing the fundamental aspects of the
mechanical response of submicron metallic parts. In this model, the displace-
ment jumps along a cross sectional plane of the sample produced by the slip
of a dislocation are introduced as eigenstrains around that plane using the
Eshelby formalism. The framework also accounts for surface nucleation by
implementing a model derived from MD simulations. The stochastic nature
of the mechanical response is captured by means of a kinetic Monte Carlo
selection process. In this kMC approach, the events correspond to the sin-
gle slip events produced by dislocations, whose probabilities are dictated by
physically-based laws. This approach is implemented in an FFT-based solver
resulting in a very efficient and computationally-cheap framework that allows
to model complex 3D geometries. Secondly, we revisit micropillar compres-
sion experiments of single crystalline bcc submicron pillars (D1µm) with
focus on tungsten (W). We show that not only our method can successfully
capture fundamental aspects of the mechanical response at these scales, but
also allows for detailed investigation of size-effects, strain-rate effects and
statistics of strain-bursts.
2. Methodology
The framework developed assumes dislocation-mediated plasticity as the
underlying inelastic deformation micromechanism. We consider pre-existing
dislocations as well as newly developed dislocations due to surface nucleation.
The amount of pre-existing dislocations Ndon a pillar of diameter D= 2R,
height H, and initial dislocation density ρ, is determined as
Nd=int(ρπ R2H
Ld
) (1)
where Ldis the average length of a dislocation segment, here taken as R
for simplicity. Then, Ndslip planes are randomly selected along the pillar
axis with normal directions consistent with the available slip systems.
5
Surface nucleation can potentially happen at any position on the pillar
surface. From a practical point of view, in order to consider a dislocation
nucleation and its subsequent slip in our framework, a large but finite number
of nucleation sites NSN is considered. In practice, 100 < NSN <200. The
positions of the resulting slip planes are randomly selected along the pillar
axis, and their orientations are also based on the available slip systems.
In our framework, dislocations are not explicitly considered, but the effect
of their extension when reaching the surface of the pillar. This effect consists
in a relative displacement between the upper and lower parts of their plane
and experimentally would define a slip trace. This description is equivalent to
the definition of a closed dislocation loop in phase-field dislocation dynamics
as an inclusion with an eigenstrain (an Eshelby inclusion) (Wang et al., 2001;
Rodney et al., 2003). In our case, only the final stage of the loop is considered
which consists then in a loop with a shape defined by the intersection of the
pillar with the slip plane; see Sec. 2.1 for details. The different parts of the
model are introduced in the next sections.
2.1. Kinematics of a slip event
First we derive the elastic displacement introduced by a dislocation loop
considering it as an equivalent Eshelbian inclusion. Under this representa-
tion, the loop is idealized by moving the upper side of surface S, denoted
by S+, by bwith respect to the lower side S, as shown in Fig. 1.a. Let n
be the normal to the loop, then the relative displacement between the upper
and lower parts of the loop in direction nis given by
JuKn(x) = bδS(x),(2)
where bis the burgers vector of the loop and δS(x) denotes the surface Dirac
delta function, that is infinite if xSand zero elsewhere. If the loop is
fully embedded in a material, the resulting elastic fields of this incompatible
deformation can be obtained using Green’s function (Mura, 2013). In our
model, the only stage considered is when the loop has filled all the pillar
cross section.
The displacement jump due to a full plane slip has been considered in
the past using a strong discontinuity approach in the context of finite ele-
ments and discrete dislocation dynamics (Romero et al., 2008) . However,
introducing this discontinuity in other numerical frameworks as FFT imply
smoothing out the jump to make the model numerically tractable. This can
6
be done by distributing the jump along a thin band of size h, such that it
renders the same total displacement as in the initial ideal case. In a simplified
way, the line integral along a path parallel to ncorresponds to
Zz=h/2
z=h/2
ε(x)dz=JuKn(3)
Figure 1: Dislocation eigenstrain formalism and the application to a single crystalline
pillar. (a) A dislocation loop Lacting as a boundary of a dislocation surface S. By
sweeping the upper part S+with respect to Sin an amount b, a dislocation is effectively
introduced. (b) For a dislocation sweeping a pillar, the dislocation can be replaced by a
plate-like Eshelbian inclusion of thickness h.
Extending the use of the equivalent Eshelbian inclusion, Figure 1.b de-
picts a scenario in which the cross-section of cylindrical pillar has been fully
swept in an amount of bby a single dislocation loop. The associated displace-
ment is linearly distributed over a thickness hin the direction perpendicular
to the loop. As explained later on (Sec. 2.4), the simulation domain is dis-
cretized in a box of [Nx, Ny, Nz] voxels, so in practice, his taken as twice
the distance between voxels (Capolungo and Taupin, 2019). As presented in
the Supplementary Material file, the method is relatively insensitive to the
election of the voxel size. In consequence, the strain εEI G (i)(x) associated
with a slip event i, is taken as:
εEI G (i)(x) = (b
h(sini), if xϵ plastic region
0, elsewhere (4)
7
where bis the burgers vector modulus, his the thickness of the slip
band and siand nistand for the burgers vector direction (slip) and normal
direction to the slip plane, sinibeing the Schmid tensor. Consequently,
for a pillar that underwent Mslip events, the total eigenstrain field εEIG (x)
is the superposition of all the eigenstrains applied,
εEI G(x) =
M
X
i=1
εEI G (i)(x) (5)
2.2. Constitutive laws for the slip events
The laws linking the stress at a given plane with its slip rate will be
presented in next sections.
2.2.1. Pre-existing dislocations
The model assumes that the pillar contains a certain number of disloca-
tions which will drive the slip events. Two types of pre-existing dislocations
are considered, these are single-arm sources and screw dislocation segments.
The former aid in the introduction of source truncation effects, while the
latter allow for source exhaustion effects.
Single-Arm Sources. A dislocation with a pinning point in the interior of the
crystal and a free end on the surface acts as a source of dislocations in small
pillars and is commonly termed as Single-Arm Source (SAS) (Parthasarathy
et al., 2007). In the present model, these sources are randomly distributed
in the pillar volume, assigning to each source a particular slip plane. The
critical resolved stress for activating a SAS is determined by its arm length
λ,
τSAS
CRS S =αSAS µ b
λ+τ0(6)
where αSAS is a source-strength coefficient that considers the nature of
the source, its length and material, here taken as 0.6 after (Rao et al., 2007).
µis the shear modulus, bthe Burgers vector and τ0the friction stress. In
their presentation of the SAS model, Parthasarathy et al. (2007) include a
Taylor-type term in eq. 6 to account for dislocation forest interactions. This
term is not considered here since, for D 1µm specimens and typical dislo-
cation densities of the order of 1012m21013m2, the number of dislocations
present in the pillar becomes too small to homogenize their effect (Nd10).
8
Under such conditions, weak interactions are expected. Thus, we essentially
neglect hardening. For pillar diameters above 1 micron, hardening should be
incorporated, as explained in Sec. 4.
In order to generate a random arrangement of SAS, for each source a
random pining point is generated in the pillar interior. Then, a slip plane
among the characteristic active planes of the metal lattice is randomly as-
signed. The source is defined as a straight line of length λconnecting the
pinning point with the nearest point in the elliptical boundary resulting from
the intersection of the plane with the pillar surface. For more details on this
model, the reader is referred to (Parthasarathy et al., 2007).
This model has been widely used on experimental studies (Abad et al.,
2016), DDD simulations (Hu et al., 2019), Crystal Plasticity simulations (Gu
et al., 2021), as well as on recent discrete slip plane models (Wijnen et al.,
2021). In the present work, it is assumed that single-arm sources remain
immutable after activation. Changes in SAS could be potentially included in
the model, for example the incorporation of SAS destruction as proposed in
the work by Cui et al. (2015) on coated submicron pillars.
Screw dislocation segments. In addition to SASs, we also consider the possi-
bility of pre-existing pure-screw segments. In essence, such dislocations are
not pinned and can sweep across the slip plane, exiting the cylinder as they
reach its surface. Under these operating conditions, screw segments allow
to capture the so-called ”dislocation-starvation” behavior (Greer and Nix,
2006). In this case, the resolved stress necessary to move the dislocation is
only the lattice friction,
τscrew
CRS S =τ0(7)
The position of the screw dislocation within the pillar cross-section is
randomly chosen. As shown in section 3.3, this allows capturing different
cases of mechanical annealing observed in the literature, with variable stress
drop due to the activation of these sources. It is worth noting that in our
implementation, dislocation cross-slip due to image forces in the vicinity of
the free surfaces is not considered. This is further discussed in Sec. 4.
2.2.2. Surface nucleation model
For pillars and nanowires in the range of a few tenth of nanometers
to a few hundred of nanometers, virtually defect-free microstructures are
not unlikely. Under these conditions, mechanical testing of nanopillars and
9
nanowires reveal extremely high tensile stress and atomistic simulations point
to surface nucleation of dislocations as the source of plasticity (Zhu et al.,
2008). In consequence, the framework developed includes surface nucleation
effects.
The surface nucleation model adopted here was initially proposed after
atomistic simulations (Zhu et al., 2008; Ryu et al., 2011) and later adopted
on discrete dislocation dynamics (DDD) simulations (Ryu et al., 2015; Hu
et al., 2017, 2019). Briefly, within a time span t, the probability of a surface
nucleation event in a pillar of surface Sis given by
P=ν0exp[F(σ, T )
kBT]·S
b2·t. (8)
Here, ν0is an attempt frequency and kBthe Boltzmann constant.Thus, the
ratio S
b2stands for the number of potential nucleation sites. Finally, F(σ, T )
is an activation free energy that depends on stress and temperature through
F(σ, T ) = (1 T
Td
)·F0(σ) (9)
with
F0(σ) = A·(1 σ
σathm
)αSN .(10)
Here, Ais the zero-stress, zero-temperature activation energy, while αSN
and Tdare constants corresponding to the stress and temperature dependen-
cies, respectively. Tdis usually taken as half the bulk melting temperature
(Zhu et al., 2008). Finally, σathm is an athermal nucleation stress, correspond-
ing to the tensile stress required for surface nucleation at zero-temperature.
The possibility of surface nucleation is considered once P= 1.
The parameters A,Td,αSN and σathm are material-dependent and their
calibration require dedicated studies, either based on atomistic simulations
(Zhu et al., 2008) or on experiments (Chen et al., 2015). To the best of the
authors’ knowledge, there are no such studies on W. Therefore, we opt to fit
these parameters to the available experimental data, following suggestions
by Chen et al. (2015). Nanopillar compression and nanowire tensile test
for [100]-oriented W are extremely scarce in the literature. Srivastava et al.
(2021) report yield stresses of the order of 3.5 GPa in their 100 nm diameter
pillar compression tests, whereas ordoba et al. (2017) report yield stresses
of 6.6 ±1.2 GPa for their 93 nm suspended wires under bending. Therefore,
10
Symbol Property Value Unit
TdDisorder temperature 1800 K
A zero σ, zero Tactivation energy 6 eV
αSN Fitting exponent stress dependence 4 -
σathm Athermal nucleation stress 16 GPa
Table 1: Surface nucleation model parameters.
we opted to fit the parameters of eqn. 8 to a 100 nm target yield strength
of 4.5 GPa, in between the mean value of Srivastava et al. (2021) and the
lower bound of the range reported by ordoba et al. (2017). The result-
ing parameters are included in Table 1 after the fitting procedure presented
in the Supplementary Material file. The surface nucleated dislocations are
considered as pure screw segments, as described in the previous subsection.
2.2.3. Dislocation mobility and displacement rates
The displacement rate of the pillar in the slip direction sidue to the slip
produced in a plane iby the movement of a dislocation can be obtained using
a reasoning similar to Orowan’s equation,
˙
ui
p=bivi
Disi(11)
where biand viare the Burgers vector modulus, and dislocation velocity
corresponding to source i, respectively. Dicorresponds to the length in
direction siswept by a single dislocation when fully crossing the slip plane
i. If the unit vector estands for the pillar axis and Dfor its diameter, then
Di=D/(ni·e), and the vertical displacement rate corresponds to
˙ui
p=˙
ui
p·e=bivi
D/ni·esi·e=biviMi
D(12)
where Miis the Schmid factor of slip system associated to source iprojected
in direction ee.
In order to compute the displacement rate for each plane i, the associated
dislocation velocity vimust be determined. Dislocation-mediated plasticity
in bcc metals can take place in 48 bcc slip systems including the {110},{112}
and {123}families of slip planes. Interestingly, some works propose the
11
Symbol Property Value Unit
µ0Attempt frequency 9.1 1011 s1
hDistance between Peierls valleys 6/3a0
wKink pair width 11 b
ζMean dislocation segment width 25 b
τPPeierls stress 2.03 GPa
H0Kink-pair energy at 0 K 1.63 eV
pEnergy profile parameter 0.86 -
qEnergy profile parameter 1.69 -
BPhonon drag coefficient 9.8 104Pa*s
Table 2: Parameters for the dislocation mobility law. Obtained from molecular dynamics
simulations (Cereceda et al., 2013; Stukowski et al., 2015; Cereceda et al., 2016; Po et al.,
2016)
decomposition of {112}and {123}slip planes on alternating {110}slip planes
(Christian, 1983; Marichal et al., 2013; Ruestes et al., 2014). Based on this,
and for the sake of simplicity, we consider bcc metals with only {110}slip
planes (¯
b= 1/2a0<111 >dislocations). That is, |¯
b|is unique and equal
to a03/2, with a0the bcc lattice parameter.
The mobility laws for bcc tungsten dislocations have been deeply studied
using atomistic simulations (Cereceda et al., 2013; Stukowski et al., 2015;
Cereceda et al., 2016; Po et al., 2016). For tungsten, and considering rela-
tively low temperatures, dislocation velocities can be estimated by:
v=(µ0h(ζw)
bexp(H(τeff )
kBT),H > 0
τeff b
B,H0(13)
H(τef f ) = H01(τeff
τ
)pq(14)
In the expression, µ0is an attempt frequency, his the distance between
Peierls valleys, ζis the mean dislocation segment length for the slip system, w
is the kink-pair width and H0is the kink-pair formation energy. Parameters
pand qallow for the adjustment of the ”tail” (τeff 0) and ”top” (τef f τ)
of the energy profile H, respectively (Kocks et al., 1975). The parameters
for W are presented in Table 2.
The upper part of the definition (∆H > 0) corresponds to the kink-
pair thermally-activated regime, typical of bcc metals (Kocks et al., 1975).
12
The lower part, in turn, corresponds to the phonon-drag-dominated regime,
typical of high stress / high strain-rate conditions (Po et al., 2016). In Eq.
(14) τeff represents the effective resolved stress on the plane i,
τeff =|τi| τi
a(15)
where τiis the resolved shear stress, averaged over the plane, for source i
resulting from the Cauchy stress projected in the particular system to which
that source belongs to,
τi=σ: (sini).(16)
In Eq. (15) τais the resistance to slip due to athermal barriers, as dislocation
groups, precipitates, which are not considered in the present study (τi
a= 0).
Finally, τrepresents the part of the resistance to slip due to thermally
activatable obstacles. In the case of pure W the only thermal barrier is the
Peierls resistance, hence τ=τP.
In essence, the use of eq. 13 for the calculation of the pillar displace-
ment rates (eq. 12) allows to incorporate the thermally-activated motion
character of screw dislocations at low stresses and temperatures, as well as
the phonon-drag-dominated regime for newly surface-nucleated dislocations
under high stresses. Equations 12-13 allow for the effect of each plane iin
the displacement rate of the pillar.
2.3. Kinetic Monte Carlo procedure for stochastic slip events
In our model, the activation of the previously described microscopic slip
mechanisms is controlled by a kinetic Monte Carlo selection process. Con-
sidering a displacement-controlled micropillar compression experiment with
an applied strain rate of ˙ε0, the velocity of the upper part of the pillar cor-
responds to ˙u0=H˙ε0. This displacement is accommodated by either elastic
or plastic deformation. The plastic contribution to the displacement of the
pillar corresponds to the superposition of the displacement produced by slip
in the active planes
˙up=X
i
˙ui
p
with ˙ui
pgiven by eq. 12. The elastic displacement rate then corresponds to
˙uE= ˙u0˙up.(17)
Yu et al. (2021) proposed a rejection-free kinetic Monte Carlo algorithm
to obtain the stress-strain evolution in a deforming single crystal in which
13
the strain rates of the different slip systems provide the set of event rates.
In this work, this idea is generalized to a full deforming specimen solving
the macroscopic stress-strain response as well as the spatial distribution of
the fields involved. At time tn, each of the iescape pathways has a rate
constant ˙ui
p, that characterizes the probability per unit time that the system
escapes to that state i. These rates can be used to assemble an array of
partial sums representing the accumulated rate of all the objects up to and
including object j,
rj=
j
X
i=1
˙ui
p+r0(18)
with r0= ˙ε0Then, the total escape rate is the sum of all the rates
rtot =
N
X
i=1
˙ui
p+r0(19)
Nbeing the total number of sliding planes considered. In agreement with
rejection-free kMC theory, the next event to be executed is the kth process
satisfying
rk1< ξ1rtot rk(20)
If the event selected corresponds to the rate r0, the next event will cor-
respond to an elastic event, i.e. no inelastic deformation will be introduced
in the plastic layers. Otherwise, a plastic event is chosen, for which the
corresponding eigenstrain will be applied. The pillar geometry, array of dis-
placement rates, and the resulting pillar shape after the whole process are
exemplified in Figure 2.
Then the next time step is computed as:
δtn+1 =logξ2ε
rtot
(21)
such that the time is advanced as:
tn+1 =tn+δtn+1 (22)
ξ1and ξ2are random numbers uniformly distributed in (0,1] and εis a
normalization factor (∆ε<= 1) to adjust the time step within physical time
scale bounds (Yu et al., 2021). Once an event is executed, the strain fields
and stress fields are updated.
14
Figure 2: Schematic of stochastic selection process. a) a number of plastic regions are
randomly selected both for pre-existing dislocations (blue) and possible surface nucleation
sites (red). b) Upon loading and based on eqns. 12 through 19, a sampling array of events
in built. The total rate rtot is composed of the prescribed applied strain rate r0and the
vertical displacement rates of each plastic region. In general, the value of each vertical
displacement rate ˙ui
pis different from the rest of the displacement rates during each time
step, as represented by the different box sizes. c) After selection of a number of events
using rejection-free kMC, the final deformed state of the pillar reflects the operation of
pre-existing dislocations as well as a surface nucleation site. Arrow indicates most probable
event.
Time scale bounds. The absolute time scale emanating from the sampling
of eq. 17 using eqns. 18 through 21 represents the maximum time step
compatible with the rate equation under consideration. As shown by Yu
et al. (2021), it is indeed possible to use an arbitrarily smaller time step
without invalidating the method. A physically-reasonable upper bound for
the time step is that of the total time for a single dislocation to sweep the
pillar section:
ti=Di
vi(23)
In addition, the existence of tialso poses a constraint. On each plastic
event, the eigenstrain applied cannot be larger than that of eq. 4, which
corresponds to a dislocation fully sweeping a pillar on a time ti. Thus, tiis
15
an upper bound for the time increment δtn+1. As consequence, δtn+1/ti1.
In practice, the eigenstrain applied once a plastic event is selected is:
εEI G (i)(x) = ((δtn+1
ti)( b
h) sˆn), if xϵ plastic region
0, elsewhere (24)
Weakest link principle and deformation localization. Under the weakest link
principle, plastic flow occurs when the stress is high enough to activate the
weakest source available in the crystal (Norfleet et al., 2008). In our model,
weaker sources with a lower activation stress will tend to produce higher dis-
location velocities (eq. 13) and thus render higher plastic displacement rates
(eq. 12). In consequence, a sample with several SASs, out of which one is
significantly longer than the others, will result in a higher probability of acti-
vation of such source, thus favoring localization in the corresponding plastic
region. In contrast, for a sample with several SASs of commensurate length,
the kMC selection process will translate into the sequential activation of sev-
eral SASs, favoring delocalization of deformation into several plastic regions.
These aspects are further discussed in Sec. 4 after the results presented in
Sec. 3.
2.4. Solving the mechanical problem with an FFT algorithm
In order to obtain the strain and stress distribution in the pillar, which
drives the activation of slip in the plastic regions, an elastic problem with
eigenstrains has to be solved at each time step. An FFT-based approach will
be used for this purpose.
The total strain is the sum of the eigenstrains (input) and the elastic strain
εe, which appears to enforce total strain compatibility and stress equilibrium,
ε=εe+εEI G.(25)
Under linear elasticity, stress is related to strain through
σ=C:εe=C: (εεEI G) (26)
where Cis the stiffness tensor of the crystal.
The mechanical equilibrium corresponds to
· [C: (εεEI G)] = 0(27)
16
and rearranging terms, the result is a linear differential equation
· [C:ε] = −∇·[C:εEIG].(28)
If the right-hand side of Eq. (28) is known, as in our case, the total strain ε
can be directly obtained by the convolution of this term with the correspond-
ing Green’s function derivative. Moreover, if the medium is homogeneous and
periodic, closed expressions for Green’s function exist in Fourier space, and
the solution is just a multiplication.
However, when considering the present problem in a FFT framework,
the actual geometry has to be embedded in a cuboidal periodic domain
which also contains the outer free space. In this case, the domain can be
considered as an heterogeneous microstructure formed by two phases and
this microstructure can be described by
C=(C1, if xin 1
C2, if xin 2
(29)
In the simulation domain, the metallic pillar occupies the region (Ω1) and
is surrounded by an infinite compliant medium which occupies the region
(Ω2). For the solution of Eq. 28, considering Eq. 29, we follow the Fourier-
Galerkin approach. The details of the method can be found in (Vondˇrejc
et al., 2014; Zeman et al., 2017) and here only the final equations are recalled.
The problem is discretized in a box of [Nx, Ny, Nz] voxels which correspond
to the same number of frequencies in the discrete Fourier space. The value of
C, strain and stress fields are approximated by trigonometrical polynomial,
whose coefficients are given by the FFT of the value of these fields at the
voxels. Let Gbe the projection operator that, by convolution, extracts the
compatible part of a tensor field (Vondˇrejc et al., 2014). The mechanical
equilibrium of the discrete stress field translates into,
G σ=0(30)
where indicates a convolution. Introducing the value of σfrom eq. 26,
G (C(x) : (εεE IG)) = 0.(31)
The total strain εis decomposed into a homogeneous average strain ten-
sor, represented by the macro-scale applied strain ET, and a periodic fluctu-
ating micro-scale strain field ˜ε, which is the unknown.
ε=˜ε+ET(32)
17
Combining the last two equations results in a linear system in which the
strain fluctuation at the voxels are the unknowns.
G (C(x) : ˜ε) = −G (C(x):(ETεEI G)) (33)
This equation can be solved easily by transformation to Fourier space where
the operator has a closed expression and convolutions are transformed into
products. If a linear discrete operator GC() is defined as
GC() = F1(ˆ
G:F(C(x):())) (34)
with Fand F1the discrete Fourier transform operations and ˆ
Gthe projec-
tion operator in Fourier space, the equation to solve is
GC(˜ε) = −GC(ETεEIG)).(35)
To solve this linear problem efficiently, an iterative solver has to be used. The
convergence rate and quality of the solution depend on the phase property
contrast, which here is infinite because one of the phases is empty. Following
Lucarini et al. (2022) this problem can be solved efficiently and with minimal
noise by using a discrete projection operator (here rotated discrete G-operator
(Willot, 2015)) and a minimum residual iteration method as linear solver.
It is important to highlight that the framework presented is thought to
study general geometrical domains in 3-D with heterogeneous microscopic
fields. While this may not be regarded as fundamental for a simple cylin-
drical pillar, it is essential for the modeling of complex geometries, such as
nanoporous metals and nanoarchitected meta-materials.
For a set of initialization parameters, our method can be summarized by
the algorithm 1 presented in the Appendix.
3. Application to single-crystalline W pillar compression
Tungsten is a prototypical refractory metal with a high melting point
and good corrosion resistance. Its outstanding mechanical properties at high
temperatures, together with its low sputtering rate, make this material an
ideal candidate for plasma-facing components in fusion energy devices (Rieth
et al., 2013). In addition, tungsten is one of the most studied bcc metals
by micropillar compression testing (Schneider et al., 2009; Kim et al., 2010;
Abad et al., 2016; Srivastava et al., 2021) and its dislocation mobility laws are
18
well-documented (Cereceda et al., 2013, 2016; Po et al., 2016). These aspects
make W an ideal candidate to test the applicability of our framework.
Single crystalline tungsten pillars were simulated using the framework
described in the previous section. The pillar dimensions were varied in the
range from 25 nm to 1 µm at a strain-rate of 103s1. The crystalline orien-
tation and loading axis were consistent with the [100] direction. In all cases,
an initial dislocation density of ρ= 5·1012 m2was taken. Note that such an
election renders an initial dislocation count Ndin the range of 1 for the small-
est pillars to 8 for the largest ones. The number of potential nucleation sites
was taken as NSN = 128. The parameters of the surface nucleation model
are presented in Table 1. The parameters of the physically-based dislocation
mobility law are presented in Table 2. A list of the slip systems considered
can be found in Table A.3. The models used were discretized in a grid of
128 ·128 ·128 voxels with a kMC normalization factor of ε= 103. See
Supplementary Material file for sensitivity of the results to the discretization.
In particular, we focus on three important aspects: size effects (Sec.3.1),
strain-rate effects (Sec.3.2) and mechanical annealing effects (Sec.3.3). For
validation purposes, the results of the model are later compared with exper-
imental results in Sec. 4.
3.1. Size effects
Figure 3 presents the stress-strain curves for all the simulations performed
in this study. For each of the pillar dimensions probed, ten simulations were
conducted. Blue lines correspond to source truncation (SAS) dominated
cases, whereas red lines correspond to surface nucleation-dominated ones.
The envelope of the mechanical response for all the stress-strain curves is
illustrated by the corresponding colored areas. Black lines correspond to the
average of the results.
For the largest pillars, with a diameter of 1 µm, the plastic regime is
characterized by a rather smooth output with minimum scatter among the
curves. This is a consequence of having several single-arm sources distributed
along the sample, allowing the activation of several slip planes in a sequential
way. For 500 nm diameter pillars, the plastic regime is again smooth, but now
the curves appear more scattered, as judged by the upper and lower limits of
the blue-shaded region. For a smaller pillar diameter, the average number of
dislocations is close to one. As their pinning position is randomly selected,
significant scatter is then translated to the loading curves. The mean flow
stress is higher than before, as expected for a decreased pillar diameter.
19
For 200 nm diameter pillars, again a significant scatter is seen for the load-
ing curves. Interestingly, one case rendered a significantly high flow stress,
associated to surface nucleation of dislocations instead of the activation of
the SAS mechanism. This corresponds to the red curve, for which a SAS
length of 8 nm was randomly generated. The short length thus renders an
activation stress that is higher than that of the SN model. This behavior is
further analyzed in the Supplementary Material.
For 100 nm diameter pillars, now several cases appear both for the ac-
tivation of SA sources (blue) as well as surface nucleation (red). Both the
mean flow stresses for the SAS cases as well as the average (black) increase
with respect to the previous pillar diameter. For 50 nm diameter pillars, the
surface nucleation mechanism is dominant, with few cases of activation of SA
sources at higher stresses compared to previous dimensions. Finally, for 25
nm diameter pillars, surface nucleation was found to be the only operating
mechanism. Note that as surface nucleation becomes the dominant mecha-
nism, the curves show smaller scatter. This is consistent with experimental
observations (Kiener and Minor, 2011; Huang et al., 2011).
20
Figure 3: Stress-strain response for W micro/nanopillars under compression. For each
of the pillar dimensions probed, ten simulations were conducted. Shaded regions corre-
spond to the envelope of the mechanical response for all the stress-strain curves, where
blue corresponds to single-arm source dominated cases and red corresponds to surface
nucleation dominated cases. Blue and red lines correspond to the average of each subset
(SAS-dominated and SN-dominated, respectively), whereas black lines correspond to the
average of all the results, irrespectively of their nature.
3.2. Strain-rate effects
In order to evaluate the strain rate sensitivity of the flow strength, W
pillar simulations were conducted at different strain rates (103/s, 102/s
21
and 101/s) and for different pillar diameters (1 µm, 500 nm, 200 nm and
100 nm). Figure 4 presents the logarithm of the flow stress as function
of the logarithmic applied strain rate obtained by the model for pillars of
different diameter and compared to that of Srivastava et al. (2021) at room
temperature. The log-log flow stress data was then fitted with a power-law
function of the strain rate, σ˙ϵm. The strain rate sensitivity parameter m
is then m=(lnσ)/∂(ln ˙ϵ).
The resulting curves are approximately linear in the double-logarithm
diagram, indicating a strain rate sensitivity mapproximately constant in the
strain range explored. Comparison with Srivastava et al. (2021) shows not
only similar strain rate sensitivity parameters but also a similar transition
trend from positive to null values as the pillar diameter decreases.
22
Figure 4: Size-dependent strain-rate effects on the flow stress of W pillars. Empty circles
and dashed lines correspond to experimental results by Srivastava et al. (2021). For
comparison purposes, the colors are the same as the ones used in Srivastava et al. (2021)
and correspond to: blue - 1 µm; red - 500 nm; magenta - 200 nm; green - 100 nm.
3.3. Mechanical annealing
In addition to single-arm sources, pre-existing pure screw dislocations can
also be considered, as presented in the methods section. These are valuable
for capturing mechanical annealing effects (Shan et al., 2008; Huang et al.,
2011; Shan, 2012) associated with dislocation starvation (Greer and Nix,
2006). Figure 5 presents three distinctive stress-strain curves corresponding
to 100 nm diameter pillar simulations where one pre-existing screw dislo-
cation has been considered. The red curve corresponds to a pillar without
SASs and a screw segment located in the middle of the cross-section (SN-
dominated - case 1), while the green curve corresponds to a pillar without
SASs and a screw segment located halfway between the center of the pillar
23
and its surface (SN-dominated - case 2). In both cases, after the opera-
tion of the screw dislocation, the source has exhausted. In consequence, the
simulated test continues elastically until activation of the SN mode. A simi-
lar scenario would have been found for a case in which a pre-existing screw
dislocation habits a pillar with a SAS of sufficiently small length. In our
implementation, the position of the screw segment is randomly chosen in the
pillar cross section considering a uniform distribution. Specific knowledge
of the material under consideration would allow to modify such distribution
accordingly and without invalidating our methodology. Differences in load
drops for the activation of screw segments are discussed in Sec. 4. In con-
trast, the blue curve corresponds to a pillar with a screw dislocation in the
middle of the cross section and a sufficiently long SA source. Upon loading,
and after an initial elastic stage, the screw dislocation is activated -recall
that the critical resolved shear stress is lower than that of SASs -. After
exhaustion of the screw source, the mechanical response proceeds elastically
until reaching a stress corresponding to the CRSS of the SA source, which
gets activated several times for the rest of the simulation.
24
Figure 5: Stress-strain response for 100 nm diameter pillar simulations containing a pure
screw dislocation. Red and green curves correspond to pillars without SASs and a screw
dislocation located in the middle of the cross section (red - case 1) and halfway between
the center of the pillar and its surface (green - case 2). Blue curve corresponds to a pillar
with a sufficiently long SAS and a screw dislocation located in the middle of the cross
section.
4. Discussion
Comparison with other methods. Several aspects of our proposed frame-
work are worth discussing. The framework proposed aims at modelling the
full elastic-plastic response with 3D resolution of microfields. Its particular
application to nano-pillars and submicron-pillars allows to obtain the size
effect in the yield like other simple models (Parthasarathy et al., 2007; Jen-
nings et al., 2013), but here the microfields and post yield behavior are also
recovered, with the aided benefit of obtaining the 3-D deformation pattern
25
(see below). Our model can therefore be placed in between DDD approaches
and continuum descriptions, such as crystal plasticity. Regarding CP, con-
tinuum microfields are used but deformation is localized here. In this aspect,
our method shares a similarity with the concept of thin discrete slip bands
used by Wijnen et al. (2021) but the nature of our model is stochastic and
based on the displacement discontinuity created by slip events. Like in DDD
approaches (El-Awady et al., 2009; Srivastava et al., 2013; Cui et al., 2014;
Ryu et al., 2013, 2015, 2020), important aspects of the dislocation mobil-
ity laws and underlying physics are considered. The selection of the plastic
events does not take place using the weakest-link principle, like in (El-Awady
et al., 2009; Wijnen et al., 2021), but a Monte Carlo selection process over
the displacement rates calculated using an approach similar to Orowan’s
equation.
However, unlike DDD approaches, dislocations are not modeled explicitly,
but their effect is introduced by means of the concept of Eshelby inclusions,
such that the effect of dislocation gliding is introduced through an equiva-
lent eigenstrain. The total strain is thus the superposition of two fields, (ε
and the inelastic strain obtained as a sum of eigenstrains due to slip events
εEI G). Therefore, the calculation of the strain and stress fields becomes lin-
ear, in contrast to models based on localized crystal plasticity as (Wijnen
et al., 2021). This aspect, together with an inherently fast FFT solver, re-
sults in an efficient and computationally cheap stochastic framework capable
of simulating experiments with 3D spatial resolution at strain rates lower
than 103/s on nano and submicron samples within a few hours on an off-
the-shelf workstation. The present implementation is based on small strains
since we are focusing on the initial flow of the pillars, but can be extended to
finite strains by assuming a non-linear elasticity energy and solving the FFT
problem in finite strains (Zeman et al., 2017). Other aspects for further ex-
tension of the model include non-Schmid effects. This could be done following
steps already taken with recent advances on the development of a physically-
informed continuum crystal plasticity model for tantalum (Lee et al., 2023).
The framework would also allow for the incorporation of twinning as a com-
plementary deformation mechanism. Discrete dislocation plasticity models
have proven useful for the understanding of dislocation slip-mediated twin-
ning mechanisms (Wang et al., 2022) and could be used as input for the
extension of the framework presented here.
26
Flow stress size dependence. The proposed framework allows to obtain
the statistical distribution of the mechanical response of a pillar as a function
of its size, as shown in Fig 3. To analyze these results, in Figure 6 the
flow stress for different pillar sizes obtained with our model is compared to
selected experimental references (Schneider et al., 2009; Abad et al., 2016;
Srivastava et al., 2021; Kim et al., 2010) in a double logarithm representation.
Flow stress values presented correspond to the average stress in the plastic
regime of the curves presented in Fig. 3, while vertical bars correspond
to standard deviation of all the results presented in Fig. 3, capturing the
dispersion of results produced by the stochastic selection of the SAS length.
Schneider et al. (2009) performed load-controlled compression tests on [100]-
oriented W pillars spanning from the nm scale (200 nm) to the micrometer
scale (6 µm). Later on, the same group produced an experimental study
probing temperature effects on W pillars with similar orientation (Abad et al.,
2016). Our model successfully captures these experimental measurements as
well as an overall σ1 dependence of the yield strength in the 200
nm - 1 µm range (black dashed line). Srivastava et al. (2021) performed
displacement-controlled compression tests on [100]-oriented W nanopillars in
the range of 100 nm - 1 µm. Their results are systematically lower than
those reported in Schneider et al. (2009); Abad et al. (2016) yet following a
similar yield strength trend increase with decreasing size (σ1). Kim
et al. (2010) performed displacement-controlled compression tests on [100]-
oriented W nanopillars in the range of 200 - 900 nm. Their results represent
a lower bound for the experimental data set chosen. Yet their values again
follow similar trends (σ1). To the best of the authors knowledge, there
are no reports on [100]-oriented W pillars with diameters below 100 nm,
probably due to the associated experimental difficulties. At the limit of a
100 nm, we predict higher yield stresses than those reported in Srivastava
et al. (2021), yet our estimations agree with their measurements if we take
into account the standard deviation.
For diameters below 200 nm, the exponential dependence σ1 yields
to a surface nucleation-dominated trend (red dashed line), indicating a crossover
between predictions from the SAS and SN models. The critical sample size
observed is approximately 120 nm. Below this threshold, the flow stress de-
termined by the SN model is lower than that by the SAS model, highlighting
the comparative difficulty of activating a single-arm source over nucleating
a dislocation from a free surface, emphasizing a preference for surface nu-
cleation. In contrast, when dealing with samples that exceed this critical
27
size, plasticity is typically ruled by the activation of pre-existing disloca-
tion sources, such as those represented by the SAS mechanism. A handful
of previous studies had focused on the critical size for bcc and fcc metals
using both experimental and computational techniques. For fcc materials,
Hu et al. (2019) used DDD simulations to predict the the flow stress and
dominant yielding mechanisms in Cu nanopillars (100 nm < D < 800 nm),
reporting on a cross-over from SAS activation to SN operation as D reaches
110nm. Similarly, Shan et al. (2008) utilized transmission electron mi-
croscopy (TEM) to investigate the behavior of pre-existing dislocations in Ni
pillars of different diameters. The authors observed that in the case of a 160
nm diameter pillar, the pre-existing dislocations gradually exited the pillar,
followed by the emergence of new dislocations, which can be attributed to
the phenomenon known as SN. Conversely, in the case of pillars with a di-
ameter of 290 nm, dislocations persisted even after undergoing compression,
which can be tentatively associated with SAS. For bcc metals, experiments
conducted on bcc Mo pillars suggest a critical size of around 200 nm (Huang
et al., 2011; Shan, 2012). We conclude that our critical size prediction for
[100]-oriented W pillars is in agreement with previous observations for other
metals at these scales.
28
Figure 6: Log-log size-dependence of the flow stress compared with experimental literature.
Black dashed line indicates 1 dependence, in agreement with a SAS dominated regime.
Blue dashed line indicates the estimated power law exponent, while light-blue shadow
indicates ±values (αW0.31 ±0.11). Red dashed line indicates a surface nucleation
(SN) trend. Its intersection indicates a cross-over transition size. Vertical bars correspond
to standard deviation.
The results can be interpreted through the well-known power law relation,
namely
σy=σ0+k1DαW(36)
where σ0is the strength of the bulk material, k1is a constant and αWis the
power law exponent. In practice, σ0is usually neglected (Brinckmann et al.,
2008; Abad et al., 2016; Srivastava et al., 2021) as the second term in eq. 36
dominates at these small scales. As a reference, Srivastava et al. (2021) report
on exponent of 0.319 for their 103s1experiments, close to the αW= 0.33
value obtained by Abad et al. (2016) for the same strain rate. In turn, Kim
et al. (2010) informed a power law exponent in the range of 0.33 < αW<0.55.
On a lower end, Schneider et al. (2009) report on α1= 0.21. By linear fitting
eq. 36 on the SAS-dominated regime (D > 120 nm), we obtain an exponent
29
αW0.31±0.11, in agreement with most of the experimental evidence. This
further supports the ability of the single-arm source model to understand the
power-law relation on the strength vs size dependence of micropillars (Lee
and Nix, 2012). The red dashed line, connecting the average yield stress
values obtained for the smaller pillars, presents a slope that seems to be an
indication of a possibly weaker size dependence.
The underlying cause for the values of αWare worth a discussion. Brinck-
mann et al. (2008) used the screw dislocation cross-slip mechanism concept,
derived from high strain-rate MD and DD simulations (Weinberger and Cai,
2008; Greer et al., 2008), to provide an explanation for their bcc Mo pil-
lar compression test results. They argued that the relatively low power law
exponent of bcc metals (αM o = 0.45) could be attributed to such effects.
Interestingly, Kim et al. (2010) performed pillar compression tests on W, re-
porting on strikingly similar power-law exponents (αW= 0.44 ±0.11). One
could then assume that the same mechanism takes place in bcc W. Yet, our
power-law exponent estimations are on the lower bound of this regime, while
our model does not incorporate a single characteristic of the cross-slip mecha-
nism presented in (Weinberger and Cai, 2008). Therefore, our results suggest
that screw dislocation mobility alone could indeed provide adequate power-
law exponents provided the underlying physics of the dislocation mobility
law are fully incorporated ( eq. 13).
It is worth highlighting that previous experimental observations on bcc
Mo suggest that size effect itself has a strong size effect (Huang et al., 2011;
Shan, 2012). In their in-situ TEM bcc Mo studies, Huang et al. (2011) found
a strong increase in the power law exponent as the pillar diameter decreases
beyond 200 nm. In contrast, we predict a break down in the power-law
exponent as the W pillar diameter decreases below 200 nm. This difference
in behavior could be attributed to different activation energies between Mo
and W. Small variations in activation energy can strongly influence yield
stress at the nanoscale. A break down in exponent can also be inferred
from dislocation dynamics simulations on Cu (Hu et al., 2019). We attribute
the break down in power law exponent to the change of mechanism, from
SAS-dominated to surface nucleation dominated, for which a weaker stress
dependence on diameter is expected (Zhu et al., 2008; Huang et al., 2015;
Hu et al., 2019). Certainly, this alternative explanation of the results leaves
space for future research and informed discussion.
30
Strain-rate sensitivity. Our results show a size dependence of the strain
rate sensitivity parameter, which changes from positive mto nearly null val-
ues (Figure 4). Huang et al. (2015) and Srivastava et al. (2021) observed the
same positive to null trends for their bcc Fe and W pillar experiments, re-
spectively. The strain rate sensitivity exponent mdepends on the activation
volume Vaand on the resolved stress on the active plane τthrough (Kocks
et al., 1975):
m=kBT
τVa
.(37)
In the previous equation, both the critical shear τand the activation vol-
ume Vacan be size dependent, and the origin of size dependency of mmight
have both contributions. Regarding the effect of activation volume, Huang
et al. (2015) rationalized that in Fe this might be the main contribution to size
effect in m. This size dependent min Fe arises from a competition between a
size-independent friction term -that rules for large pillar diameters (D >500
nm)- and a size-dependent term -that rules for smaller pillar diameters-, with
larger activation volumes for increasingly smaller pillars (100 nm <D<500
nm). They also argue that as the pillar diameter further decreases (D <100
nm), the activation volume becomes again smaller. Still, Fe is known to have
an activation volume that may display larger variations with stress compared
to other bcc metals with higher melting point (Christian and Masters, 1964).
Tungsten, in constrast, and particularly for low homologous temperatures,
displays a very low activation volume (Va10b3) (Kiener et al., 2019), con-
sistent with the dominance of the kink mechanism at low temperatures. Our
calculations of the activation volume (see Suppl. Mat.) render that for the
kink-pair activated mechanism, the stress-dependent activation volume is in
the range of 4 b3< V kink
a<11 b3for the stress range of the SAS-dominated
regime, whereas for the surface nucleation mechanism is in the range of 2.5
b3< V SN
a<4.5 b3for the stress range of the SN-dominated regime.
Therefore, considering the relatively low variations in the activation vol-
ume Vain W, changes in the strain rate sensitivity exponent mcan be un-
derstood based on the size-dependence of the stress reached on the pillar due
to the size dependent strength of SAS, as well as on the stress-dependence
of the dislocation mobility law. Similar conclusions on the importance of τ
were reached by Srivastava et al. (Srivastava et al., 2021) for their strain-rate
sensitivity studies on W pillars.
Future investigations using our proposed framework could also focus on
temperature effects, as a recently developed CP model using similar disloca-
31
tion mobility laws has proven useful in exploring the temperature dependence
of deformation localization in irradiated tungsten (Li et al., 2021).
Mechanical annealing. Earlier MD simulation studies predict that such
behavior cannot take place in bcc pillars. Such simulations suggested that
the combined effects of image forces, together with the dislocation core struc-
ture, would facilitate dislocation multiplication by a surface-assisted cross-
slip mechanism (Weinberger and Cai, 2008). SEM observations of multi-
ple slip traces on compressed bcc pillars seem to support this mechanism
(Schneider et al., 2009). On the other hand, in-situ TEM compression of Mo
nanopillars (Huang et al., 2011; Shan, 2012) and Fe–3% Si pillars (Zhang
et al., 2012) show clear evidence that mechanical annealing by dislocation
starvation can indeed take place in bcc metals. Our proposed computational
framework takes into account such scenarios. Interestingly, load drops due
to mechanical annealing are, in general, less marked than in our Figure 5, as
shown in (Shan et al., 2008). Other works do report stress drops of the order
of 0.5 - 1.0 GPa (Huang et al., 2011; Zhang et al., 2012). Both situations
can be taken into account in our method by an adequate selection of the
statistical distribution of pre-existing screw segments.
Deformation patterns and localization. Figure 7 presents the typical
deformation patterns obtained in our simulations, together with a compari-
son with a selected reference (Srivastava et al., 2021). For our largest pillar
diameter (D= 1µm), the deformation pattern reveals the activation of sev-
eral SASs on a variety of slip systems (see Supplementary Material movie), in
agreement with scanning electron microscopy (SEM) analysis of W micron-
sized pillars (Schneider et al., 2009; Abad et al., 2016). As the pillar diameter
decreases to the submicron regime (D= 500 nm and D= 200 nm) and for
the dislocation density chosen, only one SAS is present. Upon activation,
significant deformation takes place and concentrates on a single slip band,
in agreement with SEM micrographs of W pillars with comparable diame-
ters (Srivastava et al., 2021). Further decrease of pillar diameter into the
nanometer regime (D= 100 nm) triggers surface nucleation instead of SAS
activation. The former takes place on multiple slip systems.
As explained in Sec. 2.3, our model favors the activation of weaker
sources. Localization of plastic activity in one or a few number of planes
depends on the number of potential slip sites (e.g. number of SAS) and
the statistical distribution of the slip resistance on that planes (linked with
32
the length of each SAS). Our 1 µm diameter pillars with ρ= 5 ·1012 m2
have multiple SASs of variable length. They show a tendency for a rather
delocalized (homogeneous) distribution of deformation. In contrast, 200 nm
- 500 nm diameter pillars with similar dislocation density have 1 - 2 SASs on
average. In such cases, localization is favored (Figure 7). Cui et al. (2018)
presented a 2D Monte-Carlo model of DD source activation coupled with
crossslip channel widening to reproduce and physically explain the transi-
tion in the mechanism of plastic flow localization in irradiated materials from
irradiation-controlled to dislocation source-controlled. They showed that as
the size decreases, the spatial correlation of plastic deformation decreases due
to weaker dislocation interactions and less frequent cross-slip, thus producing
thinner dislocation channels. For a low irradiation damage, they show that
as the diameter decreases from 1.5 µm to 300 nm, deformation transitions
from rather homogeneous to highly localized. Our results broadly agree with
this picture.
Statistical analysis of burst displacement. In order to analyze the
statistics of the plastic bursts obtained in the simulations, the complementary
cumulative distribution function (CCDF) of the burst displacement magni-
tude Uis obtained. The CDDF is computed using the method presented by
Cui et al. (2016) and its result for a 1 µm diameter pillar is represented in Fig-
ure 8. The Figure suggest that, under the displacement-controlled conditions
(strain control) explored here, the system does not exhibit power-law scaling,
in agreement with Cui et al. (2016). In addition, the data spans less than two
orders of magnitude. In contrast, typical load-displacement controlled com-
pression experiments usually span several orders of magnitude, with bursts
often exceeding 10 nm (Alcal´a et al., 2020; Srivastava et al., 2021). Detailed
tracking of the plastic strain rate (Inset of Figure 8) shows that the dynamical
behavior is quasiperiodic, in agreement with displacement-controlled simu-
lations (Cui et al., 2016) and experiments (Papanikolaou et al., 2012). In
contrast, experiments under load-controlled conditions lead to a completely
different scenario, where plasticity does not lead to marked stress drops,
allowing to trigger the operation of several SAS simultaneously and thus
exhibiting power law scaling. As a result, avalanche events can develop,
leading to a highly correlated dynamical response, consistent with the con-
cept of self-organized criticality (Bak et al., 1988). The framework presented
here, inherently implies the activation of one source at a time, leading to a
quasiperiodic response that it is consistent with displacement-control condi-
33
tions.
On the incorporation of strain hardening effects. In all the cases pre-
sented in Figure 3, the flow stress keeps constant, in agreement with the
assumptions introduced in Sec. 2. This is consistent with an scenario in
which the slip planes swept by the SAS exhibit a scarcity of dislocations in
intersecting planes, and where the pinning point of the SAS is stable un-
der the prevailing stress levels. Consequently, the activated SAS does not
encounter other defects, hence preventing the formation of new junctions
or the reduction of the source length. In addition, experiments on W sin-
gle crystalline pillars (Schneider et al., 2009) do not show hardening for the
relatively low strains explored in Figure 3.
For larger samples, our method could be extended to incorporate harden-
ing effects by including a Taylor hardening-like term in eq. 6 and a disloca-
tion density evolution law (Cui et al., 2014). The latter should incorporate
size-dependent terms and bulk-like terms. Size-dependent terms include the
generation of dislocations due to the operation of existing sources (1)
and the escape of dislocations reaching a free surface ( 1/D) (Cui et al.,
2014). Bulk-like terms include dislocation multiplication due to forest dislo-
cations (ρ) and the annihilation of closely spaced dislocations of opposite
signs ( ρ) (Devincre et al., 2008). If self and pair interactions must be
taken into account (e.g. for micron scale pillars above 1 µm), the correspond-
ing hardening coefficient matrix can also be incorporated (Yu et al., 2021;
Lee et al., 2023). It must be emphasized that for submicron pillars, simple
calculations show that the contributions of the bulk-like terms are nearly an
order of magnitude smaller than the size-dependent terms (Cui et al., 2014).
Thus, by neglecting the former and provided the dislocation generation rate
is balanced with the escape rate from free surface which, together with a
relatively low dislocation density, the results presented in Sec.3.1 would be
reproduced.
To wrap up, the results presented here focus solely on tungsten and a
relatively simple geometry. Yet, these are not restrictions. For instance,
FeCrAl alloys and refractory multi-principal element alloys (RMPEAs) are
also promising candidate materials for use under extreme conditions in the
nuclear industry. FeCrAl alloys have already been probed using discrete dis-
location dynamics (Pachaury et al., 2023) as well as with crystal plasticity
approaches using Arrhenius-type rate equations (Gong et al., 2023). With
respect to RMPEAs, kink-migration of screw dislocations have been shown
34
to play a dominant role in their deformation (Balbus et al., 2024) and it is
expected that developments in dislocation mobility laws (Shen and Spearot,
2021) would allow for systematic investigations on these promising alloys us-
ing upper scale methods. Our proposed framework offers flexibility to explore
the micropillar compression behavior of such alloys by the incorporation of
specific dislocation mobility laws. To further extend its potential, the use
of a 3D FFT solver offers opportunities for the exploration of a variety of
complex geometries, including nanoporous metals and nanoarchitected meta-
materials.
Figure 7: Residual deformation patterns for different pillar diameters explored, together
with SEM observations by Srivastava et al. (2021). Reprinted with permission from Else-
vier. Frame colors correspond to dominating deformation mechanism (Blue - Single-Arm
source, Red - Surface nucleation).
35
Figure 8: Complementary cumulative distribution function (CCDF) of burst displacement
for one example of our 1 µm diameter pillars. Inset, corresponding evolution of plas-
tic strain rate showing typical quasiperiodic strain bursts, consistent with displacement-
controlled compression conditions. Note the breakdown from power law scaling (dashed
line).
5. Conclusions
A novel approach is presented to simulate the deformation of submicron
specimens, accounting for the discrete events produced by the slip of in-
ternal dislocation and surface nucleated ones in a stochastic manner. The
framework considers the slip events as eigenstrain fields that produce a dis-
placement jump across a slip plane and whose activation is driven by a Monte
Carlo (MC) method. Physically-based laws are incorporated to account for
activation probabilities, dislocation mobility and surface nucleation.
Two factors of stochasticity are taken into account: a random selection
of the position of defects and a random selection of plastic events after a
MC process on a sampling array of possible displacement rates, computed
using a similar concept as in Orowan’s equation. Implementation on a fast
FFT solver, results on an efficient, computationally-cheap algorithm which
36
allows to simulate accurately the deformation of a specimen in a wide range
of strain rates and sizes in a fraction of the time needed using techniques as
MD or DDD.
The framework developed has been applied to the compression of tung-
sten (W) on [100]-oriented pillars in the 25 nm to 1 µm diameter range.
We find an adequate agreement between simulations and experimental data.
The study a size-dependence of flow stress, characterized by distinct power-
law exponents for regimes dominated by source truncation (D >120 nm,
αW0.31 ±0.11) and a break down of the power law scaling as surface sur-
face nucleation becomes dominant (D <120 nm). Pre-existing dislocations
naturally compete with surface nucleation of new dislocations. The former
are favored for pillars above 120 nm diameter, whereas the latter is favored
below such critical size. In addition, strain-rate sensitivity effects are ade-
quately captured, including a size-dependence of the strain-rate sensitivity
exponent. This is attributed not only to changes in activation volume and
source strength but also to the stress-dependent dislocation mobility law,
whose form considers both the kink-pair regime, dominating at the relatively
low stresses found in submicron pillars, as well as the phonon drag regime,
that rules for very high stresses found in pillars of 200 nm and below. The
framework inherently implies the activation of one source at a time, leading
to a quasiperiodic response that it is consistent with displacement-control
conditions.
Finally, the framework presented here opens the possibility to provide
valuable analysis and interpretations of other small-scale testing techniques,
such as microtensile testing and microbending. Future developments of dis-
location mobility laws for FeCrAl alloys and multi-principal element alloys
would allow for a systematic investigation of these relevant metals, while
the possibility to include other micro and nanoscale geometries further ex-
tends the potential applications map to include nanolattices and nanoporous
materials, among other nanostructures of interest.
6. Acknowledgements
This project has received funding from the European Union’s Horizon
Europe research and innovation programme under the Marie Sklodowska-
Curie grant agreement no. 101062254. Funded by the European Union.
Views and opinions expressed are however those of the author(s) only and
37
do not necessarily reflect those of the European Union. Neither the European
Union nor the granting authority can be held responsible for them.
Appendix A. Appendix Section
Algorithm 1 kMC discrete slip approach with eigenstrains
1: procedure Initialization
Define pillar geometry, C(x), in the discrete domain.
Set positions, eigenstrain fields and strength of each plastic region i
Set loading conditions ET(t),∆ε.
2: procedure Solution
3: while t < tT OT do:
4: Get rnd no. ξ1,ξ2(0,1]
5: Solve the equilibrium (Eq. 35) using FFT solver to obtain σ(Eq.
26) , |τi|(Eq. 16)
6: for each region do
7: if |τi|> τCRS S then
8: Compute: P(SN sites only), vi(Eq. 13) , ti(Eq. 23) and
˙ui
p(Eq. 12).
9: else
10: vi= 0
11: Compute accumulated rates rj,rtot
12: δtn+1 =logξ2ε
rtot
13: if ξ1<r0
rtthen
14: δtn+1 =dt0
15: next event is elastic: εE IG (i)= 0
16: else
17: next event is plastic
18: Compute eigenstrain: εEIG (i)(Eq. 24)
19:
20: Update total eigenstrain field εEI G (Eq. 5)
The slip systems considered in this study are presented in Table A.3.
38
iSlip system sini
1 [1¯
11](011) [1¯
11] [011]
2 [¯
1¯
11](011) [¯
1¯
11] [011]
3 [111](0¯
11) [111] [0¯
11]
4 [¯
111](0¯
11) [¯
111] [0¯
11]
5 [¯
111](101) [¯
111] [101]
6 [¯
1¯
11](101) [¯
1¯
11] [101]
7 [111](¯
101) [111] [¯
101]
8 [1¯
11](¯
101) [1¯
11] [¯
101]
9 [¯
111](110) [¯
111] [110]
10 [¯
11¯
1](110) [¯
11¯
1] [110]
11 [111](¯
110) [111] [¯
110]
12 [11¯
1](¯
110) [11¯
1] [¯
110]
Table A.3: Slip systems considered in this study
References
Abad, O.T., Wheeler, J.M., Michler, J., Schneider, A.S., Arzt, E., 2016.
Temperature-dependent size effects on the strength of ta and w micropil-
lars. Acta Materialia 103, 483–494.
Alcal´a, J., cen´sek, J., Varillas, J., A. El-Awady, J., Wheeler, J.M., Michler,
J., 2020. Statistics of dislocation avalanches in fcc and bcc metals: dis-
location mechanisms and mean swept distances across microsample sizes
and temperatures. Scientific reports 10, 19024.
Bak, P., Tang, C., Wiesenfeld, K., 1988. Self-organized criticality. Physical
review A 38, 364.
Balbus, G., Rao, S., Senkov, O., Payton, E., 2024. Orientation dependent
plasticity of the refractory multi-principal element alloy monbti investi-
gated via micropillar compression. Acta Materialia 262, 119401.
Brinckmann, S., Kim, J.Y., Greer, J.R., 2008. Fundamental differences in
mechanical behavior between two types of crystals at the nanoscale. Phys-
ical Review Letters 100, 155502.
Capolungo, L., Taupin, V., 2019. Gd 3: generalized discrete defect dynamics.
Materials Theory 3, 1–21.
39
Cereceda, D., Diehl, M., Roters, F., Raabe, D., Perlado, J.M., Marian, J.,
2016. Unraveling the temperature dependence of the yield strength in
single-crystal tungsten using atomistically-informed crystal plasticity cal-
culations. International Journal of Plasticity 78, 242–265.
Cereceda, D., Stukowski, A., Gilbert, M., Queyreau, S., Ventelon, L.,
Marinica, M.C., Perlado, J., Marian, J., 2013. Assessment of interatomic
potentials for atomistic analysis of static and dynamic properties of screw
dislocations in w. Journal of Physics: Condensed Matter 25, 085702.
Chen, L.Y., He, M.r., Shin, J., Richter, G., Gianola, D.S., 2015. Measur-
ing surface dislocation nucleation in defect-scarce nanostructures. Nature
materials 14, 707–713.
Christian, J., 1983. Some surprising features of the plastic deformation of
body-centered cubic metals and alloys. Metallurgical transactions A 14,
1237–1256.
Christian, J.W., Masters, B., 1964. Low-temperature deformation of body-
centred cubic metals. ii. mechanism of thermally activated flow. Proceed-
ings of the Royal Society of London. Series A. Mathematical and Physical
Sciences 281, 240–257.
ordoba, R., Lorenzoni, M., Pablo-Navarro, J., Mag´en, C., erez-Murano,
F., De Teresa, J.M., 2017. Suspended tungsten-based nanowires with en-
hanced mechanical properties grown by focused ion beam induced deposi-
tion. Nanotechnology 28, 445301.
Cruzado, A., Gan, B., Jim´enez, M., Barba, D., Ostolaza, K., Linaza, A.,
Molina-Aldareguia, J., Llorca, J., Segurado, J., 2015. Multiscale modeling
of the mechanical behavior of in718 superalloy based on micropillar com-
pression and computational homogenization. Acta Materialia 98, 242–253.
Cui, Y., Lin, P., Liu, Z., Zhuang, Z., 2014. Theoretical and numerical in-
vestigations of single arm dislocation source controlled plastic flow in fcc
micropillars. International Journal of Plasticity 55, 279–292.
Cui, Y., Po, G., Ghoniem, N., 2016. Controlling strain bursts and avalanches
at the nano-to micrometer scale. Physical review letters 117, 155502.
40
Cui, Y., Po, G., Ghoniem, N., 2018. Size-tuned plastic flow localization in
irradiated materials at the submicron scale. Physical review letters 120,
215501.
Cui, Y.n., Liu, Z.l., Zhuang, Z., 2015. Theoretical and numerical investiga-
tions on confined plasticity in micropillars. Journal of the Mechanics and
Physics of Solids 76, 127–143.
Dehm, G., Jaya, B.N., Raghavan, R., Kirchlechner, C., 2018. Overview on
micro-and nanomechanical testing: New insights in interface plasticity and
fracture at small length scales. Acta Materialia 142, 248–282.
Devincre, B., Hoc, T., Kubin, L., 2008. Dislocation mean free paths and
strain hardening of crystals. Science 320, 1745–1748.
El-Awady, J.A., Wen, M., Ghoniem, N.M., 2009. The role of the weakest-
link mechanism in controlling the plasticity of micropillars. Journal of the
Mechanics and Physics of Solids 57, 32–50.
Fan, H., Wang, Q., El-Awady, J.A., Raabe, D., Zaiser, M., 2021. Strain rate
dependency of dislocation plasticity. Nature communications 12, 1845.
Gong, M., Xie, D., Sun, T., Zhang, X., Shao, L., Wang, J., 2023. Bridging
microscale to macroscale mechanical property measurements of fecral al-
loys by crystal plasticity modeling. International Journal of Plasticity 165,
103608.
Greer, J.R., De Hosson, J.T.M., 2011. Plasticity in small-sized metallic sys-
tems: Intrinsic versus extrinsic size effect. Progress in Materials Science
56, 654–724.
Greer, J.R., Nix, W.D., 2006. Nanoscale gold pillars strengthened through
dislocation starvation. Physical Review B 73, 245410.
Greer, J.R., Weinberger, C.R., Cai, W., 2008. Comparing the strength of fcc
and bcc sub-micrometer pillars: Compression experiments and dislocation
dynamics simulations. Materials Science and Engineering: A 493, 21–25.
Gu, Y., Eastman, D.W., Hemker, K.J., El-Awady, J.A., 2021. A statisti-
cal model for predicting size effects on the yield strength in dislocation-
mediated crystal plasticity. Journal of the Mechanics and Physics of Solids
147, 104245.
41
Healy, C.J., Ackland, G.J., 2014. Molecular dynamics simulations of
compression–tension asymmetry in plasticity of fe nanopillars. Acta Ma-
terialia 70, 105–112.
Hu, J., Liu, Z., Van der Giessen, E., Zhuang, Z., 2017. Strain rate effects
on the plastic flow in submicron copper pillars: Considering the influence
of sample size and dislocation nucleation. Extreme Mechanics Letters 17,
33–37.
Hu, J., Song, H., Liu, Z., Zhuang, Z., Liu, X., Sandfeld, S., 2019. Predicting
the flow stress and dominant yielding mechanisms: analytical models based
on discrete dislocation plasticity. Scientific Reports 9, 20422.
Huang, L., Li, Q.J., Shan, Z.W., Li, J., Sun, J., Ma, E., 2011. A new regime
for mechanical annealing and strong sample-size strengthening in body
centred cubic molybdenum. Nature communications 2, 547.
Huang, R., Li, Q.J., Wang, Z.J., Huang, L., Li, J., Ma, E., Shan, Z.W., 2015.
Flow stress in submicron bcc iron single crystals: sample-size-dependent
strain-rate sensitivity and rate-dependent size strengthening. Materials
Research Letters 3, 121–127.
Jennings, A.T., Weinberger, C.R., Lee, S.W., Aitken, Z.H., Meza, L., Greer,
J.R., 2013. Modeling dislocation nucleation strengths in pristine metallic
nanowires under experimental conditions. Acta materialia 61, 2244–2259.
erusalem, A., Fern´andez, A., Kunz, A., Greer, J.R., 2012. Continuum mod-
eling of dislocation starvation and subsequent nucleation in nano-pillar
compressions. Scripta Materialia 66, 93–96.
Kiener, D., Fritz, R., Alfreider, M., Leitner, A., Pippan, R., Maier-Kiener,
V., 2019. Rate limiting deformation mechanisms of bcc metals in confined
volumes. Acta Materialia 166, 687–701.
Kiener, D., Minor, A.M., 2011. Source-controlled yield and hardening of cu
(1 0 0) studied by in situ transmission electron microscopy. Acta Materialia
59, 1328–1337.
Kiener, D., Motz, C., Dehm, G., 2009. Micro-compression testing: A critical
discussion of experimental constraints. Materials Science and Engineering:
A 505, 79–87.
42
Kim, J.Y., Jang, D., Greer, J.R., 2010. Tensile and compressive behavior
of tungsten, molybdenum, tantalum and niobium at the nanoscale. Acta
Materialia 58, 2355–2363.
Kocks, U.F., AS, A., MF, A., 1975. Thermodynamics and kinetics of slip .
Konstantinidis, A.A., Aifantis, K.E., De Hosson, J.T.M., 2014. Capturing
the stochastic mechanical behavior of micro and nanopillars. Materials
Science and Engineering: A 597, 89–94.
Lee, S., Cho, H., Bronkhorst, C.A., Pokharel, R., Brown, D.W., Clausen,
B., Vogel, S.C., Anghel, V., Gray III, G.T., Mayeur, J.R., 2023. Defor-
mation, dislocation evolution and the non-schmid effect in body-centered-
cubic single-and polycrystal tantalum. International Journal of Plasticity
163, 103529.
Lee, S.W., Nix, W.D., 2012. Size dependence of the yield strength of fcc and
bcc metallic micropillars with diameters of a few micrometers. Philosoph-
ical Magazine 92, 1238–1260.
Li, Z., Liu, Z., Zhuang, Z., Cui, Y., 2021. Temperature dependent deforma-
tion localization in irradiated tungsten. International Journal of Plasticity
146, 103077.
Lin, P., Liu, Z., Cui, Y., Zhuang, Z., 2015. A stochastic crystal plasticity
model with size-dependent and intermittent strain bursts characteristics at
micron scale. International Journal of Solids and Structures 69, 267–276.
Lin, P., Liu, Z., Zhuang, Z., 2016. Numerical study of the size-dependent de-
formation morphology in micropillar compressions by a dislocation-based
crystal plasticity model. International Journal of Plasticity 87, 32–47.
Lucarini, S., Cobian, L., Voitus, A., Segurado, J., 2022. Adaptation and
validation of fft methods for homogenization of lattice based materials.
Computer Methods in Applied Mechanics and Engineering 388, 114223.
Marichal, C., Van Swygenhoven, H., Van Petegem, S., Borca, C., 2013. {110}
slip with {112}slip traces in bcc tungsten. Scientific reports 3, 2547.
Mura, T., 2013. Micromechanics of defects in solids. Springer Science &
Business Media.
43
Ng, K., Ngan, A., 2008. Stochastic nature of plasticity of aluminum micro-
pillars. Acta Materialia 56, 1712–1720.
Norfleet, D., Dimiduk, D., Polasik, S., Uchic, M., Mills, M., 2008. Dis-
location structures and their relationship to strength in deformed nickel
microcrystals. Acta Materialia 56, 2988–3001.
Pachaury, Y., Warren, G., Wharry, J.P., Po, G., El-Azab, A., 2023. Plastic-
ity in irradiated fecral nanopillars investigated using discrete dislocation
dynamics. International Journal of Plasticity , 103676.
Papanikolaou, S., Dimiduk, D.M., Choi, W., Sethna, J.P., Uchic, M.D.,
Woodward, C.F., Zapperi, S., 2012. Quasi-periodic events in crystal plas-
ticity and the self-organized avalanche oscillator. Nature 490, 517–521.
Parthasarathy, T.A., Rao, S.I., Dimiduk, D.M., Uchic, M.D., Trinkle, D.R.,
2007. Contribution to size effect of yield strength from the stochastics of
dislocation source lengths in finite samples. Scripta Materialia 56, 313–316.
Po, G., Cui, Y., Rivera, D., Cereceda, D., Swinburne, T.D., Marian, J.,
Ghoniem, N., 2016. A phenomenological dislocation mobility law for bcc
metals. Acta Materialia 119, 123–135.
Raabe, D., Ma, D., Roters, F., 2007. Effects of initial orientation, sample ge-
ometry and friction on anisotropy and crystallographic orientation changes
in single crystal microcompression deformation: A crystal plasticity finite
element study. Acta Materialia 55, 4567–4583.
Rao, S.I., Dimiduk, D., Parthasarathy, T.A., Uchic, M., Tang, M., Wood-
ward, C., 2008. Athermal mechanisms of size-dependent crystal flow
gleaned from three-dimensional discrete dislocation simulations. Acta Ma-
terialia 56, 3245–3259.
Rao, S.I., Dimiduk, D., Tang, M., Uchic, M., Parthasarathy, T., Woodward,
C., 2007. Estimating the strength of single-ended dislocation sources in
micron-sized single crystals. Philosophical Magazine 87, 4777–4794.
Rieth, M., Dudarev, S.L., De Vicente, S.G., Aktaa, J., Ahlgren, T., An-
tusch, S., Armstrong, D., Balden, M., Baluc, N., Barthe, M.F., et al.,
2013. Recent progress in research on tungsten materials for nuclear fusion
applications in europe. Journal of Nuclear Materials 432, 482–500.
44
Rodney, D., Le Bouar, Y., Finel, A., 2003. Phase field methods and disloca-
tions. Acta Materialia 51, 17–30.
Romero, I., Segurado, J., LLorca, J., 2008. Dislocation dynamics in non-
convex domains using finite elements with embedded discontinuities. Mod-
elling and Simulation in Materials Science and Engineering 16, 035008.
Ruestes, C.J., Bringa, E.M., Stukowski, A., Nieva, J.R., Tang, Y., Meyers,
M., 2014. Plastic deformation of a porous bcc metal containing nanometer
sized voids. Computational Materials Science 88, 92–102.
Ryu, I., Cai, W., Nix, W.D., Gao, H., 2015. Stochastic behaviors in plastic
deformation of face-centered cubic micropillars governed by surface nucle-
ation and truncated source operation. Acta Materialia 95, 176–183.
Ryu, I., Gravell, J., Cai, W., Nix, W.D., Gao, H., 2020. Intrinsic size depen-
dent plasticity in bcc micro-pillars under uniaxial tension and pure torsion.
Extreme Mechanics Letters 40, 100901.
Ryu, I., Nix, W.D., Cai, W., 2013. Plasticity of bcc micropillars controlled
by competition between dislocation multiplication and depletion. Acta
Materialia 61, 3233–3241.
Ryu, S., Kang, K., Cai, W., 2011. Entropic effect on the rate of dislocation
nucleation. Proceedings of the National Academy of Sciences 108, 5174–
5178.
Schneider, A., Kaufmann, D., Clark, B., Frick, C., Gruber, P., onig, R.,
Kraft, O., Arzt, E., 2009. Correlation between critical temperature and
strength of small-scale bcc pillars. Physical Review Letters 103, 105501.
Shan, Z., 2012. In situ tem investigation of the mechanical behavior of
micronanoscaled metal pillars. JOM 64, 1229–1234.
Shan, Z., Mishra, R.K., Syed Asif, S., Warren, O.L., Minor, A.M., 2008.
Mechanical annealing and source-limited deformation in submicrometre-
diameter ni crystals. Nature materials 7, 115–119.
Shao, S., Abdolrahim, N., Bahr, D.F., Lin, G., Zbib, H.M., 2014. Stochastic
effects in plasticity in small volumes. International Journal of Plasticity
52, 117–132.
45
Shen, Y., Spearot, D.E., 2021. Mobility of dislocations in fenicrcocu high
entropy alloys. Modelling and Simulation in Materials Science and Engi-
neering 29, 085017.
Srivastava, K., Gr¨oger, R., Weygand, D., Gumbsch, P., 2013. Dislocation
motion in tungsten: atomistic input to discrete dislocation simulations.
International Journal of Plasticity 47, 126–142.
Srivastava, P., Jiang, K., Cui, Y., Olivera, E., Ghoniem, N., Gupta, V., 2021.
The influence of nano/micro sample size on the strain-rate sensitivity of
plastic flow in tungsten. International Journal of Plasticity 136, 102854.
Stukowski, A., Cereceda, D., Swinburne, T.D., Marian, J., 2015. Thermally-
activated non-schmid glide of screw dislocations in w using atomistically-
informed kinetic monte carlo simulations. International Journal of Plas-
ticity 65, 108–130.
Tang, H., Schwarz, K., Espinosa, H., 2008. Dislocation-source shutdown and
the plastic behavior of single-crystal micropillars. Physical Review Letters
100, 185503.
Uchic, M.D., Dimiduk, D.M., 2005. A methodology to investigate size scale
effects in crystalline plasticity using uniaxial compression testing. Materi-
als Science and Engineering: A 400, 268–278.
Uchic, M.D., Dimiduk, D.M., Florando, J.N., Nix, W.D., 2004. Sample
dimensions influence strength and crystal plasticity. Science 305, 986–989.
Vondˇrejc, J., Zeman, J., Marek, I., 2014. An fft-based galerkin method
for homogenization of periodic media. Computers & Mathematics with
Applications 68, 156–173.
Wang, C., Sun, C., Cai, W., Qian, L., Guo, X., Fu, M., 2022. Evolution
of partial dislocation slip–mediated deformation twins in single crystals: a
discrete dislocation plasticity model and an analytical approach. Interna-
tional Journal of Plasticity 152, 103230.
Wang, Y.U., Jin, Y., Cuitino, A., Khachaturyan, A., 2001. Nanoscale phase
field microelasticity theory of dislocations: model and 3d simulations. Acta
Materialia 49, 1847–1857.
46
Weinberger, C.R., Cai, W., 2008. Surface-controlled dislocation multipli-
cation in metal micropillars. Proceedings of the National Academy of
Sciences 105, 14304–14307.
Wijnen, J., Peerlings, R.H., Hoefnagels, J.P., Geers, M.G., 2021. A discrete
slip plane model for simulating heterogeneous plastic deformation in single
crystals. International Journal of Solids and Structures 228, 111094.
Willot, F., 2015. Fourier-based schemes for computing the mechanical
response of composites with accurate local fields. Comptes Rendus
ecanique 343, 232–245.
Yaghoobi, M., Voyiadjis, G.Z., 2016. Size effects in fcc crystals during the
high rate compression test. Acta Materialia 121, 190–201.
Yang, W., Ruestes, C.J., Li, Z., Abad, O.T., Langdon, T.G., Heiland, B.,
Koch, M., Arzt, E., Meyers, M.A., 2021. Micro-mechanical response of
ultrafine grain and nanocrystalline tantalum. journal of materials research
and technology 12, 1804–1815.
Yu, Q., Mart´ınez, E., Segurado, J., Marian, J., 2021. A stochastic solver
based on the residence time algorithm for crystal plasticity models. Com-
putational Mechanics 68, 1369–1384.
Zeman, J., de Geus, T.W., Vondˇrejc, J., Peerlings, R.H., Geers, M.G., 2017.
A finite element perspective on nonlinear fft-based micromechanical simu-
lations. International Journal for Numerical Methods in Engineering 111,
903–926.
Zhang, H., Schuster, B.E., Wei, Q., Ramesh, K.T., 2006. The design of
accurate micro-compression experiments. Scripta Materialia 54, 181–186.
Zhang, L., Ohmura, T., Sekido, K., Hara, T., Nakajima, K., Tsuzaki, K.,
2012. Dislocation character transition and related mechanical response in
a body-centered cubic single crystal. Scripta Materialia 67, 388–391.
Zhang, X., Aifantis, K., 2011. Interpreting strain bursts and size effects in
micropillars using gradient plasticity. Materials Science and Engineering:
A 528, 5036–5043.
47
Zhu, T., Li, J., Samanta, A., Leach, A., Gall, K., 2008. Temperature and
strain-rate dependence of surface dislocation nucleation. Physical Review
Letters 100, 025502.
48
... Reproducing such experiments is also challenging for most of mesoscale and continuum methods, such as dislocation dynamics [26][27][28][29][30][31], the phase field method [32][33][34][35], stochastic discrete slip models [36], and stochastic crystal plasticity models [37]. These approaches have been widely used to study various deformation phenomena: DD provides insight into dislocation interactions and collective motion at the mesoscale, the phase-field method captures microstructural evolution during phase transformations and crack propagation, stochastic discrete slip models describe intermittent plasticity and size effects in small-scale materials, and stochastic crystal plasticity models enable large-scale simulations of plastic deformation in polycrystalline materials. ...
Article
Full-text available
Tailored control of the pattern forming instability that is initiated through selective dissolution in metal alloys has enabled the production of high-performing nanoporous metals. Nanoporous metals are “forged in fire,” evolving from exposure to harsh environments. This makes them uniquely adapted to extreme conditions with the potential for service in applications that require tolerance to corrosive environments, high-rate mechanical loading, and high-energy irradiation. Nanoporous metals are metastable structures and successful integration into extreme applications requires an atomistic understanding of how nanoporous morphologies evolve under relevant conditions. In this article, we explore our existing understanding of the evolution of nanoporous metals within extreme environments and how that understanding can be used to design mitigation strategies to drive operational longevity. Graphical abstract
Article
Full-text available
An FFT framework which preserves a good numerical performance in the case of domains with large regions of empty space is proposed and analyzed for its application to lattice based materials. Two spectral solvers specially suited to resolve problems containing phases with zero stiffness are considered (1) a Galerkin approach combined with the MINRES linear solver and a discrete differentiation rule and (2) a modification of a displacement FFT solver which penalizes the indetermination of strains in the empty regions, leading to a fully determined equation. The solvers are combined with several approaches to smooth out the lattice surface, based on modifying the actual stiffness of the voxels not fully embedded in the lattice or empty space. The accuracy of the resulting approaches is assessed for an octet-lattice by comparison with FEM solutions for different relative densities and discretization levels. It is shown that the adapted Galerkin approach combined with a Voigt surface smoothening was the best FFT framework considering accuracy, numerical efficiency and h-convergence. With respect to numerical efficiency it was observed that FFT becomes competitive compared to FEM for cells with relative densities above ≈7%. Finally, to show the real potential of the approaches presented, the FFT frameworks are used to simulate the behavior of a printed lattice by using direct 3D tomographic data as input. The approaches proposed include explicitly in the simulation the actual surface roughness and internal porosity resulting from the fabrication process. The simulations allowed to quantify the reduction of the lattice stiffness as well as to resolve the stress localization of ≈ 50% near large pores.
Article
Full-text available
Dislocations in high entropy alloys (HEAs) are wavy and have natural pinning points due to the variable chemical and energetic landscape surrounding the dislocation core. This can influence the critical shear stress necessary to initiate dislocation motion and the details associated with sustained dislocation glide. The objective of this work is to determine the relationship between Schmid shear stress and dislocation velocity in single phase FCC FeNiCrCoCu HEAs using molecular dynamics simulations, with comparisons made to dislocation motion in homogeneous Ni and Cu. Simulations are performed for four different dislocation character angles: 0° (screw), 30°, 60° and 90° (edge). Several key differences are reported, compared to what is previously known about dislocation motion in homogeneous FCC metals. For example, the drag coefficient B in the phonon damping regime for HEAs has a nonlinear dependence on temperature, whereas this dependence is linear in Ni. Mobility relationships between different types of dislocations common in homogeneous FCC metals, such as the velocity of screw and 60° dislocations being lower than edge and 30° dislocations at the same shear stress, do not necessarily hold in HEAs. Dislocation waviness is measured and is found to correlate with the ability of dislocations to glide under an applied shear stress, including the temperature dependence of the drag coefficient B. These results confirm that the influence of HEA chemical complexity on dislocation motion is important and this data can be used to guide development of analytical or empirical models for dislocation mobility in HEAs.
Article
Full-text available
The deformation of crystalline materials by dislocation motion takes place in discrete amounts determined by the Burgers vector. Dislocations may move individually or in bundles, potentially giving rise to intermittent slip. This confers plastic deformation with a certain degree of variability that can be interpreted as being caused by stochastic fluctuations in dislo-cation behavior. However, crystal plasticity (CP) models are almost always formulated in a continuum sense, assuming that fluctuations average out over large material volumes and/or cancel out due to multi-slip contributions. Nevertheless, plastic fluctuations are known to be important in confined volumes at or below the micron scale, at high temperatures, and under low strain rate/stress deformation conditions. Here, we develop a stochastic solver for CP models based on the residence-time algorithm that naturally captures plastic fluctuations by sampling among the set of active slip systems in the crystal. The method solves the evolution equations of explicit CP formulations, which are recast as stochastic ordinary differential equations and integrated discretely in time. The stochastic CP model is numerically stable by design and naturally breaks the symmetry of plastic slip by sampling among the active plastic shear rates with the correct probability. This can lead to phenomena such as intermittent slip or plastic localization without adding external symmetry-breaking operations to the model. The method is applied to body-centered cubic tungsten single crystals under a variety of temperatures, loading orientations, and imposed strain rates.
Article
Full-text available
Tungsten is considered as the plasma-facing material in the nuclear fusion device, which has to withstand coupled irradiation, mechanical and thermal conditions. To guarantee the safe operation of energy devices, it is essential to accurately predict its response. As a typical body center cubic material, the plasticity of tungsten is believed to exhibit the strong temperature dependence and non-Schmid effect. It is far from well understood how these plasticity features influence the behavior of irradiated tungsten. To disclose this mystery, a crystal plasticity model is developed, which considers the kink-pair mechanism of screw dislocations, the contribution of edge dislocations, and the interaction between dislocations and irradiation defects. The model is firstly verified through comparing with the available experimental results at different temperatures, and then used to disclose the temperature dependence of deformation localization in irradiated tungsten. It is found that post-irradiation mechanical tests at room temperature underestimate the occurrence of dislocation channel and deformation localization at higher temperatures, while in the low temperature regime, the thermal-activated dislocation mobility law may lead to a short-term deformation localization without clear dislocation channels. Conditions for the localization of deformation in irradiated tungsten are discussed, which is hoped to guide the reliable design of tungsten components used in fusion and fission reactors.
Article
Full-text available
In small-scale mechanical tests, such as micropillar compression tests, plastic deformation is often localized in narrow slip traces. These slip traces result from a few dislocation sources with relatively low nucleation stresses that are present in the material. In order to accurately simulate such small-scale experiments, the stochastics of the underlying dislocation network must be taken into account, which is usually done by performing discrete dislocation dynamics simulations. However, their high computational cost generally restricts these simulations to small and simple geometries and small applied displacements. Furthermore, effects of geometrical changes are usually neglected in the small strain formulation adopted. In this study, a discrete slip plane model for simulating small-scale experiments on single crystals is proposed, which takes the most important characteristics of dislocation plasticity for geometries in the micrometer range into account, i.e. the stochastics and physics of dislocation sources. In the model, the properties of all lattice planes are sampled from a probability density function. This results in a heterogeneous ow stress within a single crystal, unlike the uniform properties assumed in conventional crystal plasticity formulations. Moreover, the slip planes can be grouped together in bands via a weakest-link principle. The resulting equations are implemented in a standard crystal plasticity finite element model, using a finite deformation formulation. Within this setting, only the collective dislocation motion on glide planes is modeled, resulting in a significantly lower computational cost compared to frameworks in which the dynamics of individual dislocations are considered. This allows for simulating multiple realizations in 3D, up to large deformations. A small case study on micropillar compression tests is presented to illustrate the capabilities of the model.
Article
Deformation twinning is an essential plastic deformation mechanism that realizes the trade–off between strength and ductility. Twins nucleate and grow by the coordinated slip of partial dislocations on consecutive {111}–type slip planes. The route of twin growth is responsible for the evolution of twin morphology, which affects the non–coplanar dislocation slips by adjusting the twinning–associated mean free path. Incorporating such twinning mechanisms is critical for the accurate modelling and simulation of deformation behavior. In this study, a discrete dislocation plasticity (DDP) model was developed by integrating the source introduction methods and source activation criteria of Shockley partial dislocation. In the model, two twin nucleation mechanisms, i.e., the internal source and surface source, were considered concurrently, and the additional effect of stacking fault energy on the motion of partial dislocations was introduced. The evolution of partial dislocation slip–mediated deformation twins in micron–sized pillars of twinning–induced plasticity steel under uniaxial compression was investigated. The predicted twin morphologies and stress–strain curves from DDP simulation both agreed well with the experimental results, highlighting the inherent characteristics of partial–dislocation–based twinning behavior. The simulation results showed that the formation of nanometer–sized sharp twin tips was caused by the strong interaction between the front and rear dislocations on adjacent slip planes. In addition, a novel analytical model verified with the DDP simulation was proposed by considering the kinetics of the newly formed twin embryos. The competition between new twin activation and near–twin merging in determining the evolution of twin thickness was analyzed using the analytical model. The dependence of flow stress and twin morphology on the density and distribution of internal sources was demonstrated by considering the new twinning route. This research thus advances the understanding of partial dislocation slip–mediated twinning mechanisms.