Você está na página 1de 10

Grain boundaries exhibit the dynamics

of glass-forming liquids
Hao Zhang
a
, David J. Srolovitz
b
, Jack F. Douglas
c,1
, and James A. Warren
d
a
Department of Chemical and Materials Engineering, University of Alberta, Edmonton, AB, Canada T6G 2V4;
b
Department of Physics, Yeshiva College,
Yeshiva University, New York, NY 10033; and
c
Polymers Division and
d
Metallurgy Division, National Institute of Standards and Technology,
Gaithersburg, MD 20899
Edited by Johanna M. H. Levelt Sengers, National Institute of Standards and Technology, Gaithersburg, MD, and approved March 24, 2009
(received for review January 14, 2009)
Polycrystalline materials are composites of crystalline particles or
grains separated by thin amorphous grain boundaries (GBs).
Although GBs have been exhaustively investigated at low tem-
peratures, at which these regions are relatively ordered, much less
is known about them at higher temperatures, where they exhibit
signicant mobility and structural disorder and characterization
methods are limited. The time and spatial scales accessible to
molecular dynamics (MD) simulation are appropriate for investi-
gating the dynamical and structural properties of GBs at elevated
temperatures, and we exploit MD to explore basic aspects of GB
dynamics as a function of temperature. It has long been hypoth-
esized that GBs have features in common with glass-forming
liquids based on the processing characteristics of polycrystalline
materials. We nd remarkable support for this suggestion, as
evidenced by string-like collective atomic motion and transient
caging of atomic motion, and a non-Arrhenius GB mobility describ-
ing the average rate of large-scale GB displacement.
glass formation grain-boundary mobility molecular dynamics
polycrystalline materials string-like collective motion
M
ost technologically important materials are polycrystalline
in nature (1), and it is appreciated that the grain bound-
aries (GBs) of these materials, the interfacial region separating
the crystal grains (see Fig. 1A), significantly influence the
properties of this broad class of materials (2). In particular, the
dynamical properties of GBs, such as the GB mobility (M), play
an important role in the plastic deformation and evolution of
microstructure during material processing and service (3).

The atomic organization in the GBs represents a compromise


between the ordering effects of adjacent grains, and packing
frustration [or reduced packing efficiency (6)] is also charac-
teristic of glass-forming (GF) fluids, in which particle ordering
is likewise limited in range (7). This simple observation leads us
to expect similarities between the dynamics of GBs and GF
fluids, and below we provide evidence for this relationship. By
implication, GB migration should then be sensitive to impurities,
geometrical confinement, and applied stressesbasically any
factor that affects particle-packing efficiency (810). To illus-
trate this point and test our perspective of GB dynamics, we
quantitatively interpret differences in the effect of large tensile
and compressive deformations on M in terms of measures of
cooperative atomic motion drawn from the physics of GF fluids.
Nearly 100 years ago, Rosenhain and Ewen (11) suggested that
metal grains in cast iron were cemented together by a thin
layer of amorphous (i.e., noncrystalline) material identical
with or at least closely analogous to the condition of a greatly
undercooled liquid. Although this conceptual model was able
to rationalize processing characteristics of ferritic materials (11),
it was not possible to validate it at the time through direct
observation or simulation. Sixty years later, Ashby (12) simu-
lated GB dynamics by using a model system of layers of
macroscopic bubbles (bubble rafts); he also suggested simi-
larities between GBs and GF liquids. Again, the inability to test
this hypothesis on physically realistic GBs limited its acceptance.
Recent molecular dynamics (MD) simulations of polycrystalline
metals and silicon by Wolf and coworkers (13, 14) suggest that
the disordered structure within GBs can be well described by
atomic radial distribution functions characteristic of GF liquids.
Here we directly address what, if any, characteristics GBs share
with GF liquids by performing a series of MD simulations on the
migration of model GBs at elevated temperatures (T) and by
comparing the GB MD and GB mobility M(T) to the MD and
large-scale structural relaxation of GF liquids. In Section II we
describe GB geometry and our GB model, quantify cooperative
atomic motion within the GBs by using measures applied before to
GF liquids, compare the T dependence of M(T) to the transport
properties (e.g., viscosity) of GF liquids, and then determine the
characteristic temperatures of GB fluids that define the broad
transition in the GB mobility and which have their counterparts in
GF liquids. Further comparison of these transitions is included in
SI to more precisely identify the type of GF liquids that most
resemble our simulated GB fluid dynamics. We also illustrate the
value of this perspective of polycrystalline materials by applying it
to explain changes in M(T) arising from the form of applied stress
(e.g., tensile versus compressive deformation) to which the poly-
crystalline material is subjected.
Results
I. Cooperative GB Atomic Motion and GB Mobility. GB geometry and
model. The crystallography of GBcan be minimally specified by five
variables: three parameters to specify the relative orientation of one
grain with respect to the other and two to indicate the GB
inclination (3) (note the distinct orientations of grains indicated in
Fig. 1B; a misorientation angle describes a rotation of one grain
that will cause it tocoincide inorientationwiththe other). Although
these variables characterize the basic GB crystallography, they do
not fully specify GB atomic structure. GB structure has been
variously described in terms of dislocations (15), structural units
(16), coincidence site lattices (CSLs) (3), etc. All of these
descriptions of GB structure have their limitations at elevated T, at
which the GBs exhibit appreciable disorder.
Because the geometrical parameters of the CSL provide a
prevalent method for GB characterization, we describe aspects
of this classification scheme required for our discussion below.
For certain misorientation angles between adjacent crystals, a
Author contributions: D.J.S., J.F.D., andJ.A.W. designedresearch; H.Z. andJ.F.D. performed
research; H.Z. and J.F.D. analyzed data; H.Z. performed molecular dynamics simulations;
and J.F.D. wrote the paper.
The authors declare no conict of interest.
This article is a PNAS Direct Submission.
1
To whom correspondence should be addressed. E-mail: jack.douglas@nist.gov

In inorganic crystalline materials, the GBs are often as wide as several nanometers (in
metals the GB widths are more typically in the 0.2- to 1-nm range). When the grain size is
on the nanoscale, the fraction of the material in these GB regions becomes appreciable [in
excess of 10%for polycrystalline materials with a grain size of 10 nm(4)], and the GBs can
then come to dominate material properties (5).
This article contains supporting information online at www.pnas.org/cgi/content/full/
0900227106/DCSupplemental.
www.pnas.orgcgidoi10.1073pnas.0900227106 PNAS May 12, 2009 vol. 106 no. 19 77357740
P
H
Y
S
I
C
S
common superlattice can be defined if the adjacent crystal
lattices are formally allowed to interpenetrate each other. This
lattice is called a CSL (3), and the corresponding GB is char-
acterized by a coincidence number , the ratio of the volumes
of the CSL and individual grain lattice cells. For most misori-
entation angles, there is no CSL, but there is a countable infinity
of misorientation angles for which the CSL exists (just a few of
these being important in applications), and these GBs are
termed special or type. On the other hand, in the common
case where the GBs are more disordered so that the CSLs are not
defined, the GBs are termed general or non- type. Low
value GBs have relatively lower GB energies (3), and materials
having a high density of GBs with this symmetric structure give
rise to enhanced corrosion resistance and other desirable ma-
terial properties. However, most materials involve random
GBs having a broad population of and non- GBs that can be
altered by changing the annealing history. For generality, we thus
investigate both of these basic GB types. Visualizations of and
non-GBs, which are helpful in developing an intuition for these
structures, are given in ref. 17.
Fig. 1B shows a representative non- GB having a misorien-
tation angle of 40.23 and the coordinate systemdefining this
structure and to which we refer below. The system actually
simulated comprises only two crystal grains and the interfacial
region between them, a bicrystal. The temperature T/T
m
range
considered ranges from 0.48 to 0.86, where the reduced T is
defined in terms of the bulk Ni melting temperature, T
m
1,624
K (18) and where T is controlled with a HooverHolian ther-
mostat (19). An embedded-atom potential (20) describes inter-
actions between the 22,630 Ni atoms in our simulation.
Compressive stresses were applied to the bicrystal (constant
stress arises from the application of constant strain along the x
and y directions, as illustrated in Fig. 1B, where x, y, and z axes
define laboratory coordinates and [100], [010], and [001] define
crystallographic axes), giving rise to a change in the elastic free
energy of each grain (the crystals are elastically anisotropic) and
a thermodynamic driving force for GB displacement. Compu-
tation details were described in our previous articles (21, 22).
Cooperative particle motion within the GB. Cooperative particle dy-
namics is one of the most characteristic features of the dynamics
of GF fluids. In particular, both atomistic simulations and
experiments on colloidal and granular fluids demonstrate that
this cooperative motion takes the form of string-like motion
(2326). To examine whether a similar dynamics occurs in the
GB regions of polycrystalline materials, we apply methods
originally developed to identify this type of motion in GF liquids
(23) to our MD simulations of GB atomic dynamics. As a first
step in identifying the collective particle motion, we identify the
mobile atoms in our system (i.e., those that move a distance
in a time t that is larger than the typical amplitude of an atomic
vibration but smaller than the second nearest-neighbor atomic
distance). Because there are no defects in either crystal, the
average displacement of the atoms within the grains after any t
is within the scale of the mean vibrational amplitude. Therefore,
the identified mobile atoms within t are all located within GB
region. Next, mobile atoms i and j are considered to be within a
displacement string if they remain near one another as they move
(9, 18, 23) (see Fig. 1C). We indeed find strings in our GB
dynamics simulations as in GF liquids and provide some char-
acterization of these structures below to determine how their
geometry compares to their counterparts in GF fluids. In
particular, the average string length,
n t
n2

nPn, t . [1]
Fig. 1. Illustration of string-like cooperative atomic motion within a GB. (A) Schematic microstructure of polycrystalline metal. Different colors indicate the
individual grains having different orientations, and the black line segments represent GBs. (B) Equilibrium boundary structure projected onto the xz plane for
a 40.23 [010] general tilt boundary at T 900 K (x, y, and z axes are lab-xed Cartesian coordinates, whereas [100], [010], and [001] refer to crystallographic
axes). Upper and lower grains rotate relatively to each other by 40.23 along the common tilt axis [010]. The misorientation angle 40.23 does not correspond
to a special value [ refers to the ratio of the volume of the coincidence site lattice (CSL) to the volume of crystal lattice]. The atoms are colored according to
their coordination numbers q (orange, q 12; others, q 12). The simulation cell was chosen to have the GB plane normal to the z axis. (C) Representative string
within GB plane. Yellow and blue spheres represent the atoms at an initial time t 0 and a later time, t*. (D) Snapshot of string-like cooperative motion within
the GB region at T 900 K at t t*. The rectangular box illustrates the simulation cell in the xy plane. Biaxial strain
xx
and
yy
are applied to xy plane to
induce driving force that arises from the elastic energy difference between two grains, as shown in the diagram above the box.
7736 www.pnas.orgcgidoi10.1073pnas.0900227106 Zhang et al.
provides a natural measure of the scale of cooperative particle
motion in strongly interacting liquids where P(n,t) is the
probability of finding a string of length n a time interval t.
Previous work (8, 22) has established that the average length of
these strings in GF liquids grows after cooling, along with the
effective activation energy for structural relaxation. This finding
accords with the AdamsGibbs theory of relaxation in GF liquids
(27), in which the strings are identified (23) with the vaguely
defined cooperatively rearranging regions of the Adams
Gibbs theory. Strings are thus of practical interest, because they
are correlated with the relative strength of the temperature
dependence of transport properties (see below), perhaps the
most important property of GF fluids.
Similarity of GB mobility to transport properties of GF liquids. The GB
mobility is defined by the rate of displacement of the GB after
the application of a stress to the polycrystalline material so this
motion occurs in a direction orthogonal to the plane of the GB
in which collective atomic motions primarily occur. Fig. 2 shows
the displacement of the GB position as a function of t for eight
representative T values. Because the structure of GBs is more
disordered than that of the crystal grains, the coordination
number for most atoms within GB region is normally reduced.
On the basis of this observation, the mean GB position can be
calculated by averaging the positions of those atoms having this
reduced coordination number (see Fig. 1B legend). The bound-
ary velocity v was simply obtained from the average slope of the
boundary position versus t, where final displacements from 1 to
5 nm were considered, depending on T.
In the classical theory of GB migration (28), the temperature
dependence of the GBmobility (ratio of the velocity v to the driving
force p) obeys an Arrhenius temperature dependence {i.e., M
M
0
exp[Q/(k
B
T)]}, where M
0
is a constant and Q is the activation
energy for boundary migration. Fig. 2 Inset (opencircles) shows that
this expectation is not satisfied. Instead, if we fit the GB mobility
data to the VogelFulcher (VF) equation (29),
M M
VF
exp
Q
VF
k
B
T T
0

, [2]
where M
VF
, Q
VF
, and T
0
are material-specific constants, then the
fit becomes significantly better. A strong temperature depen-
dence of large-scale transport properties is a characteristic, even
defining property, of GF liquids, and the VF equation (Eq. 2)
phenomenologically describes structural relaxation and diffusion
at low temperatures in an astounding number of GF materials at
temperatures higher than the glass transition temperature, T
g
.
Eq. 2 does not apply below T
g
, at which an Arrhenius temper-
ature dependence of relaxation is again normally recovered.
Because T
g
is generally well above the temperature T
0
at which
the rate of molecular diffusion formally vanishes in the VF
mobility relation (see Eq. 2), the extrapolation temperature (30)
T
0
cannot be literally identified with a condition of vanishing
mobility. The same situation is true for the GB mobility, M(T),
so that T
0
only serves to characterize the strength of the
non-Arrhenius temperature dependence of the mobility over a
restricted temperature range above T
g
.
A best fit of the data in Fig. 2 to Eq. 2 yields the characteristic
temperature, T
0
(509 18) K (error estimate is based on a 95%
confidence interval throughout this article), and belowwe compare
the ratio, D ' Q
VF
/k
B
T
0
, the fragility parameter in GF liquids,
for our GB dynamics to D for appropriate classes of GF fluids. We
conclude that the temperature dependence of the GB mobility
obeys the same phenomenological relationship as relaxation in GF
liquids, providing some support for the physical picture of poly-
crystalline materials described in our introduction.
We tested the physical sensibility of T
0
in relation to GF fluids by
comparing this quantity to the melting temperature T
m
. For me-
tallic glasses, the ratio of the melting (eutectic) temperature to T
0
has been estimated to be T
m
/T
0
2.8 (SI Text and Table S1), which
is reasonably close to the corresponding ratio 1624/509 K3.2 that
we found for GB migration. We conclude that the magnitude of T
0
that we estimated is quite reasonable in comparison to the phe-
nomenology of GF fluids. Next, we will consider the commonality
between the MDof GBmotion and GFliquids at a molecular scale.
Cooperative molecular motion in GB and GF liquids. As discussed above,
string-like cooperative atomic motion is prevalent in all GF
liquids examined to date [including water, polymer fluids, me-
tallic GF liquids, concentrated colloidal suspensions, and even
strongly driven granular fluids (8, 23, 26, 31, 32)]. It is apparently
a universal property of the dynamics of strongly interacting
fluids, where a strong reduction in the particle mobility and an
enormous change in the rate of structural relaxation are found
in association with the growth of string-like correlated motion
after approaching the glass transition. We next examine the
nature of the atomic motion occurring in GB migration to
determine whether it follows this general pattern of frustrated
fluid dynamics. Fig. 1C shows a typical displacement string in
our simulation appearing in the plane of the GB region. The
initial atom positions are shown in yellow (t 0), and their
positions at a later time t are shown in blue (displacements are
shown by using arrows; see Fig. 2 C and D). This string-like
atomic motion in the GB region occurs predominantly along a
direction parallel to the tilt axis (see Fig. 1D) but in a direction
orthogonal to the ultimate direction of GB displacement.
Atomistic simulations of GF liquids suggest that the distribu-
tion of string lengths P(n) is an approximately exponential
function of n,
Pn exp n/ n. [3]
Fig. 3 shows the distribution of string lengths at t t*, where
the string length n(t) exhibits a maximum during GB migration
[time dependence of n(t) not shown; see refs 2224]. Interest-
ingly, the distributions of n in GF liquids and in the GBs are
essentially the same (10), and even the magnitude of n is
comparable to values found in GF liquids for the corresponding
T range (see below). Evidently, n increases after cooling (see
Fig. 3 Inset), where n is smaller for GBs than the non- GBs,
Fig. 2. Temporal evolution of the average boundary position at eight
different temperatures. (Inset) Logarithm of the boundary mobility as a
function of T [open circles versus 1/T (top axis) compare the data to an
Arrhenius relationship, whereas the lled circles compare to the VF equation
(bottom axis)]. The nonlinearity of the Arrhenius plot indicates that this
relationship does not apply to GB mobility data.
Zhang et al. PNAS May 12, 2009 vol. 106 no. 19 7737
P
H
Y
S
I
C
S
which suggests that atoms in the GBs are less frustrated than
those in the more disordered non- GBs.
Comparison of the characteristic temperatures of GB and GF liquids. We
next consider other aspects of the phenomenology of GF liquids
that have significance for understanding the transport properties of
polycrystalline materials. In particular, glass formation is generally
accompanied by thermodynamic changes similar to those observed
in rounded thermodynamic transitions, and correspondingly, both
of these transitions are characterized by multiple characteristic
temperatures (31). We summarize these for GF fluids, and below
we determine their analogs for GB dynamics.
The dynamics of GF liquids is characterized by a number of
temperatures, which in decreasing order include T
A
(demarking
the onset of the cooperative atomic motion), a crossover tem-
perature T
c
(separating high and low T regimes of glass forma-
tion), the glass transition temperature T
g
(below which aging and
other nonequilibrium behavior is overtly exhibited), and finally
T
0
(characterizing the end of the glass-transformation process)
(31, 33). To further our analysis, it is natural to consider the
analogous characteristic temperatures for the GB MD. T
A
is
experimentally defined as the temperature at which the Arrhe-
nius temperature dependence of structural relaxation no longer
holds. On the basis of the entropy theory of glass formation (33),
the apparent activation energy E
a
(T) below T
A
follows a uni-
versal quadratic T dependence, E
a
(T)/E
a
(T 3 T
A
) 1
C
0
(T/T
A
1)
2
, where C
0
is a constant. The GB migration
mobility data in Fig. 2 Inset fits this relation (SI) well near T
A
(950 K T 1,400 K), and we estimate T
A
to equal 1,546 K. To
determine the other characteristic temperatures (see SI), we
performed a series of simulations to determine the mean-square
displacement of atoms r
2
within the GBs.
Following Starr et al. (30), we define the DebyeWaller factor
(DWF) as the mean-square atomic displacement r
2
after a fixed
decorrelation time t
0
characterizing the crossover from ballistic to
caged atom motion. Fig. 4 shows corresponding data for the GB
dynamics where the same criterion is chosen for defining t
0
as
described by Starr et al. (30). In GF liquids, r
2
at low T exhibits a
well-defined plateau after t
0
that persists up to the structural
relaxation time of the fluid (a time normally many orders of
magnitude larger than t
0
for T T
g
), and the height of this plateau
defines the size of the cage in which particles are transiently
localized by their neighbors. In Fig. 4, we observe a progressively
flattening of the GB data for r
2
at intermediate times that
illustrates the progressive caging of atomic motion after cooling in
these relatively high-temperature simulations. A well-defined cage
(plateau) is apparent only for the lowest temperature simulations
indicated. Because r
2
is slowly varying with t in the time interval
between t
0
and in the glass state and because is on the order of
minutes for TT
g
, any experimental estimate of r
2
yields a similar
value in the glass state over this broad range of time scales. This
observation basically explains why dynamic neutron and x-ray
scattering measurements of r
2
with instrumental time scales on the
order of 10
9
s are of direct relevance to the physics of glasses. Fig.
4 shows the DWF u
2
(T) for the GB atomic motion based on the
same criterion (i.e., r
2
(t
0
) u
2
) used by Starr et al. (30) for GF
fluids. For reasons described below, u
2
(T) is divided by
2
(T), the
square of the equilibrium interatomic distance in the crystal. Fig. 4
Inset shows the original GB data from which u
2
was determined
(the broken line denoting t
0
)
.
In addition to the mean particle-displacement scale u
2
(T) in
the caging regime, the distribution function G
s
(r,t), for the
atomic displacement r(t), also provides valuable information
about how much GB atomic motion resembles the dynamics of
GF liquids, and this property also allows for the determination
of the characteristic temperature T
c
(34, 35). In particular, we
calculate the GB self-intermediate scattering function F
s
(q,t)
(35), which is the Fourier transform of G
s
(r,t), over a broad
temperature range, 1,050 K T 1,150 K. Specifically, F
s
(q,t)
is defined as F
s
(q,t) exp{iq[r
i
(t) r
i
(0)]}, where the
Fig. 3. String-length distribution function P(n) for 36.9 boundary (GB)
and 40.2 boundary (non- GB) at 800 and 1,400 K. (Inset) Average string
length n as a function of T for the GB, illustrating the growth of the scale of
collective atomic motion after cooling. The scale of n at a corresponding
reduced temperature is similar to that of previous simulation observations on
GF liquids (9, 23).
Fig. 4. DWF u
2
as a function of T for the GB and crystal. The solid red circles
represent the DWF for GB, and the solid pink circles indicate the DWF for the
lower grain. We observe that u
2
for the GB atomic motion (upper curve) is
substantially larger than the scale of atomic motion in the grain (lower curve),
with u
2
in the grain only obtaining a value comparable to the Lindemann
value for temperatures near T
m
1,621 K (18). The VF temperature T
0
,
determined from the M(T) data in Eq. 2, nearly coincides with the T at which
u
2
extrapolates to 0. The temperature T
A
1,546 K approximates the onset
of thesupercoolingregimeandwas alsodeterminedfromtheM(T) datainFig.
2 (see SI). The crossover temperature T
c
is determined from the GB uid
structural relaxation time (see Fig. 5), and T
g
is estimated by the condition
u
2
/
2
0.125, where is the interatomic distance inthe crystal, a Lindemann
condition for glass formation (33). For the crystal grain, u
2
shows linear
dependence over the entire T range. However, for GB, u
2
varies nearly
linearly with T (harmonic localization) at lowtemperatures, but for T T
c
the
harmonic approximation no longer applies. (Inset) r
2
(t) data fromwhich u
2

' r
2
(t
0
) was determined. The vertical broken line indicates the inertial
decorrelation time, t
0
(30).
7738 www.pnas.orgcgidoi10.1073pnas.0900227106 Zhang et al.
Fourier transform variable q is termed the scattering wavevec-
tor because of its measurement interpretation. Because the
variation of F
s
(q,t) with T is an apparently universal property of
GF liquids, an examination of F
s
(q,t) for the GB dynamics
provides an opportunity to examine how much the GB atomic
motion resembles the MD of GF liquids.
In Fig. 5, we see that F
s
(q,t) for the GB dynamics develops a
progressive tendency to flatten out at time scales after the
inertial dynamics regime, as in the GB data for r
2
(t) shown in
Fig. 4. In each case, this behavior arises from particle caging or
transient particle localization after cooling. In GF fluids, F
s
(q,t)
is an isotropic function of the magnitude of the scattering
wavevector (i.e., q q) and F
s
(q,t) characteristically exhibits
a stretched exponential variation, F
s
(q,t) exp[(t/)

], at
long time scales t, where the amount of stretching in this
phenomenological relation is quantified by the extent to which
is 1. Despite the application of stress in the GB system, F
s
q,t)
of the GB dynamics is isotropic to within numerical uncertainty
(18, 22); moreover, we see from Fig. 5 that a stretched expo-
nential fits the long time decay of the GB data for F
s
(q,t) quite
well. In particular, we find that the fitted values of are nearly
constant ( 0.34) for the lower T data where the secondary
decay of F
s
(q,t) is well developed. We also find that is well
described (Fig. 5 Inset) by an apparent power law, (T T
c
)

,
in the restricted T range indicated (see below). Both of these
observations on GB dynamics are characteristic for GF liquids
(34), where the fitted temperature T
c
is conventionally termed
the mode-coupling temperature. In particular, T
c
and from
our GB simulations are estimated to equal T
c
923 23 K and
2.58 0.4, a value of that is typical for GF liquids (34).
Note that the apparent power law T scaling of in GF liquids
is limited to a T range between T
A
and T
c
, and our GB data are
correspondingly restricted to such an intermediate T range.
Mode-coupling theory is an idealized mean field theory of the
dynamics of supercooled liquids, the power-law divergence in
predicted by this theory does not actually occur in practice, and
this theory does not accurately predict T
c
. Instead, T
c
has the
physical significance of prescribing a crossover temperature
separating well-defined high- and low-temperature regimes of
glass formation (33).
To complete our comparison of the characteristic tempera-
tures of GF liquids with those of the GBs, we must also
determine the low-T glass regime temperatures, T
g
and T
0
.
Equilibrium simulations of liquids are normally limited to T
T
c
because of the growing relaxation and equilibration times of
cooled liquids. The same difficulty holds for studying GB
motion. Thus, T
g
must be obtained through extrapolation of
high-T simulation data. With this difficulty in mind, we observe
that u
2
in Fig. 4 exhibits a linear T dependence up to 940 K,
a T near T
c
. Temperatures above T
c
define a different regime of
behavior for u
2
and high-T regime of glass formation more
broadly (33). The u
2
data in Fig. 4 extrapolate to 0 at a T close
to the VF temperature (T
0
) (see Eq. 2) determined from our GB
mobility data in Fig. 2. This finding accords with previous
simulation observations by Starr et al. (30) on GF fluids where
the temperature at which u
2
extrapolates to zero coincides
within simulation uncertainty with the VF temperature charac-
terizing the T dependence of the structural relaxation time at
low T, where was determined, as described above, from F
s
(q,t).
Thus, we find another striking correspondence between GB
atomic dynamics and the dynamics of GF liquids.
Physically, T
g
corresponds to a condition in which particles
become localized in space at essentially random positions through
their strong interaction with surrounding particles, and an (arguably
nonequilibrium) amorphous solid state having a finite shear mod-
ulus emerges under this condition. A particle localization
delocalization also underlies crystallization, and the Lindemann
relation is known to provide a good rough criterion for the melting
transition. Correspondingly, the same instability condition (33, 36)
has been advocated generally for GF liquids, which allows a direct
estimation of T
g
. Following Dudowicz et al. (33), we define T
g
by
the Lindemann condition, u
2

1
2
/ 0.125, and we estimate T
g

695 K, which is a typical magnitude for metallic GF liquids (37). In
practice, experimental estimates of T
g
depend somewhat on the rate
of sample cooling and other sample history effects, which leads to
some uncertainty in this characteristic temperature. The Linde-
mann criterion for T
g
is just a rough criterion for a roughly defined
quantity. We have now defined all of the analogs of the character-
istic temperatures of GF fluids for our simulations of GBdynamics;
next, we compare to the relationships between these temperatures
for both GB and GF liquids.
II. Perspective on Nature of Applied Stress and Impurities on GB
Dynamics. Our paradigm for GB dynamics emphasizes the im-
portance of string-like cooperative motion in understanding the
transport properties of both polycrystalline materials and GF
liquids, and we now apply this perspective to investigate a
formerly puzzling phenomenon relating to GB migration under
large deformation conditions. Various types of perturbations
(e.g., hydrostatic pressure, molecular and nanoparticle additives,
nanoconfinement) can be expected to influence molecular pack-
ing and, thus, the extent of packing frustration in the fluid and
should influence the collective string dynamics. In previous
work, we found that a variation of T and the GB type both
influenced (18) the average string length n, so a sensitivity of
n to thermodynamic conditions is established. We can also
expect that varying the type of loading conditions, such as
applying compressive stress, tensile stress, or even a constant
hydrostatic pressure, will influence the character of string for-
mation, because these forms of applied stress naturally affect
molecular packing differently. These resulting changes in string
geometry should then be directly reflected in M, providing an
interesting test of the formal relationship between the dynamics
of GB and GF fluids.
In previous simulations, we showed that large compressive
versus tensile stresses led to appreciable changes in the T
dependence of M (21), where M at 800 K differed for these
modes of stress by a factor of order 10, even when driving forces
Fig. 5. The self-intermediate scattering function for GB particles in the T
range of 1,050 and 1,150 K (dened in the text). The dashed curves are a t of
the stretchedexponential relation, F
s
(q,t) exp[(t/)

] tothe long-time data,


where the short-time decay arises fromthe inertial atomic dynamics. (Inset) A
power t of toTT
c
, whereT
c
andareadjustableparameters as inprevious
measurements and simulations.
Zhang et al. PNAS May 12, 2009 vol. 106 no. 19 7739
P
H
Y
S
I
C
S
had the same magnitude. This sensitivity of M to the mode of
applied stress is difficult to explain in terms of conventional GB
migration theories, but this effect is readily understood from our
perspective of GB dynamics. In particular, a reexamination of
our former simulation results indicates that n for 2% tensile
and compressive strains at T 800 K equals 1.63 and 2.13,
respectively (SI). If one assumes that the apparent activation
energy Q for GB migration can be scaled by n [i.e., Q n E
0
,
where E
0
is the high temperature activation energy (near T
A
)],
then the observed change of n (and thus Q) in tension and
compression accounts for the change in magnitude of M. We
suggest that the main origin of this shift in the scale of collective
motion n derives from a shift of T
g
with deformation, com-
pressive deformation acting similarly to an increase in the
hydrostatic pressure, which generally increases T
g
, whereas the
extensional deformation has an opposite effect. Temperature
and pressure studies will be required to confirm this interpre-
tation of the origin of the deformation-induced changes in M in
terms of the influence of the mode of deformation on cooper-
ative GB atomic motion.
The addition of impurities and nanoscale confinement can also
be expectedtoaffect the cooperativity of atomic motions instrained
polycrystalline materials, as recently shown in simulations of GF
liquids (8, 9). Specifically, if the impurities help relieve packing
frustration, then n should be greatly attenuated (8) and the T
dependence of M should be weakened (i.e., the glass formation
becomes stronger), whereas if the impurities disrupt molecular
packing, then the scale of collective motion should become ampli-
fied and the Tdependence of Mshould be amplified. Large changes
in M, and the resulting properties of polycrystalline materials, are
then expected from the application of strains and the presence of
impurities through the influence of these effects on the scale of
collective motion in the GB region.
Discussion
We conclude that the atomic dynamics within the GB region of
polycrystalline materials and the GB mobility at elevated T
exhibit many features in common with GF liquids. Highly
cooperative string-like atom motion in the plane of the GB can
greatly affect the average rate of GBmotion transverse to the GB
plane. This understanding of GB dynamics is expected to shed
significant light on the mechanical properties of polycrystals.
Indeed, we expect the viscoelastic and highly temperature-
dependent properties of the complex GB fluid enveloping the
crystalline grains to have a large impact on the plastic defor-
mation of these materials. This viewpoint of the polycrystalline
materials is contrasted with recent work that attributes the
deformational properties of polycrystalline materials to simply
the presence of solid crystalline grains within the uncrystallized
fluid melt (38, 39). In our view, the uncrystallized material is a
complex fluid that imparts its own viscoelastic effects on the
polycrystalline material. Moreover, the frustrated atoms within
the GB region should exhibit a high sensitivity to impurities,
pressure, and geometrical confinement as in the case of GF
liquids so that we can anticipate significant changes in the plastic
deformation properties of polycrystalline materials arising from
a modulation of the collective motion in the GB regions through
these perturbations. This perspective of polycrystalline materials
offers the promise of an increased control of the properties of
semicrystalline materials based on further quantification of this
phenomenon. Although this conceptual view of polycrystalline
materials was intuitively recognized by scientists and engineers
involved in the fabrication of iron materials at the beginning of
the last century (11), the present work puts this working model
of the deformation properties of polycrystalline materials on a
sound foundation through direct MD simulation.
ACKNOWLEDGMENTS. We thank Robert Riggleman of the University of
Wisconsin and Anneke Levelt Sengers of the National Institute of Standards
and Technology for helpful comments and questions about the work, and we
acknowledge the support of the U.S. Department of Energy (DE-FG02-
99ER45797) and the National Institute of Standards and Technology. Useful
group discussions were facilitated by the Department Of Energy Ofce of
Basic Energy Sciences Computational Materials Science Network program.
1. Ahluwalia R, Lookman T, Saxena A(2003) Elastic deformation of polycrystals. Phys Rev
Lett 91:055501.
2. Sutton AP, Balluf RW (1995) Interfaces in Crystalline Materials (Clarendon, Oxford).
3. Gottstein G, Shvindlerman LS (1999) Grain Boundary Migration in Metals: Thermody-
namics, Kinetics, Applications (CRC Press, Boca Raton, FL).
4. Zhu X, Birringer R, Herr U, Gleiter H (1987) Phys Rev B Condens Matter 35:90859090
5. Van Swygenhoven H, Weertman JR (2006) Mater Today 9:2431.
6. Rouxel T (2007) Elastic properties and short- to medium-range order in glasses. J Am
Ceram Soc 90:30193039.
7. Dzugutov M, Simdyankin SI, Zetterling FHM (2002) Decoupling of diffusion from
structural relaxationandspatial heterogeneity ina supercooledsimple liquid. Phys Rev
Lett 89:195701.
8. Riggleman RA, Yoshimoto K, Douglas JF, de Pablo JJ (2006) Inuence of connement
on the fragility of antiplasticized and pure polymer lms. Phys Rev Lett 97:045502.
9. Riggleman RA, Douglas JF, de Pablo JJ (2007) Tuning polymer melt fragility with
antiplasticizer additives. J Chem Phys 126:234903.
10. Riggleman RA, Douglas JF, De Pablo JJ (2007) Characterization of the potential energy
landscape of an antiplasticized polymer. Phys Rev E Stat Nonlin Soft Matter Phys
76:011504.
11. Rosenhain W, Ewen D (1913) The intercrystalline cohesion of metals. J Inst Met
10:119149.
12. Ashby MF (1972) Boundary defects, and atomistic aspects of boundary sliding and
diffusional creep. Surf Sci 31:498542.
13. Keblinski P, Phillpot SR, Wolf D, Gleiter H(1996) Thermodynamic criterionfor the stability
of amorphous intergranular lms in covalent materials. Phys Rev Lett 77:29652968.
14. Wolf D (2001) High-temperature structure and properties of grain boundaries: long-
range vs. short-range structural effects. Curr Opin Solid State Mater Sci 5:435443.
15. Read WT, Shockley W (1950) Dislocation models of crystal grain boundaries. Phys Rev
78:275289.
16. Sutton AP, Vitek V (1983) On the structure of tilt grain-boundaries in cubic metals. 1.
Symmetrical tilt boundaries. Philos Trans R Soc London Ser A 309:168.
17. Randle V, Engler O (2000) Introduction to Texture Analysis: Macrotexture, Microtex-
ture and Orientation Mapping (Gordon & Breach, Amsterdam, The Netherlands).
18. Zhang H, Srolovitz DJ, Warren J, Douglas JF (2006) Characterization of atomic motions
governing grain boundary migration. Phys Rev B Condens Matter 74:115404.
19. Hoover WG, Holian BL (1996) Kinetic moments method for the canonical ensemble
distribution. Phys Lett A 211:253257.
20. Voter AF, Chen SP (1987) Accurate interatomic potentials for Ni, Al and Ni3Al. Mater
Res Soc Symp Proc 82:175180.
21. Zhang H, Mendelev MI, Srolovitz DJ (2004) Computer simulation of the elastically
driven migration of a at grain boundary. Acta Mater 52:25692576.
22. ZhangH, Srojovitz DJ, Douglas JF, WarrenJ (2007) Atomic motionduringthemigration
of general [001] tilt grain boundaries in Ni. Acta Mater 55:45274533.
23. Donati C, et al. (1998) Stringlike cooperative motion in a supercooled liquid. Phys Rev
Lett 80:23382341.
24. Donati C, Glotzer SC, Poole PH, Kob W, Plimpton SJ (1999) Spatial correlations of
mobility and immobility in a glass-forming LennardJones liquid. Phys Rev E Stat Phys
Plasmas Fluids Relat Interdiscip Top 60:31073119.
25. Marcus AH, SchoeldJ, Rice SA(1999) Experimental observations of non-Gaussianbehav-
ior and stringlike cooperative dynamics in concentrated quasi-two-dimensional colloidal
liquids. Phys Rev E Stat Phys Plasmas Fluids Relat Interdiscip Top 60:57255736.
26. Weeks ER, Crocker JC, Levitt AC, SchoeldA, Weitz DA(2000) Three-dimensional direct
imaging of structural relaxation near the colloidal glass transition. Science 287:627
631.
27. Adam G, Gibbs JH (1965) On temperature dependence of cooperative relaxation
properties in glass-forming liquids. J Chem Phys 43:139146.
28. Porter DA, Easterling KE (1993) Phase Transformations in Metals and Alloys (Chapman
& Hall, New York).
29. Angell CA(1995) Formationof glasses fromliquids andbiopolymers. Science267:1924
1935.
30. Starr FW, Sastry S, Douglas JF, Glotzer SC (2002) What do we learn from the local
geometry of glass-forming liquids? Phys Rev Lett 89:125501.
31. Douglas JF, Dudowicz J, Freed KF (2006) Does equilibriumpolymerization describe the
dynamic heterogeneity of glass-forming liquids? J Chem Phys 125:144907.
32. Keys AS, Abate AR, Glotzer SC, Durian DJ (2007) Measurement of growing dynamical
length scales and prediction of the jamming transition in a granular material. Nat Phys
3:260264.
33. Dudowicz J, Freed KF, Douglas JF (2007) Generalized entropy theory of polymer glass
formation. Adv Chem Phys 137:125223.
34. Kob W, Andersen HC (1994) Scaling behavior in the beta-relaxation regime of a
supercooled LennardJones mixture. Phys Rev Lett 73:13761379.
35. Hansen JP, McDonald IR (1986) Theory of Simple Liquids (Academic, New York).
36. Larini L, Ottochian A, De Michele C, Leporini D (2008) Universal scaling between
structural relaxation and vibrational dynamics in glass-forming liquids and polymers.
Nat Phys 4:4245.
37. Wang WH (2005) Elastic moduli and behaviors of metallic glasses. J Non-Cryst Solids
351:14811485.
38. Gourlay CM, Dahle AK (2007) Dilatant shear bands in solidifying metals. Nature
445:7073.
39. Martin CL, Braccini M, Suery M(2002) Rheological behavior of the mushy zone at small
strains. Mater Sci Eng A Struct Mater 325:292301.
7740 www.pnas.orgcgidoi10.1073pnas.0900227106 Zhang et al.
Supporting Information
Zhang et al. 10.1073/pnas.0900227106
SI
Comparative Fragility of GBs and GF Liquids. A precise comparison
of the dynamics of GB and GF liquids requires some discussion
of the conventional jargon for classifying different types of GF
liquids. The term fragility is basic to this classification scheme,
and facilitates the determination of the specific type of GF fluid
that most resembles GB dynamics.
An Arrhenius T dependence of structural relaxation is
normally taken as an ideal behavior in GF fluids, and the
relative deviation from this behavior is quantified by the term
fragility. Fluids in which this deviation is large are said to be
fragile, whereas those that are nearly Arrhenius are said to be
strong. Alternatively, the glass transition occurs rather
abruptly after cooling in a fragile GF fluid, and the glass
transition occurs over a broad T range for a strong GF liquid. In
conventional engineering terminology, the word short is used
instead of the abstract term fragile, reflecting the short time to
work with such materials before they solidify while cooling.
Because the glass transition occurs over a narrower relative T
range in fragile liquids, the ratios of the characteristic temper-
atures for glass formation provide a convenient and model
independent measure of fragility that can be used to classify GF
liquids (1). This scheme then allows for a direct appraisal of how
the GB dynamics relates to model GF liquids. The relatively
large temperature ratio, T
c
/T
g
1.33, for the GB dynamics that
we find corresponds to a strong GF liquid in comparison to many
polymeric glass-formers, but this ratio is larger for network-
forming glasses such as B
2
O
3
and SiO
2
, which are at the strong
end of the fragility spectrum (meaning simply that the relaxation
in these systems is more nearly Arrhenius, and presumably less
cooperative, for all T). Metallic GF liquids, thus, are glass-
formers of intermediate fragility. Polymers having a simple
monomer structure, a flexible chain backbone, and flexible side
groups can also be characterized as having an intermediate
fragility (strong by the standards of polymers alone) (1), and this
class of fluids is a natural candidate for comparison with the GB
dynamics. Table S1 summarizes the T ratios found in our GB
dynamics simulations and the corresponding analytically pre-
dicted ratios (1) for low molecular mass polymeric GF liquids
having a similar intermediate fragility (FF polymer model of ref.
1). This conclusion is also confirmed through a consideration of
the GB mobility data in Fig. 2 of the main text, which when fitted
to Eq. 2 yields the fragility parameter D Q
VF
/k
B
T
0
3.59
0.3. Theoretical estimates for D range from 5 to 3 for strong
and fragile high molecular mass polymeric GF liquids, respec-
tively (1). An intermediate fragility is again indicated for GB
dynamics.
In Table S1, we also include a comparison to recent simulation
estimates of the corresponding T ratios for a relatively strong GF
polymer liquid, rendered a strong glass-former through the
addition of a molecular additive (2). Thus, we obtain yet another
striking confirmation of at least a qualitative commonality
between the dynamics of GB and GB fluids, although the overall
breadth of the glass-formation temperature range, as quantified
by T
A
/T
0
, is somewhat larger for the GB fluid than the GF liquids
in this comparison. This correspondence applies both to the
phenomenology of macroscopic transport properties (e.g., M)
and to the dynamics on an atomic scale.
Temperature Dependence of the GB Mobility (M) and Activation
Energy E
a
(T) and the Determination of T
A
. See Fig. S1.
Effect of Applied Strain on the GB String Length. See Fig. S2.
1. Dudowicz J, Freed KF, Douglas JF (2007) Generalized entropy theory of polymer glass
formation. Adv Chem Phys 137:125223.
2. Riggleman RA, Douglas JF, de Pablo JJ (2007) Tuning polymer melt fragility with
antiplasticizer additives. J Chem Phys 126:234903.
Zhang et al. www.pnas.org/cgi/content/short/0900227106 1 of 4
Fig. S1. Reduced apparent activation energy, E
a
(T)/E
a
(T 3T
A
), for the GB mobility, M(T), as a function of reduced temperature, [(T T
A
)/T
A
]
2
. This allows us
to estimate the onset temperature for string-like cooperative motion, T
A
. See Fig. 2 in the main text for original GB mobility data.
Zhang et al. www.pnas.org/cgi/content/short/0900227106 2 of 4
Fig. S2. Effects of applied strain or pressure on the string-length distribution function for a 5 [010] tilt GB at 800 K. In this study, the GB tilt axis was chosen
to be oriented along the [010] axis, the tilt angle was taken to be 36.8, and one of the boundary planes is [001]. Under the strain conditions (i.e.,
0
0%,
2%, and 2%), the average string length n was found to equal 1.98, 1.63, and 2.13, respectively, where n is determined by the slopes of these curves,
depending on the nature of the applied strain.
Zhang et al. www.pnas.org/cgi/content/short/0900227106 3 of 4
Table S1. Characteristic temperature ratios
Model strong polymer GF liquids
Ratio GB dynamics Analytic estimates (1) MD (2)
T
c
/T
g
1.3 1.4 1.4
T
c
/T
0
1.8 1.7 1.5
T
A
/T
c
1.7 1.6 1.4
T
A
/T
g
2.2 2.2 2.0
T
A
/T
0
(3) 3.0 2.6 2.2
1. Dudowicz J, Freed KF, Douglas JF (2007) Generalized entropy theory of polymer glass formation. Adv Chem Phys 137:125223.
2. Riggleman RA, Douglas JF, de Pablo JJ (2007) Tuning polymer melt fragility with antiplasticizer additives. J Chem Phys 126:234903.
3. Anexperimental estimate of the ratioTm/T0 2.8 is givenby BuschR, LiuW, andJohnsonWL (1998) Thermodynamics andkinetics of the Mg65Cu25Y10 bulk metallic glass formingliquid.
J Appl Phys 83:41344141.
Zhang et al. www.pnas.org/cgi/content/short/0900227106 4 of 4

Você também pode gostar