You are on page 1of 10

Impact Behavior of Intrinsically Brittle

Nanoparticles: A Molecular Dynamics


Perspective
B. Daneshian and H. Assadi
(Submitted August 16, 2013; in revised form October 2, 2013)
Impact behavior of intrinsically brittle materials at the nanoscale is a topic of growing interest, for
instance, in aerosol deposition and cold spraying of ceramic materials. In this work, we examine the
behavior of single-crystalline brittle nanoparticles upon impact on a rigid substrate, within the framework
of a molecular dynamics model. The model is based on Lennard-Jones formulation, where brittleness is
brought about by using a relatively small cut-off interaction distance. Simulations have been carried out
for different values of particle size and velocity. The results show that despite the induced brittleness,
particles start to deform without breaking into fragments, as the particle size falls below a critical value.
They also indicate that the deformation of particles can be accompanied by poly-crystallization and
bonding to the substrate. The necessary conditions for deformation and bonding are also predicted in
view of an analytical model of impact and fracture, considering the effects of particle size and surface
energy. The results are summarized into a parameter selection map, providing an overview of the
conditions for successful deposition of intrinsically brittle materials, in terms of particle size and velocity.
The predictions are interpreted with respect to the results of the relevant kinetic spraying studies.
Keywords aerosol deposition, cold spray, molecular dynam-
ics, nanoparticles, particle bonding
1. Introduction
Ceramic materials are often viewed as being com-
pletely brittlei.e., lack the capacity to deform plasti-
callyat room temperature. Recent studies on solid-state
coating processes, namely, aerosol deposition (AD)
(Ref 1-5) and cold spraying (CS) (Ref 6-8) appear to
contradict this view. By using AD, for instance, dense
layers of intrinsically brittle ceramic materials can be
successfully deposited at room temperature. In this
method, particles of ceramic materials with a diameter of
approximately 100-300 nm are accelerated to relatively
high velocities of 100-500 m/s before impinging a substrate
in a partially evacuated chamber (Ref 1-5). AD may be
regarded as a close variation of CS. Both processes work
on the basis of high velocity impact and bonding of solid
particles. In conventional CS, nevertheless, particles are
typically metallic, hence deformable, have a relatively
larger diameter of 5-50 lm, and are accelerated to mark-
edly higher velocities of up to 1200 m/s. Bonding and
deposition mechanisms in CS are well studied and rela-
tively well understood. A main supposition in CS is that
material deposition takes place essentially via plastic
deformation and adiabatic shear instability at the inter-
acting surfaces (Ref 9). Based on this view, material
ductility is considered as a prerequisite for bonding
mechanism in CS. This is in contrast with the observation
that ceramic materials, which exhibit no ductility at low
temperatures, can in fact be deposited via AD or CS. The
mechanism through which intrinsically brittle ceramic
particles may deform and get accommodated in a dense
and pore-free AD or CS deposit has thus remained an
open question.
An interesting aspect of AD is that it works only if
particles are smaller than a certain size; much smaller than
the typical particle size as used in CS (Ref 1-5). In case of
cold spraying of brittle materials, also, deposition is found
to be feasible only when especially tailored spray materi-
alsoften in the form of agglomerates of sub-micron
particulates (Ref 6)are used. Although the length scale
seems to play an important role in these solid-state
spraying processes, the inuence of particle size on the
deformation behavior of intrinsically brittle particles
during high-velocity impact is yet to be fully understood.
On the other hand, the effect of size on the deformation
behavior of materials, in general, is a well-established
subject. It has long been known that cracking in brittle
materials is scale-dependent (Ref 10-13) and that it may
be suppressed completely if the sample is sufciently small
(Ref 12, 13). More than three decades ago, Kendall
(Ref 13) demonstrated that crack propagation and frag-
mentation of ceramic particles under quasi-static com-
pressive loading is possible only when the particle size is
larger than a certain value. Below this critical particle size,
the material has no option but to deform plastically
without fracture. Further to Kendalls nding, it has been
B. Daneshian and H. Assadi, Department of Materials
Engineering, Tarbiat Modares University, Tehran, Iran.
Contact e-mail: ha10003@modares.ac.ir.
JTTEE5
DOI: 10.1007/s11666-013-0019-4
1059-9630/$19.00 ASM International
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
shown more recently that pillars of intrinsically brittle
materials can be deformed plastically at room temperature
when their thickness falls below a critical value (Ref 14-
16). In view of these studies, it seems reasonable to con-
ceive that a critical particle size must also exist in kinetic
spraying processes, such as AD and CS, marking the
transition between deformation and fracture.
In this context, the present study aims to provide a
computational basis for understanding the effect of size on
the impact behavior of nanoparticles of intrinsically brittle
materials. Computer models of particle impact allow
detailed analysis of various physical phenomena at very
short time and length scales, which cannot be monitored
experimentally. There are several methods that can be
used for the simulation of particle impact, deformation,
and fracture at the continuum level. These include ele-
ment vanishing technique (Ref 17), mesh splitting method
(Ref 18), node duplicating technique (Ref 19), and
material point method (Ref 20, 21). In order to simulate
the mechanical behavior of brittle materials at the nano-
scale; however, one should also take into account the
surface energy effects, which become particularly promi-
nent at nanoscale. In this respect, molecular dynamics
(MD) (Ref 22, 23) may be regarded as a suitable method
of simulation. In this method, the movements and inter-
actions of atoms or molecules are calculated based on the
Newtonian equations of motion. The method is particu-
larly suited for the study of physical and mechanical
properties of materials at small length scales (Ref 23, 24).
There are various examples of particle-based simulation of
cluster deposition in the literature (Ref 25-32). The most
relevant study, which employs MD and focuses particu-
larly on the impact behavior of brittle nanoparticles, has
been carried out by Ogawa (Ref 33-35). Ogawa explored
the inuence of various parameters, such as particle
velocity and crystallographic orientations, on the impact
behavior of nanoparticles in AD.
Here, we explore the problem further by focusing
particularly on the effect of particle size on the impact
behavior of nanoparticles. In order to illustrate general
features of a brittle-ductile transition, we make use of a
basic MD model, where brittleness is mimicked simply by
incorporation of an adjustable cut-off radius into the
Lennard-Jones potential function. Simulations of particle
impact are then carried out for various values of impact
velocity and particle size. Deformation and fracture of
particles are investigated with respect to the structural
changes of the impacted particles. The overall effect of the
particle velocity and particle size on the impact behavior
of particles is subsequently discussed in view of a criterion
for deformation without fracture, taking into account the
surface energy effects.
2. Method
A molecular dynamics model based on the Lennard-
Jones potential (Ref 23) was used to study the impact
behavior of nanoparticles in two dimensions. In order to
adjust the mechanical response of the model system, to
match that of an intrinsically brittle material, the following
potential function was considered
U
e
r
r
_ _
12

r
r
_ _
6
_ _
if
r
r
<r
c
0 if
r
r
r
c
_
Eq 1
where r is the atomic distance between two atoms, r
c
is a
dimensionless cut-off radius, and r and e are the charac-
teristic length and energy parameters, respectively. In this
way, brittle behavior can be induced and adjusted to the
desired level by choosing an appropriate value of r
c
. It
should be noted that a lower value of r
c
leads to more
prominent brittle behavior, but at the same time, it could
result in unrealistically low values of strength. Therefore,
r
c
should be carefully adjusted to represent realistic
material behavior. Quantitative modeling of brittle
behavior in real ceramic systems can be more appropri-
ately pursued using more elaborate and accurate
approaches, such as those using Ewald summation in three
dimensions. However, as will be demonstrated later, our
MD model seems to be intricate enough to capture main
features of particle impact that are relevant for the pur-
pose of the present study. The equations of motion were
integrated using the Velocity Verlet algorithm. In all MD
simulations, a microcanonical ensemble (NVE) with the
time step of 5 9 10
15
s was used. Main simulations were
performed with r =0.23 nm and e =0.96 eV. Additional
simulations were also performed with e =0.32 eV, but
these resulted in qualitatively similar results. The initial
congurations corresponded to an initial temperature of
about 360 K.
Prior to the impact simulations, the mechanical
behavior of the model system under rapid loading was
examined and adjusted via simulated tensile tests. The
objective was to mimic mechanical behavior of real cera-
mic materials, which commonly exhibit a combination of
both high strength and negligible ductility. The atoms
were initially arranged in a two dimensional close-packed
gridreminiscent of the (111) plane in fcc structureand
cut into the shape of a prenotched tensile test specimen
(Fig. 1). The specimen, oriented horizontally, had three
parts: one active zone in the middle, one xed zone
positioned at the left, and one moving zone at the right
side, traveling in the x-direction with a xed velocity of
20 m/s. The tensile-test simulations were performed with
different values of r
c
, ranging from 1.2 to 2.5, in order to
identify an appropriate cut-off radius for subsequent im-
pact simulations. As mentioned before, very small values
of r
c
would be expected to result in rupture at very low
stresses, while very large values would be expected to
result in ductile behavior. Clearly, neither extreme would
represent a real ceramic material, and so, be of interest for
the purpose of the present study. Therefore, the following
criterion was conceived for the selection of cut-off radius:
The largest r
c
resulting in the highest strength with no
plastic deformation at fracture.
Subsequently, impact simulations were carried out for
single-crystalline closed-packed nanoparticles of about
10-50 nm in diameter, d
p
, using the same cut-off radius as
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
obtained from the previous step, i.e., the value that cor-
responded to the samples showing brittle behavior and
high strength in the simulated tensile tests. In this way, the
model was tuned so that all of the nanoparticles consid-
ered for the impact simulations would be intrinsically
brittle. The nanoparticles were given initial velocities, v
p
,
of 300-750 m/s before impacting a rigid substrate at right
angle. In order to make the substrate rigid, the position of
atoms within the substrate were xed, though the corre-
sponding interatomic potentials and forces were calcu-
lated, and hence active, throughout the simulations.
Figure 1 shows the model setup and the initial congura-
tions in the corresponding simulations.
3. Results
3.1 Simulated Tensile Testing: Calibration of r
c
Figure 2 shows the results of simulated tensile tests.
According to the snapshots, Fig. 2a, there is a substantial
change in the deformation behavior of the specimen when
r
c
is reduced from 2.5 to 1.5. In the latter case, a crack
grows from the notch at an early stage, leading to rupture
with negligible ductility. As shown in Fig. 2b, the peak
strength of the material in the latter case is nevertheless
high, almost the same as in the former case with noticeable
ductility. Further decrease of r
c
resulted in a decrease of
the peak strength, as shown in Fig. 2c. According to the
criterion for the selection of r
c
(Sect. 2), therefore, the cut-
off radius of 1.5 was selected. Note that this value is the
upper limit of r
c
that can be used to simulate tensile testing
with no plastic deformation at fracture.
The signicant effect of r
c
on the deformation and
fracture behavior can be interpreted as follows. For the
system to deform plastically, attractive interatomic forces
should sustain while interatomic distances increase, during
deformation and slip of the atomic planes. For the
examined model system, this condition is fullled only
when the cut-off radius is noticeably greater than 1.5. For
smaller values of r
c
, the interatomic forces drop signi-
cantly during slip, as the interatomic distances increase,
albeit temporarily, to values comparable to the cut-off
radius. As a result, plastic deformation via slip becomes
impossible at lower values of r
c
, and the material ruptures
with no indication of slip or ductility.
3.2 Impact Simulations: Structural Features
In view of these results, the cut-off radius was xed to
1.5 throughout the subsequent impact simulations. This
allowed the model material to exhibit an intrinsically
brittle behavior with high strength, in consistence with
high-strength ceramic materials. Figure 3 shows the
structural changes during the impact of a single-crystalline
brittle particle. Here, a most prominent feature is the
formation of discontinuities or cracks, both internally and
at the surface of the particle. With the progress of impact,
some of these cracks close partially or completely, while
others propagate further and result in fragmentation.
Another interesting feature is the formation of grain
boundaries, especially near the particle/substrate contact
area.
Formation of grains and grain boundaries can also be
observed in other particles of different size and impact
velocity. An example is shown in Fig. 4. The observed
poly-crystallization appears to result from simultaneous
reorientation of small atomic clusters near the contact
area. The corresponding rotations are nevertheless small,
so that only low-angle grain boundaries formed. This is
also indicated by the calculated diffraction pattern, Fig. 4.
Dislocation formation is also evident in this example,
though this is less common a feature, as compared to grain
boundaries and cracks.
3.3 Impact Simulations: The Effect of Impact
Velocity and Particle Size
Figure 5 shows snapshots of MD simulations after
impact, for various particles and impact velocities. At
relatively high impact velocities, the particles bond
initially to the substrate (Fig. 5a). This stage may or may
not be followed by cracking and fragmentation of the
Fig. 1 The model setup and the initial congurations in the MD
simulations of tensile testing (a) and particle impact (b)
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
particle. Note that fragmentation is undesirable, as it is
associated with unsuccessful or inefcient deposition. The
resulting effect of impact at high velocities can, thus, be
classied into two categories: (i) bonding, and (ii) frag-
mentation. An important nding is that fragmentation
becomes more likely when the particle size or the impact
velocity is increased. There is also a third behavior, that is
(iii) rebounding of the particle from the substrate, as
shown in Fig. 5b. Rebounding occurs at relatively low
impact velocity, here at 300 m/s and lower, in which case
the particle remains almost intact and un-deformed as it
detaches from the substrate after impact.
4. Discussion
The results of simulations indicate that brittle behavior
can be mimicked by imposing a cut-off radius on the
Lennard-Jones potential. This behavior is manifested as
Fig. 2 The results of MD simulation of tensile testing, showing (a) snapshots of the specimens, (b) the corresponding stress-strain
curves, and (c) the strength for various values of cut-off radius. The case with r
c
=1.5 is taken to represent brittle behavior with the highest
strength
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
the lack of ductility via slip mechanism, and formation of
cracks in both tensile testing and particle impact simula-
tions. In the impact simulations, however, there is evi-
dence for deformation without fracture. This happens only
for certain particle velocities and particles sizes, and is
always associated with a structural change, reminiscent of
nanocrystallization, e.g., via dynamic recrystallization.
This feature of the MD simulations is consistent with the
Fig. 3 The results of MD simulation for a 46-nm brittle particle impacting a rigid substrate at a velocity of 425 m/s, showing the
development of various microstructural features during impact
Fig. 4 Snapshot of the atomic positions and its respective Fourier transform (calculated diffraction pattern) for a 34.5-nm brittle particle
impacting a rigid substrate at a velocity of 375 m/s. Formation of sub-grains, labeled by letters, and dislocations are evident. The grain
structure is better visible in the inset showing the inverse Fourier transform of the ltered diffraction pattern
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
existing experimental studies. For instance, TEM obser-
vations on the AD ceramic coatings have also indicated
formation of nanocrystals, which were one order of mag-
nitude smaller than the original grain size (Ref 6, 7).
Formation of nanocrystals during impact appears to be a
key to understanding the deformation mechanism of
brittle nanoparticles (Ref 32). Further to the high-velocity
impact problem, understanding these microstructural
changes could also provide an insight to the room-tem-
perature deformation mechanism of intrinsically brittle
materials at nanoscale, which is yet to be fully understood
(It should be noted that temperature rise due to particle
impact in AD, calculable via a simple energy balance, is
not nearly enough to bring about plasticity in typical
ceramic materials such as alumina. This is also indicated
by MD simulations, where the maximum homologous
temperature is predominantly below 0.5.). Moreover, such
structural changes are expected to have a signicant effect
on the physical and mechanical properties of the deposit,
and hence, are important from the viewpoint of applica-
tions for kinetic spraying processes, such as AD and CS.
Nevertheless, detailed study of deformation mechanism
and recrystallization during impact calls for more realistic
and elaborate MD models (see, e.g., Ref 33-35) than what
Fig. 5 Snapshots of particles of various diameters after impact at various velocities, showing (a) deformation with or without facture,
and (b) rebounding from the substrate with little deformation
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
is employed in the present work. The present study is,
therefore, conned to the general aspects of impact and
deposition behavior, as explained below.
The main aspect of the impact behavior as observed in
the present MD simulations is that smaller particles are
harder to break. That is, the minimum impact velocity
required to break the particle into fragments increases
with decreasing particle size (Fig. 5a). This general
observation is consistent with previous works concerning
static loading, e.g., by Kendall (Ref 13), demonstrating
that below a certain dimension fracture would not occur in
intrinsically brittle materials. It is, therefore, reasonable to
assume that a similar critical size, marking the transition
between fracture and plastic deformation, should also
exist in the case of particle impact.
In order to provide an analytical basis for the estima-
tion of the critical particle size, we consider that there
exists a minimum kinetic energy, E
k
, which would be
necessary to bring about fracture in a defect-free particle.
This kinetic energy can be conceived to consist of two
parts as follows
E
k
E
el
E
fr
Eq 2
where E
el
is the elastic energy stored in the particle, up to
the point of fracture/deformation, and E
fr
is the fracture
energy. According to this conjecture, fracture, and frag-
mentation will occur, only if the kinetic energy of the
particle upon impact is greater than the value obtained
from Eq 2. Likewise, it means that when fracture does
occur, it consumes a part of the kinetic energy that
would otherwise be used for deformation and bonding.
This is based on the presumption that fracture and frag-
mentation counteract bonding and deposition, as men-
tioned before. Each one of the energy terms in Eq 2 can
be written as a function of material and process parame-
ters as follows
E
k

1
2
mv
2
p

ptd
2
p
8
qv
2
p
Eq 3a
E
el

m
2q
Y
2
E

ptd
2
p
8
Y
2
E
Eq 3b
E
fr
stc Eq 3c
where q is the density, Y is the stress at yield/fracture, E is
the Youngs modulus, s is the crack length, t is the thick-
ness of the 2D model system (which has an arbitrary
value), and c is the interfacial energy. The crack length, s,
correlates with the particle size, though the correlation is
not straightforward. This is so because the crack surface
has a complex geometry, generally characterized as being
rough and irregular (Fig. 6). A description of the crack
geometry follows.
The fracture surfaces or the cracks in materials are
known to have fractal geometry (Ref 36). In fact,
description of the fracture geometry has been an early
application of the fractal concept, and indeed, the reason
for the selection of the term fractal (Ref 36). According to
the fractal concept, cracks are assigned a fractal dimen-
sion, which is not necessarily an integer number; it is
between 2 and 3 in a 3D space and between 1 and 2 in a
2D space. For the present 2D model of particle impact, we
assume in a rst approximation that the fractal dimension
of the cracks is 1.5. Figure 7 shows, schematically, an
example of a crack-like geometrical feature associated
with a fractal dimension of 1.5.
Assuming that the example shown in Fig. 7 is roughly
representative of the crack geometry in the present MD
simulations, the following relationship between the crack
length and the particle diameter can be conceived
s
a

d
p
a
_ _
n
Eq 4
where a is a characteristic length, relating to the length
scale of the microstructure, and n is the fractal dimension.
Combining Eq 2, 3, and 4, with n =1.5, and rearranging for
d
p
, results in the following relation between the particle
diameter and the particle impact velocity at the threshold
of fracture
Fig. 6 Snapshots of the fragments of broken particles, showing
a rough an irregular fracture surface
Fig. 7 A schematic representation of the non-linear correlation
between the crack length and the diameter of the region of
interest, corresponding to a fractal dimension of 1.5
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
d
p

1
a
qv
2
p

Y
2
E
_ _
p
8c
_ _
2
Eq 5
In view of the main energy balance, Eq 2, it can be
shown that fracture will be viable only if the particle
diameter is greater than that given by Eq 5. If the particle
diameter is smaller than this value, then the kinetic energy
is not sufcient for the creation of new crack surfaces, and
hence, the particle will either bounce back from the sur-
face or deform plastically. In principle, no other option
will be conceivable. One may also rearrange the above
relation for v
p
, and in doing so conceive a critical particle
velocity for a given particle diameter, that would equally
mark the transition between fracture and deformation.
The critical condition for fracture in fact links the particle
diameter and particle velocity; either parameter could be
taken as the relevant critical variable interchangeably.
The above analysis can be assessed with respect to the
results of MD simulations, Fig. 8. In the present study, q
and c are 8980 kg/m
3
and 2.5 J/m
2
, respectively (The sur-
face energy is obtained by examining the potential energy
of a block as it is halved into two pieces, with at surfaces on
the planes of highest atomic densitye.g., 0, 60, or 120
orientations in Fig. 1a.). The values of Y and E are worked
out with respect to the strain-stress curves as obtained from
the tensile test simulations (Fig. 2b) to be 6 and 120 GPa,
respectively. The only unknown parameter is the charac-
teristic length, a, which is taken here as an adjustable
parameter. A good agreement between the analytical
model and the simulation is obtained by taking a =0.6 nm,
which is about 2.5 times of the interatomic distance (Fig. 8).
Figure 8 also provides an overview of the window of
deposition in AD, in analogy with the parameter selection
map in CS (Ref 37). This windowis, on one hand, limited by
fracture, which occurs as the particle diameter and/or par-
ticle impact velocity exceeds the respective value given by
Eq 5. The fracture limit corresponds to the upper bound of
the kinetic energy, beyond which deposition is not viable
because the particles break into fragments. The deposition
windowis, on the other hand, limited by a minimumimpact
velocity that is needed to incur plastic deformation and
bonding. This second limit signies the lower bound of the
kinetic energy, below which deposition in not viable be-
cause the particles bounce back. There is so far no analytical
model to describe the lower bound of the kinetic energy. In
view of the MD simulations, however, one could assume
that the minimum impact velocity depends on material
properties, but it is independent of particle size; for the
present simulations it is roughly around 300 m/s for
e =0.32 eV, and 375 m/s for e =0.96 eV. This supposition
has an important consequence. As shown in Fig. 8, there
would be a crossover of these two limiting boundaries of the
deposition window at a particle diameter of about 260 nm.
This means that deposition would never be viable if the
particle size is larger than 260 nm, regardless of the mag-
nitude of the impact velocity.
The latter nding may explain why ADworks only when
particles are smaller than a certain size. Interestingly, the
critical particle size as obtained from Fig. 8 is also quanti-
tatively consistent with the typical particle size in AD(100-
300 nm). This quantitative agreement should nevertheless
be interpreted with caution, since the present MD model is
only qualitative, and the analytical model is for a 2Dsystem.
On the other hand, the agreement between the MD simu-
lations and Eq 5 may be taken as a general verication of
the proposed analytical model. It can also be shown that the
extension of the analytical model to three dimensions lead
to an almost identical expression for the critical particle size
(see Appendix). Thus, the analytical model could serve as a
basis for the prediction of the critical particle size in AD of
real materials.
The above analysis may also shed light on CS of brittle
materials. So far, CS of ceramics has been successful only
when the feedstock is in the form of agglomerates of sub-
micron particulates. Because of the bow shock that forms
in front of the substrate, it is not feasible to use nano-
particles in cold spraying. This is so because smaller par-
ticles are more effectively decelerated in the bow shock
region; and hence, sub-micron particles would effectively
be own away and not reach the substrate, under typical
cold spraying conditions. According to the above analysis,
on the other hand, larger ceramic particles would be
expected to fragment upon impact on the substrate,
resulting in negligible deposition efciency. That is per-
haps the reason why agglomeration of nanoparticles can
be used as a workaround. In this way, the bow shock effect
can be alleviated, while the size of the particulates, which
impinge the substrate at sufciently high velocities,
remains under the critical particle size.
5. Conclusions
The present molecular dynamics simulations indicated
three distinct behaviors for the impact of intrinsically
brittle nanoparticles: (i) deformation and bonding, (ii)
fracture and fragmentation, and (iii) rebounding. These
Fig. 8 An overview of the results of MD simulations, in com-
parison with the analytical model of fracture, showing the win-
dow of deposition (gray area) for intrinsically brittle
nanoparticles
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
behaviors were shown to depend on two key factors:
particle impact velocity and particle size. Increasing the
impact velocity or the particle size beyond a limit resulted
in fragmentation, while impact below a minimum velocity
resulted in rebounding. The results of simulations have
been interpreted with respect to an analytical model, in
which the kinetic energy of particles was linked to the
fracture energy. Within the framework of the present
analysis, and for the specic material model as considered
for the study, it was shown that bonding and deposition
would be possible only when the particle size is below
0.3 lm, regardless of the value of impact velocity. This
nding is generally consistent with the observations in
aerosol deposition and cold spraying of ceramic materials,
in which the particle size, or the size of particulates within
agglomerates, are generally in the sub-micron range. The
simulations also indicate poly-crystallization of the parti-
cles especially near the contact area, which is also con-
sistent with the existing experimental observations.
Acknowledgments
B.D. is thankful to Professor Michael P. Marder of
University of Texas at Austin for valuable advice on the
calculation of surface energy.
Appendix: Extension of the Analytical
Model to Three Dimensions
For a 3D system, the kinetic, the elastic, and the frac-
ture energy terms can be rewritten as follows
E
k

1
2
mv
2
p

pd
3
p
12
qv
2
p
Eq 6
E
el

m
2q
Y
2
E

pd
3
p
12
Y
2
E
Eq 7
E
fr
s
A
c Eq 8
where s
A
is the surface area of the crack. In analogy with
the example shown in Fig. 7, the crack can be assumed to
have a fractal dimension of n =2.5 in the 3D space, so that
one may conceive the following relationship between the
crack surface area and the particle diameter
s
A
a
2

d
p
a
_ _
2:5
Eq 9
This leads to the following relationship for the velocity-
dependent critical particle size in a real 3D system
d
p

1
a
qv
2
p

Y
2
E
_ _
p
12c
_ _
2
Eq 10
The above relation is almost identical to Eq 5, with the
only exception that the numerical coefcient in the right-
hand side is changed from 8 to 12. This equation shows
that the critical particle size, below which deposition is
viable, increases with increasing Youngs modulus and
surface energy, or with decreasing particle impact velocity,
density and strength.
References
1. J. Akedo, Aerosol Deposition Method for Fabrication of Nano
Crystal Ceramic Layer, Mater. Sci. Forum, 2004, 449, p 43-48
2. J. Akedo, Aerosol Deposition of Ceramic Thick Films at Room
Temperature: Densication Mechanism of Ceramic Layers, J.
Am. Ceram. Soc., 2006, 89, p 1834-1839
3. J. Akedo, Room Temperature Impact Consolidation (RTIC) of
Fine Ceramic Powder by Aerosol Deposition Method and
Applications to Microdevices, J. Therm. Spray Technol., 2008, 17,
p 181-198
4. J. Akedo, Aerosol Deposition Method (ADM) for Nano-Crystal
Ceramics Coating Without Firing, MRS Symp. Proc., 2003, 778, p
289-296
5. J. Akedo, J. Park, and H. Tsuda, Fine Patterning of Ceramic
Thick Layer on Aerosol Deposition by Lift-Off Process Using
Photoresist, J. Electroceram., 2009, 22, p 319-326
6. S. Ravanbakhsh, H. Assadi, H. Nekoomanesh, A. Hassanzadeh,
and E. Taheri-Nassaj, Cold Spraying of Ceramic CoatingsA
Feasibility Study, Proceedings of ITSC 2011, 27-29 September,
Hamburg, Germany
7. N. Tjitra Salim, M. Yamada, H. Isago, K. Shima, H. Nakano, and
M. Fukumoto, The Understanding on Adhesion Mechanism of
Cold Sprayed TiO
2
Coating, Proceedings of ITSC 2011, 27-29
September, Hamburg, Germany
8. M. Yamada, M.E. Dickinson, K. Shima, N. Tjitra Salim, H.
Nakano, and M. Fukumoto, Deposition Behavior and Adhesion
Strength of Cold-Sprayed TiO
2
Particles, Proceedings of ITSC
2011, 27-29 September, Hamburg, Germany
9. H. Assadi, F. Ga rtner, T. Stoltenhoff, and H. Kreye, Bonding
Mechanism in Cold Gas Spraying, Acta Mater., 2003, 51, p 4379-
4394
10. A. Carpinteri, B. Chiaia, and P. Cornetti, A Scale-Invariant
Cohesive Crack Model for Quasi-Brittle Materials, Eng. Fract.
Mech., 2002, 69, p 207-217
11. A. Carpinteri, P. Cornetti, F. Barpi, and S. Valente, Cohesive
Crack Model Description of Ductile to Brittle Size-Scale Tran-
sition: Dimensional Analysis vs. Renormalization Group Theory,
Eng. Fract. Mech., 2003, 70, p 1809-1839
12. C. Gurney, and J. Hunt, Quasi-Static Crack Propagation, Pro-
ceedings of the Royal Society of London. Series A. Mathematical
and Physical Sciences, Vol 299, 1967, p 508-524
13. K. Kendall, The Impossibility of Comminuting Small Particles by
Compression, Nature, 1978, 272, p 710-711
14. J. Michler, K. Wasmer, S. Meier, F. Ostlund, and K. Leifer,
Plastic Deformation of Gallium Arsenide Micropillars Under
Uniaxial Compression at Room Temperature, Appl. Phys. Lett.,
2007, 90, p 043123
15. F. Ostlund, K. Rzepiejewska-Malyska, K. Leifer, L.M. Hale, Y.
Tang, R. Ballarini, W.W. Gerberich, and J. Michler, Brittle-to-
Ductile Transition in Uniaxial Compression of Silicon Pillars at
Room Temperature, Adv. Funct. Mater., 2009, 19(15), p 2439-
2444
16. F. O

stlund, P.R. Howie, R. Ghisleni, S. Korte, K. Leifer, W.J.


Clegg, and J. Michler, Ductile-Brittle Transition in Micropillar
Compression of GaAs at Room Temperature, Phil. Mag., 2011,
91, p 1190-1199
17. V. Tvergaard, Inuence of Void Nucleation on Ductile Shear
Fractureat a FreeSurface, J. Mech. Phys. Solids, 1982, 30, p399-425
18. X.P. Xu and A. Needleman, Numerical Simulations of Fast Crack
Growth in Brittle Solids, J. Mech. Phys. Solids, 1994, 42, p 1397-
1434
19. R.C. Batra and M.H. Lear, Simulation of Brittle and Ductile
Fracture in an Impact Loaded Prenotched Plate, Int. J. Fract.,
2004, 126, p 179-203
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d
20. F. Li, J. Pan, and C. Sinka, Modelling Brittle Impact Failure of
Disc Particles Using Material Point Method, Int. J. Impact Eng.,
2011, 38, p 653-660
21. D. Sulsky, S.J. Zhou, and H.L. Schreyer, Application of a Parti-
cle-in-Cell Method to Solid Mechanics, Comput. Phys. Commun.,
1995, 87, p 236-252
22. D. Frenkel and B. Smit, Understanding Molecular Simulation:
From Algorithms to Applications, Academic Press, London, 2001
23. J.M. Haile, Molecular Dynamics Simulation: Elementary Meth-
ods, Wiley, Weinheim, 1992
24. M.P. Allen and D.J. Tildesley, Computer Simulation of Liquids,
Oxford University Press, Oxford, 1989
25. P. Jensen, Growth of Nanostructures by Cluster Deposition:
Experiments and Simple Models, Rev. Mod. Phys., 1999, 71,
p 1695
26. R. Biswas, G.S. Grest, and C.M. Soukoulis, Molecular-Dynamics
Simulation of Cluster and Atom Deposition on Silicon (111),
Phys. Rev. B, 1988, 38, p 8154
27. H. Haberland, M. Karrais, M. Mall, and Y. Thurner, Thin Films
from Energetic Cluster Impact: A Feasibility Study, J. Vac. Sci.
Technol. A, 1992, 10, p 3266-3271
28. H. Haberland, Z. Insepov, M. Kurrais, M. Mall, M. Moseler, and
Y. Thurner, Thin Films from Energetic Cluster Impact; Experi-
ment and Molecular Dynamics Simulations, Nucl. Instrum.
Methods Phys. Res. Sect. B, 1993, 80, p 1320-1323
29. K.H. Muller, Cluster-Beam Deposition of Thin Films: A Molec-
ular Dynamics Simulation, J. Appl. Phys., 1987, 61, p 2516-2521
30. J. Kraft, O. Rattunde, O. Rusu, A. Ha fele, and H. Haberland,
Thin Films from Fast Clusters: Golden TiN Layers on a Room
Temperature Substrate, Surf. Coat. Technol., 2002, 158, p 131-135
31. H. Haberland, Z. Insepov, and M. Moseler, Molecular-Dynamics
Simulation of Thin-Film Growth by Energetic Cluster Impact,
Phys. Rev. B, 1995, 51, p 11061
32. D.M. Chun and S.H. Ahn, Deposition Mechanism of Dry
Sprayed Ceramic Particles at Room Temperature Using a Nano-
Particle Deposition System, Acta Mater., 2011, 59, p 2693-2703
33. H. Ogawa, Molecular Dynamics Simulation on the Single Particle
Impacts in the Aerosol Deposition Process, Mater. Trans., 2005,
46, p 1235-1239
34. H. Ogawa, Atomistic Simulationof theAerosol DepositionMethod
with Zirconia Nanoparticles, Mater. Trans., 2006, 47, p 1945
35. H. Ogawa, Molecular Dynamics Simulation on the Modication
of Crystallographic Orientation in Fragmented Particles in the
Aerosol-Deposition Process, Mater. Trans., 2007, 48, p 2067-2071
36. M.M. Mandelbrot, D.E. Passoja, and A.J. Paullay, Fractal
Character of Fracture Surfaces of Metals, Nature, 1984, 308,
p 721-722
37. H. Assadi, T. Schmidt, H. Richter, J.O. Kliemann, K. Binder, F.
Gartner, T. Klassen, and H. Kreye, On Parameter Selection in
Cold Spraying, J. Therm. Spray Technol., 2011, 20, p 1161-1176
Journal of Thermal Spray Technology
P
e
e
r
R
e
v
i
e
w
e
d

You might also like