Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Available online at www.sciencedirect.

com

Acta Materialia 61 (2013) 3347–3359


www.elsevier.com/locate/actamat

Shear transformation zone dynamics model for metallic


glasses incorporating free volume as a state variable
L. Li a, E.R. Homer b, C.A. Schuh a,⇑
a
Department of Materials Science and Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA
b
Department of Mechanical Engineering, Brigham Young University, 435 CTB, Provo, UT 84602, USA

Received 20 November 2012; accepted 13 February 2013


Available online 27 March 2013

Abstract

A mesoscale model, shear transformation zone dynamics (STZ dynamics), is employed to investigate the connections between the
structure and deformation of metallic glasses. The present STZ dynamics model is adapted to incorporate a structure-related state var-
iable, and evolves via two competing processes: STZ activation, which creates free volume, vs. diffusive rearrangement, which annihilates
it. The dynamical competition between these two processes gives rise to an equilibrium excess free volume that can be connected to flow
viscosity via the phenomenological Vogel–Fulcher–Tammann relation in relaxed structures near the glass transition temperature. On the
other hand, the excess free volume allows glasses to deform at low temperatures via shear localization into shear bands, even in the pres-
ence of internal stress distributions that arise upon cooling after processing.
Ó 2013 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Metallic glasses; Shear transformation zone; Free volume; STZ dynamics model; Shear bands

1. Introduction the deformation of metallic (and other) glasses. STZs are


essentially local clusters of a few atoms that can rearrange
Metallic glasses exhibit a great variety of deformation collectively in response to a shear stress. This local rear-
behaviors, depending upon conditions [1]. At low tempera- rangement involves not only the displacement of atoms,
ture and high stress, the deformation is highly inhomoge- but also an anelastic reconfiguration of atomic neighbors
neous, and the plastic strain is localized into shear bands. and a redistribution of free volume within the atomic clus-
In contrast, at high temperature and low stress, deforma- ter. This free volume redistribution is not only a transient
tion is homogeneous, exhibiting Newtonian flow rheology process, but is also believed to involve local, permanent
that gives way to exponential rheology as the stress rises. changes to the excess free volume. The local accumulation
Homogeneous flow is also associated with significant tran- of excess free volume is believed to facilitate shear localiza-
sients that follow any change in conditions (i.e., stress). The tion through local softening in the vicinity of previously
details of how all these diverse behaviors are connected in deformed regions [2].
metallic glasses pertain to the underlying deformation The concept of free volume as an important internal
mechanism of local shear shuffling in the amorphous struc- state variable for glasses pre-dates the STZ description of
ture. In particular, the shear transformation zone (STZ) glass deformation, beginning with developments by Cohen
originally proposed by Argon [2] and further observed and Turnbull [8,9]. They laid the groundwork for scalar
and studied widely in computational simulations [3–7], free volume evolution equations that have been widely
has been accepted as a “unit process”, which underlies and successfully applied by experimentalists to explain
relaxation behavior during glass formation from the
⇑ Corresponding author. Tel.: +1 617 452 2659. super-cooled liquid regime [10–12]. In addition, Spaepen
E-mail address: schuh@mit.edu (C.A. Schuh). [13] extended this approach to glass deformation on the

1359-6454/$36.00 Ó 2013 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.actamat.2013.02.024
3348 L. Li et al. / Acta Materialia 61 (2013) 3347–3359

basis of two competing processes mediated by free volume: according to a Boltzmann probability, with local biasing
free volume creation due to atomic motion driven by a from the stress state. As a consequence, the model incorpo-
shear stress vs. free volume annihilation via diffusive rear- rates a time scale and at the same time thermal effects,
rangement. This extended phenomenological model is which are missing from other mesoscale models. The acti-
capable of capturing the homogeneous and heterogeneous vation of STZs leads to stress and strain redistribution,
deformation modes in metallic glass, and is especially use- which is solved accurately via the finite element method
ful for describing deformation transients. (FEM). The updated elastic field resulting from STZ acti-
Both the STZ (as a fundamental unit process for glass vations in turn affects the rate of subsequent activation
deformation) and free volume (as an internal state variable events and thus the dynamical evolution of the system.
that evolves with time, temperature, stress and position) With this framework, the model is able to capture the basic
are simplified pictures of what actually happens in a glass. behaviors of metallic glasses, including high-temperature
There is a vast spectrum of STZs and STZ-like atomic rear- homogeneous flow, and low-temperature strain localiza-
rangements that occur in deforming glasses, and this spec- tion into shear bands.
trum is only poorly understood at present [14–16]. And One limitation of STZ dynamics as it has been imple-
“free volume” is often used as a catch-all internal variable mented to date, however, involves the use of a single, fixed
that subsumes a far greater complexity of internal states, energy barrier for STZ activation, which cannot capture
including chemical and topological short- and medium- the mechanical effects of structural evolution (e.g., local
range order, atomic stresses and strains [17–19]. Neither structural hardening or softening) as the system evolves.
picture (STZ or internal state variable) is yet fully worked One consequence of this limitation is that autocatalytic
out, and it is for this reason that simulation work on metal- accumulation of structural change (as in the dynamic soft-
lic glasses is essential to progress on understanding of glass ening that leads to shear localization) is difficult to trigger.
deformation [20]. There is an inherent lack of direct exper- In fact, when disorder pre-exists in the internal stress fields
imental approaches to these events in atomic glasses and, of the system (e.g., due to cooling from the liquid state), the
as such, atomistic simulations provide many details of plas- dynamical correlations between STZ activations are not
ticity in metallic glasses that would otherwise be unavail- large enough to overcome the internal stress fields (struc-
able [21]. Yet, even though these simulations provide tural disorder) to allow STZ activations to correlate and
significant detail, they suffer from a limitation in both form shear bands [26,31,32]. In prior STZ dynamics simu-
length and time scales (typically a few tens of nanometers lations shear bands are only seen when the initial configu-
during a few tens of nanoseconds). Atomistic methods ration is uniform and free of structural noise [26], which is
are therefore less useful for investigating large-scale events neither realistic for a glassy material nor in line with the
that occur over a long time scale, e.g., glass formation by perception that local structural softening is responsible
cooling from the melt at realistic rates. for shear banding in glasses in the first place.
In this context, mesoscale modeling methods are needed In this paper, our purpose is to continue the develop-
to bridge the gap between atomistic simulations and exper- ment of the STZ dynamics modeling framework by incor-
iments by averaging out atomistic effects and accounting porating a structural state variable into it. In line with
only for an ensemble of characteristic events. Such the above discussion, we introduce excess free volume into
approaches, on the one hand, enable simulations to access the STZ dynamics model of Homer and Schuh [26] as a
larger scales and, on the other hand, retain a description of simple internal parameter that can track local structural
elementary deformation physics. For example, Baret et al. change and allow for dynamical softening and hardening.
[22] proposed a depinning model to investigate long-range We allow system evolution via two competing processes:
spatial and temporal correlations in “depinned” local shear free volume creation via STZ activation vs. free volume
transition events. The critical stresses that lead to depin- annihilation via diffusive rearrangement; the treatment is
ning can involve a stochastic distribution in order to in the spirit of the original works by Argon [2] and Spaepen
account for the structural disorder of metallic glass. By [13]. In what follows, we first present our basic modeling
introducing an age parameter that can vary the yield stress framework followed by a two-dimensional (2D) FEM
distribution, the authors were able to show an age-depen- implementation. Then we explore the interplay between
dent shear banding behavior. A fluidity model developed evolution of excess free volume and metallic glass deforma-
by Picard [23,24] associated the characteristic transition tion over a range of thermal and mechanical loads. In par-
times with deformation and relaxation and produced com- ticular, we show that excess free volume-assisted shear
plex spatio-temporal patterns of transformation events. banding occurs at low temperatures, even in the presence
“STZ dynamics” is another mesoscale model proposed of structural disorder in “processed” samples.
by Homer et al. [25–27]. In this model a simulated metallic
glass is partitioned into an ensemble of potential STZs that 2. Modeling framework
are mapped onto a finite element mesh. The activation of
individual STZs is selected via the Kinetic Monte Carlo The STZ dynamics model is built upon that of Homer
(KMC) algorithm originally introduced by Bulatov and et al. [25–27] and uses the same general apparatus,
Argon [28–30], which uses an energy-based yield criterion consisting of a finite element mesh and a KMC algorithm.
L. Li et al. / Acta Materialia 61 (2013) 3347–3359 3349

Following Homer’s 2D implementation, we define STZs on In the following subsections, we explain our model for
a 2D irregular triangular mesh, with each element and its accumulation and loss of free volume, followed by how
immediate surrounding neighbors representing a potential these are implemented in the KMC algorithm.
STZ, as illustrated by A and B in Fig. 1. At any given time
step, one STZ is selected according to the KMC algorithm 2.1. Free volume and STZ activation
to participate in an activation event based on its local con-
ditions (e.g., stress and temperature). Our main adaptation Free volume and STZ activation are inherently linked in
of the Homer model is to add an internal state variable, our proposed model, and in fact the link is bidirectional:
excess free volume. It is represented as a normalized vol- STZs activation causes free volume accumulation, while
ume fraction fv and can vary from 0 to 1, i.e., fv = 0 corre- free volume in turn affects the activation of STZs. We treat
sponds to no excess free volume above the average these two aspects of the relationship in turn.
polyhedral volume V* in a dense random hard sphere glass, Following the view of Spaepen [13], we envision that the
while fv = 1 is an upper bound corresponding to a state dilatation required during STZ activation is not totally
where an STZ can be activated without accumulating extra transient in that it need not be eliminated immediately after
free volume [2]. The upper bound free volume f V* is only a the transition; rather, it remains in a post-activation state
fraction of V*. The factor f can vary from 0.1 to 0.6 for until it may be reduced by diffusive rearrangement. Conse-
liquid metals based on Cohen’s calculation [8]. quently, each STZ activation results in excess free volume
In our implementation, free volume is a local parameter, creation given as
and has a single value for each element in the FEM mesh.
In the most general sense, it should be regarded simply as Dfvþ ¼ ev ð1  fv Þ ð1Þ
an internal state variable that characterizes “damage” in
the structure, or the structural distance from equilibrium; This expression is derived from that of Argon [2], where ev
it need not be interpreted in a literal sense as a volume. is the local transformation dilatation, and (1  fv) corre-
In fact, unlike true free volume, our fv does not flow sponds to a first-order correction dictating that more di-
through the system from one material point to another; it lated regions accumulate less free volume upon STZ
is locally created or annihilated at each point, with no activation; a very rarefied volume of material with fv = 1
regard for volume balance. Consequently, one can envision can shear without any additional volume dilatation. When
more complex models or state variables, but the present an STZ is activated, the increase in free volume in each of
choice is viewed as presenting a good compromise between its elements is calculated according to Eq. (1).
simplicity and capturing essential physical features of glass The process of STZ activation is also affected by the
structural evolution. presence of free volume, although it otherwise follows the
framework developed by Homer and Schuh [26]. STZ acti-
vation is treated as an Eshelby inclusion problem [33]; that
is, following activation, the STZ transformation strain is
accommodated elastically by both the STZ and its sur-
rounding matrix. The activation barrier for shearing an
STZ is taken from the model proposed by Argon [2] and
Argon and Shi [34]:
7  5t
DF STZ ðfv Þ ¼ DF shear þ DF v0  gstz ðfv Þ ¼ lðT Þc20 X0
30ð1  tÞ
 
2ð1 þ tÞ 1 ^s
þ lðT Þe2v X0 þ lðT Þc20 X0  gstz ðfv Þ
9ð1  tÞ 2c0 lðT Þ
ð2Þ
where t, l(T) and X0 represent Poisson’s ratio, tempera-
ture-dependent shear modulus and volume of STZ, respec-
tively. The activation energy DFSTZ, in turn, can be
decomposed into two parts: a fv independent part DFshear,
which is the strain energy associated with an STZ shearing
by the characteristic shear strain c0; and a fv dependent part
DFv0gSTZ(fv), in which we factor out the excess free volume
Fig. 1. 2D FEM representation of a simulated sample. Region A is an dependence into a function gSTZ(fv). DFv0 incorporates the
example of a 13-element STZ that is prevalent in this 2D meshing. Region contributions from two parts: (1) a strain energy for a tem-
B is an exception that has 14 elements. Pure shear is applied when the porary dilatation to allow the atoms to rearrange into the
sample is under mechanical loading. This image also shows the redistri-
sheared position; and (2) an energy required to freely shear
bution of von Mises equivalent stresses (shaded) when STZ A and B are
activated. an STZ, where ^s is the interatomic shear resistance. Both
3350 L. Li et al. / Acta Materialia 61 (2013) 3347–3359

the dilatation energy and shear resistance can be reduced Here, the prefactor (1  fv) represents another correction
by fv through gSTZ(fv) [2], defined as for reducing the annihilation rate in proportion to the
2   amount of free volume available to annihilate, while mD is
gSTZ ðfv Þ ¼ f1 þ a½ev ð1  fv Þ g= 1 þ ae2v ð3Þ the frequency factor for diffusive rearrangement.
where a is a proportionality factor (0.8 [2]). The quadratic 2.3. Kinetic Monte Carlo
(1  fv)2 dependence comes from the fact that the activa-
tion free enthalpy increment due to a net activation dilata- A KMC algorithm is employed to control the evolution
tion ev(1  fv) is proportional to the square of this of the system through a series of individual events selected
dilatational strain term. Therefore, gSTZ(fv) varies between according to their corresponding activation rates. The
1 and 1/(1 + aev2) as fv goes from 0 to 1. The activation bar- algorithm essentially follows the form originally proposed
rier DFSTZ = DFshear + DFv0 used in the previous imple- by Bulatov and Argon [28–30] and modified by Homer
mentation of the STZ dynamics model by Homer et al. and Schuh [26], but here we introduce two types of events
[25–27] is retrieved when fv = 0. (STZ activation vs. diffusive rearrangement) that are all
With a free volume dependent activation energy enumerated and compete with each other through their
DFSTZ(fv), the activation rate of a single potential STZ is activation rates. In each KMC step:
given by
  1. The activation rates for both processes s_ STZ;i and s_ D;i
DF STZ ðfv Þ  sc0 X0 =2
s_ STZ ¼ mSTZ exp  ð4Þ at each potential site i are determined by Eqs. (4)
kT
and (7), respectively.
where s and T are the local shear stress and temperature, 2. These individual rates are then accumulated into a
respectively; mSTZ represents the attempt frequency along list and summed over the list such that the cumula-
the reaction pathway and is on the order of the Debye fre- tive activation rate s_ tot gives the rates for both pro-
quency; and k is Boltzmann’s constant. cesses over all i. For each site i, there are two
normalized rates, which are given by
2.2. Free volume annihilation via diffusive rearrangement gSTZ;i ¼ s_ STZ;i =_stot ; gD;i ¼ s_ D;i =_stot ð8Þ

Existing excess free volume can also be annihilated such that


X
through diffusive rearrangement. For the purposes of our ðgSTZ;i þ gD;i Þ ¼ 1 ð9Þ
STZ dynamics model, a diffusive rearrangement is viewed i
as a local event that involves annihilation of excess free vol-
3. Two random numbers n1 and n2 are generated
ume as
within a uniform distribution on the interval (0, 1).
Dfv ¼ ev fv ð5Þ 4. n1 is employed to determine the elapsed time of the
system by
This form is also suggested by Argon [2], and again con-
tains a correction that acknowledges the fact that free vol- Dt ¼  ln n1 =_stot ð10Þ
ume can only be annihilated in proportion to how much of 5. n2 is to select the particular process and location for
it is present in the first place; if fv = 0, no free volume the next event. This is done by comparing the ran-
reduction is possible. In the present implementation, the lo- dom number n2 with the list of activation rates. If
cal diffusional rearrangement associated with Eq. (5) oc- the random number falls in the range
curs over small regions of exactly the same size and Xk1 X
k1
shape as potential STZs, and when such an event is trig- ðgSTZ;i þ gD;i Þ < n2 6 ðgSTZ;i þ gD;i Þ þ gSTZ;k ð11Þ
gered, Eq. (5) is applied to every element in the selected i¼1 i¼1
zone (without applying any shape change). Then shear distortion c0 is applied to the selected re-
The activation energy required to trigger a single event gion i in a particular direction, determined by the
of excess free volume annihilation can be written as location of the random number n2 on the subinter-
DGD ðfv Þ ¼ DGv0 gD ðfv Þ ð6Þ val associated with the stress state [26]. Addition-
ally, fv increases in the region by an amount
where DGv0 is the activation free energy for a net diffusive
determined by Eq. (1). Otherwise, if the random
rearrangement in a region with fv = 0. gD(fv) is a factor that
number falls in the range
lowers the energy barrier in a region containing excess free
volume, which, again following Argon [2], is taken as Xk1 X
k
ðgSTZ;i þ gD;i Þ þ gSTZ;k < n2 6 ðgSTZ;i þ gD;i Þ ð12Þ
gD(fv) = (1  fv). i¼1 i¼1
Therefore, the activation rate of a single diffusive rear-
rangement event can be written as Then the diffusive rearrangement process for a partic-
  ular activation region k is selected. In this case, fv de-
DGD ðfv Þ creases in the activation region k by the amount
s_ D ¼ ð1  fv ÞmD exp  ð7Þ
kT determined by Eq. (5).
L. Li et al. / Acta Materialia 61 (2013) 3347–3359 3351

6. After a transition occurs, either STZ activation or 2.5. Discussion of the modeling approach
diffusive rearrangement, new stress and strain fields
are solved by FEM. At present, the newly introduced excess free volume rep-
resents a relatively simple way to incorporate a structural
These KMC steps can be repeated as many times as state variable into the STZ dynamics model. As will be
desired to simulate a given process, and the stochastic nat- shown in this work, this simple addition provides signifi-
ure of the algorithm will produce a realistic outcome as cant insight into the flow of excess free volume and its
long as the rates governing the individual events are cor- effects on glass deformation. However, the present imple-
rect. The two possible events described in Step 5 are exclu- mentation does have certain limitations that need to be
sive; in each KMC increment, only one of them will be addressed as knowledge of atomic scale deformation in
selected. glasses improves. For example, the present model creates
and annihilates excess free volume locally; future work
2.4. 2D FEM implementation must focus on the diffusion of this free volume rather than
the creation and annihilation thereof. Additionally, the
We implement the STZ dynamics model in the commer- excess free volume is a phenomenological variable like
cial finite element package ABAQUS via its user subrou- those used in most of the mesocale models [23,31,32,41];
tines, and its FEM solver is employed to determine the one could develop similar approaches that connect to alter-
stress and strain fields as well as the evolution of the state native state variables such as local elastic properties [42],
variable (i.e., excess free volume) for each KMC increment. local bonding such as icosahedral or non-icosahedral
Fig. 1 shows the 2D FEM mesh used for the present sim- effects [14], or dynamical state variables such as an effective
ulations. Plane-strain quadratic triangular elements disorder temperature [43–45]. Each of these state variables
(CPE6MT) are employed, and the average potential STZ provides something unique to describe structural disorder
or diffusive rearrangement region contains 13-elements of glasses, and the complex structural evolution in metallic
(refer to STZ A in Fig. 1), with each potential STZ centered glasses provides freedom in the selection and description of
on an element and including all the neighbors that share its variables to describe similar behaviors.
nodes. The number of potential STZs is thus equal to the
number of elements, so there are total of 14,846 potential 2.6. Simulation parameters
STZs defined over 14,846 elements similar to that in
Fig. 1. The validation of this geometrical STZ mapping Table 1 summarizes the parameters used in the STZ
onto the FEM mesh is discussed by Homer and Schuh dynamics simulations. For the sake of simplicity, it is
[26], who showed that the 13-element STZ gives a reason- assumed that the STZ attempt frequency is equal to that
able compromise between solution accuracy (in terms of for diffusive arrangement (tSTZ = tD). The activation bar-
convergence) and computational speed. It is also worth rier for diffusive rearrangement is taken to be equal to
noting that any element can participate in multiple STZs, the excess free volume-dependent part of STZ activation
which is physically reasonable, since atoms may participate energy, i.e., DGv0 = DFv0, considering that it is a volume-
in multiple STZ events. Under this mapping rule, there are related process. Additionally, the mechanical properties
a few STZs that may contain more or less than 13 elements of Vitreloy 1 derived from experiments [35] are used.
(refer to STZ B in Fig. 1), depending on the local mesh.
For thermal excursions, the boundary conditions as 3. Model output
shown in Fig. 1 are applied, with no tractions on the sam-
ple surfaces. For mechanical loading, a case of pure shear is We now exercise the model to briefly examine the role of
applied, and a corner node is fixed to remove the rigid body excess free volume in metallic glass deformation, with the
motion (refer to Fig. 1). During the simulations, tempera- intent of evaluating its reasonableness. In this section, we
tures are kept uniform throughout the system. first investigate the local evolution of excess free volume
at the STZ level via the creation and annihilation processes.
Table 1
List of simulation parameters used in free volume-assisted STZ dynamics simulations.
Simulation parameters Value
STZ attempt frequency tSTZ = 1.0193  1012 s1 (ref. to Eq. (4))
Diffusive arrangement frequency tD = 1.0193  1012 s1 (ref. to Eq. (7))
Free volume-independent STZ activation energy DFshear = 0.9  1030 J Pa1 l(T) (ref. to Eq. (2))
Free volume-dependent STZ activation energy DFv0 = 8.1  1030 J Pa1 l(T) (ref. to Eq. (2))
STZ shear strain c0 = 0.1 (ref. to Eqs. (2) and (4))
STZ dilatational strain ev = 0.5 (ref. to Eqs. (1), (3), and (5))
STZ volume X0 = 1.6 nm3 (ref. to Eqs. (2) and (4))
Temperature-dependent shear modulus l(T) = 37 GPa–0.004 GPa K1 T (ref. to Eq. (2))
Poisson’s ratio m = 0.352 (ref. to Eq. (2))
3352 L. Li et al. / Acta Materialia 61 (2013) 3347–3359

Samples are then processed via simulated heat treatment to later. The history of local free volumes is replete with such
obtain relaxed structures at various temperatures. Finally, peak-and-decay patterns.
samples with various internal structures are deformed Even once deformation accelerates during a localization
through creep tests at high and low temperatures, and event, the same local pattern is revealed; near the time
the effect of excess free volume on the deformation pro- marked by A, the evolution of the excess free volume accel-
cesses is analyzed. erates, and numerous creation and annihilation events are
selected within very short time intervals; the inset in
3.1. Local free volume evolution Fig. 2b shows that this selected STZ is activated four times,
with numerous relaxations in-between over the course of
We start the simulations by tracking the performance of this acceleration period. This acceleration in local evolu-
the two competing processes, i.e., free volume creation vs. tion of fvele is directly associated with the macroscopic
annihilation. For a simple demonstration, a sample initially strain evolution, as shown in Fig. 2a, where we observe
at a state of zero internal stress and zero excess free volume that this is the time where strain begins to rapidly localize
is loaded with a shear stress of 1 GPa at 300 K. Fig. 2a dis- into a shear band.
plays the resultant evolution of the macroscopic shear This exercise illustrates that fvele is determined by
strain and the volume-average excess free volume fv: fol- dynamic balancing between creation and annihilation.
lowing a transient period, both shear strain and fv rapidly Annihilation occurs more frequently because its activation
increase as a consequence of strain localization into a shear energy barrier is lower and is more significantly reduced by
band, which is illustrated in the insets (i.e., the spatial the existing fvele (refer to Eq. (6)). Usually, one creation
distribution of STZ strain and excess free volume at event is followed by two to four annihilation events, which
time A). is consistent with Spaepen’s assessment that the number of
For further details, we zoom in and explore the local diffusive jumps necessary to annihilate excess free volume is
free volume evolution of a randomly selected element in between 1 and 10 [13].
the shear band (marked as a star in the insets in Fig. 2a),
fvele , the behavior of which is displayed in Fig. 2b. Over 3.2. Thermal treatment (structural relaxation)
the course of the simulation, this particular element only
occasionally participates in an STZ or diffusive event, so Owing to their disordered structure, metallic glasses
there are many fallow periods in which other activity is contain a distribution of internal stresses [14] and, in STZ
occurring. For example, this STZ is the third to be acti- dynamics simulations, such distributions arise entropically
vated at the outset of the simulation and, at a time of upon thermal processing. The development of such a distri-
40 Gs after loading, fvele dramatically increases from 0 bution from an initially stress-free state is illustrated in
to 0.5 upon this first activation. However, immediately Fig. 3a, where the instantaneous elastic strain energy den-
after, fvele drops by half to 0.25, and then after a short per- sity in the system is plotted as a function of time for various
iod drops by half again to 0.125 as a result of annihilation annealing temperatures, under zero applied stress. Fig. 3a
processes. This is a pattern that is common in the present provides a comparison of the timescales required to anneal
model, where a single deformation event causes large rises at different temperatures as well as the equilibrium strain
in excess free volume, in turn leading to a significantly energy density that is achieved at each temperature. Note
reduced activation barrier for relaxation some short time that the logarithmic scale on the time axis obscures the

Fig. 2. (a) The macroscopic shear strain and volume-average fv evolution with respect to time. The insets show the contour plot of the spatial distribution
of plastic STZ strain and fv at the time marked A. The position of the selected element is marked as a star in the contour plots. (b) The local fv evolution of
the selected element over a series of activation events (marked by the points). The inset is a magnified view at marker A. This test is conducted at constant
shear = 1 GPa and 300 K upon a stress-free sample with zero initial excess free volume.
L. Li et al. / Acta Materialia 61 (2013) 3347–3359 3353

Fig. 3. (a) The elastic strain energy vs. time for annealing at various temperatures. The inset shows an example at 650 K to confirm the convergence of this
process to a steady state. (b) The elastic energy density in the equilibrated state vs. temperature.

achievement of a steady state for each annealing tempera- Fig. 4b. This asymmetric distribution is somewhat like
ture, which can be better illustrated by plotting the strain the exponential distribution proposed in the free volume
energy density as a function of the KMC steps, as illus- theory [8] with a large fraction of small fvSTZ balanced by
trated by the inset for the 650 K simulation. A plot of a small fraction of large fvSTZ . Furthermore, the fvSTZ distri-
the equilibrium elastic strain energy density values for each bution in turn leads to a distribution of activation energy
of the simulations shows a linear relationship with the tem- barriers for the two competing processes, as shown in
perature, as displayed in Fig. 3b. This linear trend is well Fig. 4c. Both activation energy barrier distributions take
established from 500 to 1000 K and has an extrapolated on a similar shape, yet the diffusive process DGD is situated
intercept of 0 ± 200 kJ m3 at 0 K. at smaller energy values and has a broader distribution in
During annealing, the excess free volume evolves as well, comparison with the STZ activation barrier DFSTZ.
as shown for various annealing temperatures in Fig. 4a. The distributions in Fig. 4c reflect the disordered nature
Upon annealing, fv increases dramatically and saturates of the glass structure in our model. They are not explicitly
earlier than the corresponding elastic strain energy density. input to the model, but are emergent distributions that can
Thereafter, fv oscillates around an equilibrium value fv,eq, evolve with deformation and processing. While this is an
and the amplitude of oscillation increases with annealing interesting contrast to other mesoscale models [32] that
temperature as a consequence of thermal fluctuations. An include a distribution of events a priori, it also reveals a sig-
example of fv vs. KMC steps at 650 K is displayed as an nificant point of departure between the present simple
inset as a better illustration of the achievement of a steady model and the true internal structure of a metallic glass.
state. In comparison with the STZ activation energy distributions
The annealing process produces an asymmetric distribu- measured by Argon and Kuo via mechanical spectroscopy
tion of fvSTZ at steady state, as shown more explicitly in [36], our DFSTZ distributions have a similar shape, but they

Fig. 4. (a) The evolution of volume-average excess free volume fv vs. time during annealing at various temperatures. The inset shows an example at 650 K
plotted against KMC steps. (b) The distribution of fv of each STZ when reaching steady state at various temperatures. The inset shows an example of the
spatial distribution of fv at 650 K. (c) The activation energy distribution in steady state for STZ activation DFSTZ and diffusive arrangement DGD.
3354 L. Li et al. / Acta Materialia 61 (2013) 3347–3359

are very much narrower ( 0.1 eV) than in the experiments form of Eq. (13) with a reasonable match to experimental
(1 eV). Similarly, the activation energy distribution calcu- data, especially since the temperature dependences that
lated by the atomistic simulations [16] again exhibits an went into the model as input are not linear, but exponen-
energy spectrum that is much broader (1 eV). This result tial. To appreciate how a linear form can emerge, it is
points to a greater need for an improved state description, instructive to examine the equilibrium condition of the
which may ultimately be provided by multiscale modeling. present model, by setting as equal the rates of free volume
The equilibrium volume-average excess free volume fv,eq increase and decrease:
at various annealing temperatures is plotted in Fig. 5. The  
simulated fv,eq vs. temperature is strikingly linear, and can DF STZ  sc0 X0 =2
ev ð1  fv ÞmSTZ exp 
be well described by a fitted relationship: kT
 
fv;eq ¼ afv ðT  T 0 Þ ð13Þ DGD
¼ ev fv ð1  fv ÞmD exp  ð14Þ
kT
This form is the one familiar from free volume theory
[37,38], although it is important to note that there is no in- In the true equilibrated glass, there is a wide distribution of
put to the present model specifically built in the linear form local stress states and free volumes, but as the simplest
of Eq. (13). The slope afv = 7.4  105 K1 is a measure- approximation one may replace the local state variables
ment of increase in fv,eq with respect to temperature, and with global average values (s = 0 during annealing). Given
T0 = 280 K is the temperature at which fv,eq reaches 0. the prior assumption that mSTZ = mD, we then have
These values can be compared to those experimentally  
reported for Vitreloy 1 [11]. T0 = 426.3 K is the Vogel–Ful- DF STZ ðfv Þ  DGD ðfv Þ
fv ¼ exp  ð15Þ
cher–Tammann (VFT) reference temperature of Vitreloy 1, kT
which is somewhat higher than the present simulated value where both DFSTZ and DGD depend on fv. Eq. (15) provides
of 280 K. To make a comparison with regard to afv, we an implicit relationship between fv,eq and temperature, and
recall that the experimentally measurable quantity is av = is plotted in Fig. 5 using the simulation input parameters as
fafv, with f in the range 0.1–0.6. Correspondingly, the a red solid line. This simple equilibrium model agrees rea-
range for av from our simulations is 7.4  106 to sonably with the full simulation output and, most impor-
4.5  105 K1. For Vitreloy 1, av = aliq  aglass = 1.93 tantly, exhibits a broad apparently linear regime at higher
 105 K1 (i.e., the difference in thermal expansion coeffi- temperatures, much as the simulation results do. While this
cients between the liquid and the glass [37]), which falls in expression is not linear, it is instructive to see that, on the
the middle of the simulation range. relevant scales of the present problem, it can appear so; this
Considering that no assumption is made on the temper- provides some support for the use of Eq. (13) in describing
ature dependence of fv,eq in the simulations or their inputs, the free volume state of an equilibrated glass.
the value of afv is an emergent property of the model with We may linearize Eq. (15) by assuming (as in the simu-
important implications for glass structure and deforma- lations) that DFv0 = DGv0, and performing a Taylor expan-
tion. It is encouraging that the model can reproduce the sion about a characteristic temperature (T0) and free
volume (fv0). This leads directly to Eq. (13) with the follow-
ing values of the parameters:
kln2 ðfv0 Þ
afv ¼  ð16aÞ
1ae2v ae2v
DF shear =fv0 þ DF v0 1þae2v
ð1  lnðfv0 ÞÞ þ fv0 1þae 2 ð1  2 lnðfv0 ÞÞ
v

1ae2v 2
aev 2
DF shear ð1 þ lnðfv0 ÞÞ  DF v0 f
1þae2v v0
þ 1þae 2 ð1  lnðfv0 ÞÞfv0
v
T0 ¼ ð16bÞ
kln2 ðfv0 Þ

By choosing a reference free volume content fv0 = 5% that


would be generally characteristic of the super-cooled liquid
region, Eq. (16) gives afv = 9.3  105 K1 and T0 = 405 K.
This linearization is the one shown by the broken black line
in Fig. 5; the slope matches reasonably with the simula-
tions, while the characteristic temperature is somewhat
higher than that given by extrapolation of the simulation
Fig. 5. The equilibrium free volume at various temperatures from data (T0 = 280 K) but in better agreement with the experi-
annealing simulations, plotted along with a linear relationship (ref. to mental value (T0 = 426.3 K). The fact that the equilibrium
blue dashed line) fv;eq ¼ afv ðT  T 0 Þ, where afv = 7.44  105 K1 and free volume follows the expected empirical linear relation-
T0 = 280 K. The numerical solution of Eq. (15) is shown in red; and its ship with respect to temperature to some extent validates
linearization is in black with afv = 9.43  105 K1 and T0 = 405 K
obtained from Eq. (16) with reference fv0 = 5%. (For interpretation of
the present dynamical rules for the free volume evolution.
the references to colour in this figure legend, the reader is referred to the It also provides some insight into the possible origins of
web version of this article.) these empirical relationships, which in turn provides some
L. Li et al. / Acta Materialia 61 (2013) 3347–3359 3355

support for many continuum models that assume their as fv at the time marked t1 (in the transient) and t2 (nearing
validity a priori. the steady state) are provided in Fig. 6d and h, respectively.
At t1 in Fig. 6h, fv is distributed heterogeneously, with only
3.3. High-temperature rheology a fraction of the sample having increased values. In
contrast, at t2 the higher level of accumulated excess free
The high-temperature deformation behavior of the volume fv is distributed homogeneously throughout the
relaxed structures is studied over a range of stresses at dif- sample, as is the strain (refer to Fig. 6d).
ferent constant temperatures near and above Tg = 623 K For each of the different simulations in Fig. 6 and many
(the glass transition temperature of Vitreloy 1 [35]). In all others like them, the steady-state strain rate is assessed and
the simulations, pure shear traction is applied and held at compiled in Fig. 7a as a function of the applied stress. The
a constant value. Fig. 6a–c shows the shear strain vs. time data sets show typical homogeneous glass flow for metallic
data over a range of stresses at 650 K, and the correspond- glasses at high temperatures, that is, weakly rate-dependent
ing evolutions of fv are displayed in Fig. 6e–g. At lower Newtonian behavior at low stresses vs. non-Newtonian
stresses (Fig. 6a and e) a steady-state flow is established flow with gradually enhanced rate sensitivity at high stres-
almost instantaneously with a constant strain rate and ses. This high-temperature flow generally conforms to the
minor change in fv of the same scale as its thermal fluctua- classical one-dimensional (1D) model of independent
tion range. As stress increases to 500 MPa (Fig. 6b and f), forward-and-backward STZ activation [13]. A simple
the transients become more significant, especially in fv, combination of the forward-and-backward rates of Eq.
which shows a dramatic increase followed by a relatively (4) results in a hyperbolic-sine stress-dependence on the
quick saturation at a steady-state value. Since fv saturates steady-state strain rate
so quickly, the transient in structural adjustment does not    
DF STZ ðfv Þ sc0 X0
give rise to a noticeable transient in the strain response. c_ ¼ 2  c0  mSTZ exp  sinh ð17Þ
kT 2kT
As the applied stress continues to increase to 1 GPa
(Fig. 6c and g), a longer-lived transient in both variables In principle, this 1D model contains the same inputs that
is observed. The spatial distributions of STZ strain as well the present 2D simulations do, so it can be compared with

Fig. 6. (a)–(c) The strain vs. time data for the creep response of a relaxed structure at various shear stresses at constant temperature 650 K. (e)–(g) The
corresponding fv evolution over the course of the creep tests. (d) The snapshots of STZ strains at marked time t1 and t2 in (c) for the case of shear
stress = 1 GPa. (h) Two snapshots at the same time as (d) of the spatial distributions of fv.
3356 L. Li et al. / Acta Materialia 61 (2013) 3347–3359

Fig. 7. (a) Steady-state homogeneous flow data for several high-temperature simulations, plotted along with the predicted strain rates from Eq. (17). (b)
The equilibrium fv at steady state as a function of applied load for several high-temperature simulations, plotted along with the predicted values from Eq.
(18).

the simulation output without the use of any adjustable


parameters. This is presented in Fig. 7a, where the agree-
ment between the data and Eq. (17) is good, especially as
regards the general locations of the Newtonian regime at
low stresses and its divergence into exponential flow at
higher stresses.
Fig. 7b shows a plot of the equilibrium value of the vol-
ume average excess free volume fv,eq as a function of
applied stress at various temperatures. In the Newtonian
flow region, fv,eq barely changes with applied stress, but
there is a dramatic increase with applied stresses after
entering the non-Newtonian flow region. The temperature
effect on fv,eq is not at all significant on the scale here,
and the small difference that does exist diminishes with
increasing stress. On examination of the creation and anni-
hilation processes of excess free volume fv, one can write Fig. 8. The simulated viscosity in the Newtonian flow region at various
the time rate of change temperatures vs. the corresponding excess equilibrium free volume
fv,
plotted along with the fitted Doolittle equation, g ¼ g0 exp fbv , where
    g0 = 1.58 and b = 0.64.
dfv DF STZ ðfv Þ sc0 X0
¼ 2ev ð1  fv ÞmSTZ exp  cosh
dt kT 2kT We may further appreciate the structure–property con-
  nection of a deforming glass by considering the viscosity
DGD ðfv Þ
 2ev fv ð1  fv ÞmD exp  ð18Þ of glass flow in the super-cooled liquid region. Fig. 8 shows
kT
the viscosity in the Newtonian flow region at various
Note that the stress dependence of dfv/dt is hyperbolic-co- temperatures vs. the corresponding steady-state excess free
sine rather than sine in Eq. (17). This is due to the fact that volume fv. The viscosity is calculated as g ¼ s=_c from the
both forward-and-backward STZ operations can create ex- simulation data at 50 MPa in Fig. 7a. The apparently expo-
cess free volume. Steady state is achieved when dfv/dt = 0. nential form of the data in Fig. 8 is reminiscent of the clas-
The equilibrium fv,eq can be numerically solved from Eq. sical Doolittle equation [38]:
 
(18) and is plotted in Fig. 7d as solid lines. The match to b
the simulation data is good, although the predicted values g ¼ g0 exp ð19Þ
fv
are consistently higher, and this is accentuated at high
stress. These discrepancies are most likely related to the which is proposed to describe the flow of a glass in the
free volume distribution in the simulations; when there is homogeneous regime. Fitting the present data into the
a distribution of options, STZs prefer operating more fre- form of Eq. (19) gives g0 = 1.58 and b = 0.64. Further,
quently at the locations with large fv. These preferred oper- by taking into account the linear relationship between fv
ations, on the one hand, will give rise to a fast flow rate and temperature that emerged earlier in our analysis in
and, on the other hand, result in a small amount of fv incre- the form of Eq. (13), we obtain the Vogel-Fulcher-Tamman
ment (refer to Eq. (4)), and thus relatively small fv. The 1D (VFT) equation:
 
equilibrium model of Eq. (18) replaces the distribution with b
g ¼ g0 exp ð20Þ
an average free volume and thus misses this detail. afv ðT  T 0 Þ
L. Li et al. / Acta Materialia 61 (2013) 3347–3359 3357

As noted earlier with reference to Eq. (13), it is a non-trivial a localized band of accumulation (Fig. 9c), well before
result that our simulations should conform to VFT kinet- strain begins to accumulate in a significant sense. This
ics, as the inputs to the model do not explicitly contain localized fv band facilitates shear localization, and ulti-
the VFT parameters afv and T0 or the linear form of Eq. mately leads to a nascent shear band at the same location
(13). The VFT scaling is only apparently obeyed in our (Fig. 9d). Once the shear band forms, the localized shearing
simulations as the consequence of two competing pro- promotes rapid accumulation of plastic deformation (from
cesses; this behavior is emergent. Fig. 9d to e); meanwhile fv increases at a roughly constant
rate as the shear band widens. Over the entire process, the
3.4. Low-temperature deformation fv band propagates ahead of the shear band, which is fur-
ther illustrated by the 1D STZ and fv distribution plots in
Low-temperature deformation is examined in a struc- Fig. 9b–e. We conclude that excess free volume catalyzes
ture that has been cooled at experimental rates to quench shear localization. The softening that often precedes local-
in a certain degree of disorder. This structure is achieved ization into a shear band was observed previously when
by starting with one equilibrated at 650 K (>Tg = 623 K structural relaxation and local softening are accounted
of Vitreloy 1) and allowing the sample to evolve while it for [32,39,40].
is cooled in the absence of external stress to 300 K at a rate The ability to observe shear localization in samples with
of 10 K s1. The cooled structure is out of equilibrium pre-existing structure is an important improvement to the
because of the finite cooling rate, with an average excess STZ dynamics model [26]. Fig. 10 shows the shear strain
free volume fv = 0.026 that is significantly above that evolution of the same cooled structure under the same
expected in equilibrium (fv  0, cf. Fig. 5). The specimen loading conditions (i.e., at 300 K and 2 GPa) as those in
is then tested in pure shear at 2 GPa at 300 K. Fig. 9, but now after having turned off the free volume evo-
The resultant shear strain and volume average excess lution. As reported by Homer and Schuh [26], rather than
free volume fv are plotted as a function of time in forming a shear band, the sample undergoes homogeneous
Fig. 9a. A significant transient period is clearly observed deformation without the assistance of free volume. The
in both the shear strain and fv responses. Early in this per- addition of the structural state variable allows the system
iod (Fig. 9b), the internal stress distribution resulting from to localize as a result of structural memory that is retained
the thermal treatments inhibits STZs from activating in a even after the redistribution of stress, and allows the system
correlated fashion, which is reflected in the randomly dis- to facilitate correlated events to overcome the noise
tributed spots with large fv. These uncorrelated STZs acti- induced by the internal stresses and dictate individual
vations are not associated with the development of STZ operation into a shear band. In future work we hope
significant plastic shear strain along the loading direction. the model can be exploited to study shear band formation
Shortly thereafter, though, fv responds and develops into and propagation kinetics.

Fig. 9. (a) The shear strain and excess free volume fv vs. time data of a cooled structure deformed at 300 K and 2 GPa shear load. (b)–(e) correspond to the
snapshots at different times during the creep test. For each time, the physical deformation along with the magnitude of STZ strain and fv are displayed;
additionally, a plot with the 1D profile of STZ strain and fv distributions along the vertical direction of the deformed sample is provided.
3358 L. Li et al. / Acta Materialia 61 (2013) 3347–3359

classical free volume theory. The distribution of the


excess free volume results in distributions of the activa-
tion energy barriers of both STZ transformation and
diffusive arrangement. The activation energy distribu-
tions take on a shape similar to the experimentally mea-
sured ones [36], but are much narrower (0.1 eV) than
the measured ones (1 eV).
 In shear loading at high temperatures, the excess free
volume first increases and then reaches a shear-enhanced
equilibrium value in steady-state homogeneous flow. At
high stresses, the excess free volume gradually increases
to a steady-state value an order of magnitude larger
than the starting value. Both the shear rate and excess
Fig. 10. The shear strain vs. time data of the same cooled structure under free volume in the steady-state follow expectations based
the same deformation conditions as in Fig. 9 (i.e., at 300 K and 2 GPa on simple 1D analytical models used widely to analyze
shear load), but after turning off the free volume evolution. The insets glass rheology.
show the spatial distribution of STZ strain and the corresponding 1D  Free volume catalyzes shear localization. For a
profile of STZ strain along the sample vertical direction at the marked
time, both of which illustrate the homogeneous nature of the deformation.
quenched glass structure deformed at low temperature,
strain localization into a shear band directly follows
4. Conclusions excess free volume accumulation in that band. This
highlights a key improvement over previous STZ
We further developed the STZ dynamics simulation dynamics models, where localization could be sup-
method for glasses [25–27] by incorporating a structural pressed by internal stress distributions.
state variable that can allow for dynamical hardening
and softening of a disordered solid upon heating or defor-
mation. We specifically develop on the basis of excess free Acknowledgements
volume as the internal variable, and new features added to
the framework include the following. This work was supported partially by the US Army Re-
search Office through the Institute for Soldier Nanotech-
 Excess free volume evolves via two competing processes: nologies at MIT, and partially by the US Defense Threat
STZ activation that creates free volume vs. diffusive Reduction Agency through Contract No. HDTRA-11-1-
rearrangement that annihilates it. The activation unit 0062.
of the two processes is taken to be a coarse-grained col-
lection of atoms. References
 At the same time, the excess free volume can, in turn,
affect the two competing processes by modifying their [1] Schuh C, Hufnagel T, Ramamurty U. Acta Mater 2007;55:4067.
activation energy barriers. [2] Argon AS. Acta Metall Mater 1979;27:47.
[3] Srolovitz D, Vitek V, Egami T. Acta Metall Mater 1983;31:335.
[4] Deng D, Argon AS, Yip S. Philos T Roy Soc A 1989;329:549.
We implement in two dimensions, and a series of ther- [5] Falk ML, Langer JS. Phys Rev E 1998;57:7192.
mal and mechanical simulations were performed, illustrat- [6] Lemaı̂tre A. Phys Rev Lett 2002;89:195503.
ing the interplay between the excess free volume evolution [7] Lund AC, Schuh CA. Acta Mater 2003;51:5399.
and the metallic glass deformation. The key results include [8] Cohen MH, Turnbull D. J Chem Phys 1959;31:1164.
[9] Turnbull D. J Chem Phys 1970;52:3038.
the following:
[10] Busch R, Bakke E, Johnson WL. Acta Mater 1998;46:4725.
[11] Masuhr A, Busch R, Johnson WL. J. Non-Cryst Solids 1999;250–252
 Over the course of structural relaxation at high temper- [Part 2: 566].
atures, the excess free volume saturates faster than the [12] Masuhr A, Waniuk TA, Busch R, Johnson WL. Phys Rev Lett
corresponding elastic strain energy density. After reach- 1999;82:2290.
[13] Spaepen F. Acta Metall Mater 1977;25:407.
ing equilibrium, the excess free volume fluctuates
[14] Egami T. Prog Mater Sci 2011;56:637.
around an equilibrium value with a magnitude that [15] Rodney D, Schuh C. Phys Rev B 2009;80:184203.
increases with annealing temperature. [16] Rodney D, Schuh C. Phys Rev Lett 2009;102:235503.
 In equilibrium, the excess free volume takes an asym- [17] Spaepen F. Scripta Mater 2006;54:363.
metric distribution, in which a large fraction of small [18] Falk ML. Phys Rev B 1999;60:7062.
[19] Egami T. Intermetallics 2006;14:882.
values is balanced by a small fraction of large values.
[20] Rodney D, Tanguy A, Vandembroucq D. Model Simul Mater SC
In addition, a linear relationship can be established 2011;19:083001.
between the volume-average excess free volumes and [21] Falk ML, Maloney CE. Eur Phys J B 2010;75:405.
the annealing temperatures, which is in line with [22] Baret J-C, Vandembroucq D, Roux S. Phys Rev Lett 2002;89:195506.
L. Li et al. / Acta Materialia 61 (2013) 3347–3359 3359

[23] Picard G, Ajdari A, Lequeux F, Bocquet L. Phys Rev E [35] Johnson W, Samwer K. Phys Rev Lett 2005;95:195501.
2005;71:010501. [36] Argon AS, Kuo HY. J. Non-Cryst. Solids 1980;37:241.
[24] Martens K, Bocquet L, Barrat J-L. Phys Rev Lett 2011;106:156001. [37] Williams ML, Landel RF, Ferry JD. J Chem Phys 1955;77:3701.
[25] Homer ER, Rodney D, Schuh CA. Phys Rev B 2010;81:064204. [38] Doolittle AK. J App Phys 1951;22:1471.
[26] Homer ER, Schuh CA. Acta Mater 2009;57:2823. [39] Manning ML, Langer JS, Carlson JM. Phys Rev E 2007;76:056106.
[27] Homer RE, Christopher AS. Model Simul Mater SC 2010;18:065009. [40] Moorcroft RL, Cates ME, Fielding SM. Phys Rev Lett
[28] Bulatov VV, Argon AS. Model Simul Mater SC 1994;2:167. 2011;106:055502.
[29] Bulatov VV, Argon AS. Model Simul Mater SC 1994;2:185. [41] Zhao P, Li J, Wang Y. Int J Plast 2013;40:11.
[30] Bulatov VV, Argon AS. Model Simul Mater SC 1994;2:203. [42] Tsamados M, Tanguy A, Goldenberg C, Barrat J-L. Phys Rev E
[31] Jagla E. Phys Rev E 2007;76:046119. 2009;80:026112.
[32] Vandembroucq D, Roux S. Phys Rev B 2011;84:134210. [43] Langer JS. Scripta Mater 2006;54:375.
[33] Eshelby JD. P Roy Irish Acad A 1957;241:376. [44] Bouchbinder E, Langer JS. Phys Rev Lett 2011;106:148301.
[34] Argon AS, Shi LT. Acta Metall Mater 1983;31:499. [45] Bouchbinder E, Langer JS. Phys Rev E 2011;83:061503.

You might also like