Vous êtes sur la page 1sur 15

CHAPTER 1

Nanoparticles from Mechanical Attrition


Claudio L. De Castro, Brian S. Mitchell
Department of Chemical Engineering, Tulane University,
New Orleans, Louisiana, USA
CONTENTS
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2. Historical Perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3. Principles of Milling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
4. Attrition Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
4.1. SPEX Shaker Mills . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
4.2. Planetary Ball Mills . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.3. Attritor Mills . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
4.4. Comparison of Mills . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5. Milled Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5.1. Solid-State Amorphization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5.2. Metals, Alloys, and Intermetallics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
5.3. Ceramics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
5.4. Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
5.5. Polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
6. Contamination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
7. Characterization of Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
8. Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1. INTRODUCTION
Unlike nanoparticles produced from bottom-up pro-
cesses such as self-assembly and templated synthesis,
nanoparticles from mechanical attrition are produced by
a top-down process. Such nanoparticles are formed in
a mechanical device, generically referred to as a mill,
in which energy is imparted to a course-grained mate-
rial to effect a reduction in particle size. Under cer-
tain conditions, the resulting particulate powders can
exhibit nanostructural characteristics on at least two lev-
els. First, the particles themselves, which normally pos-
sess a distribution of sizes, can be nanoparticles if their
average characteristic dimension (diameter for spheri-
cal particles) is less than 100 nm [1]. Second, many of
the materials milled in mechanical attrition devices are
highly crystalline, such that the crystallite (grain) size
after milling is often between 1 and 10 nm in diame-
ter. Such materials are termed nanocrystalline [2]. The
sizes of the nanocrystals and the nanoparticles may or
may not be the same. In some of the nanostructured
materials literature, particularly that involving bottom-up
processes, the term nanocrystal is reserved for crys-
talline particles with low concentrations of defects, such
as are found in single crystals, whereas nanoparticles
are those nanoscale particles that contain gross internal
grain boundaries, fractures, or internal disorder, whether
the crystals they contain are nanocrystalline or not [3].
However, we will see that because of the large amount
ISBN: 1-58883-009-8 / $35.00
All rights of reproduction in any form reserved. 1
Synthesis, Functionalization and Surface Treatment of Nanoparticles
Edited by M.-I. Baraton
Copyright 2002 by American Scientic Publishers
2 De Castro and Mitchell
of strain imparted to particles during the milling pro-
cess, it is virtually impossible to obtain defect-free crys-
tals via mechanical attrition. As a result, we will adhere
to the more general denitions; that is, nanocrystals are
110 nm in diameter, and nanoparticles are less than
100 nm in diameter. Thus, it is possible to have both
nanocrystalline nanoparticles and coarser particles that
contain nanocrystals. Both types of materials produced
by mechanical attrition will be addressed in this chapter.
The importance of nanoparticles lies in their inher-
ently large surface-to-volume ratio relative to that of
larger particles. These high surface areas can potentially
improve catalytic processes and interfacially driven phe-
nomena such as wetting and adhesion. Nanoparticles
have the potential for use in structural and device appli-
cations in which enhanced mechanical and physical char-
acteristics are required. As for the internal structure of
the nanoparticles, it has been found that nanocrystalline
materials have comparative advantages over their micro-
crystalline counterparts in hardness, fracture toughness,
and low temperature ductility [4, 5]. As new methods
for surface modication and postattrition processing of
nanoparticles are developed, the potential applications
for them continue to grow.
The early work on the production of nanostructured
materials by mechanical attrition has been reviewed pre-
viously [615]. However, signicant information contin-
ues to become available in this rapidly evolving eld,
particularly with regard to the range of material classes
to which it is applied. We attempt here to provide an
updated review of the subject, as mechanical attrition is
used to form nanoparticles in new materials such as poly-
mers [16] and FCC metals [17], and as improvements
in the milling process that optimize the formation of
nanoparticles continue to be made.
2. HISTORICAL PERSPECTIVE
The attrition, or milling, of materials has been a major
component of the ceramic processing and powder metal-
lurgy industries for many years. The objectives of milling
include particle size reduction (comminution or grind-
ing); amorphization; particle size growth; shape changing
(aking); agglomeration; solid-state blending (incomplete
alloying); modifying, changing, or altering properties of
a material (density, owability, or work hardening); and
mixing or blending of two or more materials or mixed
phases. However, the primary objective of milling is often
purely particle size reduction.
Mechanical attrition began as a way to simultaneously
blend and commute (decrease in size) metal powders in
a process called mechanical alloying. Mechanical alloy-
ing (MA) is a powder technique that allows production
of homogeneous materials from blended elemental pow-
der mixtures. John Benjamin and his colleagues at the
Paul D. Merica Research Laboratory of the International
Nickel Company (INCO) developed the process around
1966. The technique was the result of an intense research
effort to produce nickel-based superalloys for gas turbine
applications. Benjamin has summarized the historical ori-
gins of the process and the background work that led to
the development of MA [1820]. To summarize, they pro-
duced a nickel-chromium-aluminum-titanium alloy rst
produced in a small high-speed shaker mill and later in a
one-gallon stirred ball mill, starting the birth of MA as a
method to produce oxide dispersion-strengthened (ODS)
alloys on an industrial scale. The process developed by
Benjamin was initially referred to as milling/mixing and
later termed mechanical alloying by Ewan C. Mac-
Queen, a patent attorney for INCO. The formation of
an amorphous phase by mechanical grinding of a Y-Co
intermetallic compound in 1981 [21] and in the Ni-Nb
system by ball milling of blended elemental mixtures
in 1983 [22] established MA as a potential nonequilib-
rium processing technique. Beginning in the mid-1990s, a
number of investigations were carried out to synthesize a
variety of stable and meta-stable phases, including super-
saturated solid solutions, crystalline and quasi-crystalline
intermediate phases, and amorphous alloys [2327]. Since
then, mechanical alloying has been applied to virtually all
material classes, including metals, ceramics, and polymers.
3. PRINCIPLES OF MILLING
The fundamental principle of size reduction in mechan-
ical attrition devices lies in the energy imparted to
the sample during impacts between the milling media.
The impact process is shown in Figure 1. This model
Figure 1. Model of impact event at a time of maximum impacting force,
showing the formation of a microcompact. Reprinted with permission
from [28], E. Kuhn, ASM Handbook, Vol. 7, Materials Park, OH, 1984.
1984, ASM.
Nanoparticles from Mechanical Attrition 3
Figure 2. Process of trapping an incremental volume of powder
between two balls in a randomly agitated charge of balls and pow-
der. (ac) Trapping and compaction of particles. (d) Agglomeration. (e)
Release of agglomerate by elastic energy. Reprinted with permission
from [28], E. Kuhn, ASM Handbook, Vol. 7, Materials Park, OH, 1984.
1984, ASM.
represents the moment of collision, during which par-
ticles are trapped between two colliding balls within a
space occupied by a dense cloud, dispersion, or mass
of powder particles [28]. Figure 2 shows the process of
trapping an incremental volume between two balls. Com-
paction begins with a powder mass that is characterized
by large spaces between particles compared with the par-
ticle size. The rst stage of compaction starts with the
rearrangement and restacking of particles. Particles slide
past one another with a minimum of deformation and
fracture, producing some ne, irregularly shaped parti-
cles. The second stage of compaction involves elastic and
plastic deformation of particles [29]. Cold welding may
occur between particles in metallic systems during this
stage. The third stage of compaction, involving particle
fracture, results in further deformation and/or fragmen-
tation of the particles.
For brittle materials, particle fracture is well described
by Grifth theory. According to the theory, the stress, u
F
,
at which crack propagation leading to catastrophic failure
(fracture) occurs in the particle is approximated by
u
F

1
c
(1)
Table 1. Selected Youngs moduli for isotropic solids at room
temperature.
Material 1 (10
6
psi) 1 (10
10
N/m
2
)
Aluminum 10 7
o-Iron 30 21
Copper 16 11
Silicon 16 11
Tungsten 59 41
Titanium carbide (TiC) 45 31
Alumina (Al
2
O
3
) 58 40
Magnesia 45 31
Hard rubber (ebonite) 0.6 0.4
Polystyrene (atactic) 0.4 0.3
Polyethylene (branched) 0.03 0.02
where c is the length of the crack, 1 is the modulus
of elasticity (Table 1), and is the surface energy of
the milled substance (Table 2). When stress at the crack
tip equals the strength of cohesion between atoms, the
crack becomes unstable and propagates, leading to frac-
ture [28]. As fragments decrease in size, the tendency
to aggregate increases, and fracture resistance increases.
Particle neness approaches a limit as milling continues
and maximum energy is expended [30]. According to Har-
ris [31], the major factors contributing to grind limit are:
Increasing resistance to fracture.
Increasing cohesion between particles, with decreas-
ing particle size causing agglomeration.
Excessive clearance between impacting surfaces.
Coating of the grinding medium by ne particles that
cushion the microbed of particles from impact.
Surface roughness of the grinding medium.
Bridging of large particles to protect smaller parti-
cles in the microbed.
Increasing apparent viscosity as particle size
decreases.
Table 2. Selected surface energies of solid materials.
Material Surface energy, (J/m
2
)
Au 1.40
Cu 1.431.70
Ag 1.141.20
Ni 1.90
NaCl 0.30
Al
2
O
3
0.905
MgO 1.001.20
TiC 1.19
LiF 0.34
CaF
2
0.45
BaF
2
0.28
CaCO
3
0.23
Si 1.24
Poly(tetrauoroethylene) 0.0185
Poly(ethylene) 0.033
Poly(styrene) 0.034
4 De Castro and Mitchell
Figure 3. Change of average particle size as a function of milling time
for MA Fe-20at.%Co powders by (a) xed r.p.m. operation and (b)
cyclic operation. Reprinted with permission from [30], Y. D. Kim et al.,
Mater. Sci. Eng. A 291, 17 (2000). 2000, Elsevier Science.
Decreasing internal friction of slurry as particle size
decreases.
The grinding limit is clearly demonstrated in the studies
of Kim et al. [30]; after milling of Fe-Co powders for 30 h,
the MA process reached a steady state where the parti-
cles have become homogenized in size and shape (Fig. 3).
This work also clearly distinguishes between particle size,
which approaches a limit of 10 m with a planetary ball
mill, and crystallite size, which approaches 10 nm (Fig. 4).
Figure 4 also illustrates the strain buildup in the particles
as milling time increases and grain sizes decrease.
4. ATTRITION DEVICES
Different types of milling equipment are available
for mechanical alloying and nanoparticle formation.
Figure 4. Changes of average grain size and strain for MA Fe-20at.%
Co powders with milling time by cyclic operation. Reprinted with per-
mission from [30], Y. D. Kim et al., Mater. Sci. Eng. A 291, 17 (2000).
2000, Elsevier Science.
Figure 5. SPEX 8000D dual mixer/mill.
They differ in their capacity, efciency of milling, and
additional arrangements for heat transfer and particle
removal. A brief description of the different mills avail-
able for MA can be found below.
4.1. SPEX Shaker Mills
Shaker mills such as SPEX (Fig. 5), which mill about
1020 g of powder at a time, are most commonly used
for laboratory investigations and for alloy screening pur-
poses. The common variety of the mill has one vial (see
Fig. 6), containing the sample and the grinding media,
which is secured in the clamp and swung energetically
back and forth several thousands times a minute. The
back-and-forth shaking motion is combined with lateral
movements of the ends of the vial. With each swing of
the vial the milling media, typically hard, spherical objects
called milling balls, impact against each other and the
end of the vial, both milling and mixing at the same time.
Because of the amplitude (about 5 cm) and speed (about
1200 rpm) of the clamp motion, the ball velocities are
high (on the order of 5 m/s), and consequently the force
of the balls impact is usually great. Therefore, these mills
can be considered a high-energy variety.
The most recent design of this mill has provision
for simultaneously milling the powders in two vials
to increase the throughput. This machine incorporates
Figure 6. SPEX stainless steel vial set for SPEX 8000D mill.
Nanoparticles from Mechanical Attrition 5
forced cooling to permit extended milling times. A vari-
ety of vial materials are available for the SPEX mills,
including hardened steel, alumina, tungsten carbide, zir-
conia, stainless steel, silicon nitride, agate, plastic, and
polymethylmethacrylate. A majority of the research on
the fundamental aspects of MA has been carried out with
some version of these SPEX mills.
4.2. Planetary Ball Mills
Another popular mill for conducting MA experiments is
the planetary ball mill (referred to as Pulverisette) in
which a few hundred grams of the powder can be milled
at a time. The planetary ball mill owes its name to the
planet-like movement of its vials. These are arranged on
a rotating support disk, and a special drive mechanism
causes them to rotate around their own axes. The cen-
trifugal force produced by the vials rotating around their
own axes and that produced by the rotating support disk
both act on the vial contents, consisting of material to be
ground and the grinding balls. Since the vials and the sup-
porting disk rotate in opposite directions, the centrifugal
forces alternately act in like and opposite directions. This
causes the grinding balls to run down the inside of the
vialthe friction effectfollowed by the material being
ground. Grinding balls lift off and travel freely through
the inner chamber of the vial and collide against the
opposing inside wallthe impact effect.
Even though the disk and the vial rotation speeds could
not be independently controlled in the earlier versions of
this device, it is possible to do so in modern versions. In a
single mill there can be either two (Pulverisette 5 or 7) or
four (Pulverisette 5) milling stations. Recently, a single-
version mill was also developed (Pulverisette 6). Grinding
vials and balls are available in a variety of different mate-
rials, including agate, silicon nitride, sintered corundum,
zirconia, chrome steel, Cr-Ni steel, tungsten carbide, and
polyamide. An example of the particles that result from
attrition in a planetary mill is shown in Figure 7.
4.3. Attritor Mills
A conventional ball mill consists of a rotating horizon-
tal drum half-lled with steel balls that range from 0.318
to 0.635 cm in diameter. As the drum rotates the balls
drop on the metal powder that is being ground; the rate
of grinding increases with the speed of rotation. At high
speeds, however, the centrifugal force acting on the steel
balls exceeds the force of gravity, and the balls are pinned
to the wall of the drum. At this point the grinding action
stops. An attritor (a ball mill capable of generating higher
energies) consists of a vertical drum with a series of
impellers inside it. Set progressively at right angles to
each other, the impellers energize the ball charge, causing
powder size reduction due to the impact between balls;
between the balls and the container wall; and between
the balls, the agitator shaft, and the impellers. Some size
Figure 7. SEM of Bi
4
Ti
3
O
12
milled for different times: (a) 3, (b) 9, (c)
15, and (d) 20 h. Reprinted with permission from [43], L. B. Kong et al.,
Mater. Lett. 51, 108 (2001). 2001, Elsevier Science.
reduction appears to take place by interparticle collisions
and ball sliding. A motor rotates the impellers, which in
turn agitate the steel balls in the drum.
Attritors (Fig. 8) are the mills in which large quanti-
ties of powder (from about 0.5 to 40 kg) can be milled
at a time. Attritors of different sizes and capacities are
available. The grinding tanks or containers are avail-
able in stainless steel or stainless steel coated with alu-
mina, silicon carbide, silicon nitride, zirconia, rubber,
or polyurethane. A variety of grinding media are also
available: glass, int stones, stealite ceramic, mullite, sil-
icon carbide, silicon nitride, Sialon, alumina, zirconium
Figure 8. Attrition ball mill. Reprinted with permission from [28],
E. Kuhn, ASM Handbook, Vol. 7, Materials Park, OH, 1984. 1984,
ASM.
6 De Castro and Mitchell
silicate, zirconia, stainless steel, carbon steel, chrome
steel, and tungsten carbide.
The operation of an attritor is simple. The powder to
be milled is placed in a stationary tank with the grinding
media. The mixture is then agitated by a shaft with arms,
rotating at a high speed of about 250 rpm. This causes
the media to exert both shearing and impact forces on
the material. The laboratory attritor works up to 10 times
faster than conventional ball mills.
4.4. Comparison of Mills
The nal product size from mechanical attrition is deter-
mined by the energy input during milling, the ball-to-
powder weight ratio, and the overall temperature during
milling. Borner and Eckert [32] investigated the effect
of energy input by milling iron powders with the use of
a SPEX milling machine and a Pulverisette 5, among
other lower energy mills. They determined that the SPEX
shaker mill provides the largest input and therefore leads
to a fast decrease of grain size to less than 20 nm. The
steady-state grain size was achieved after 4 h of milling.
The Pulverisette mill provides a smaller energy impact
during the collision. After 32 h of milling the grain size
achieved was 40 nm at 90 rpm, 31 nm at 180 rpm, and
20 nm at 360 rpm (Fig. 9).
The different classications of mills and the typical
range of particle sizes they produce are summarized in
Figure 10. Notice that vibratory mills are the typical class
of mills used to produce nanoparticles. Figure 11 shows
how the different mill types can be used to attrit materials
of all types, ranging from the more commonly used brittle
materials, to the newer, plastic and viscoelastic materials.
In the next section, we describe recent research results
on nanoparticle formation and nanocrystallinity in these
different materials classes.
Figure 9. Average grain size for Fe powders vs. milling time with differ-
ent mills. Reprinted with permission from [32], I. Borner and J. Eckert,
Mater. Sci. Eng. A 226228, 541 (1997). 1997, Elsevier Science.
Figure 10. Typical size capabilities of common classes of size reduc-
tion equipment. Reprinted with permission from [28], E. Kuhn, ASM
Handbook, Vol. 7, Materials Park, OH, 1984. 1984, ASM.
5. MILLED MATERIALS
Though traditionally limited to metallic materials, in par-
ticular FCC metals, mechanical alloying is now being uti-
lized to form nanoparticles and nanocrystalline particles
from a variety of materials. We report here on recent
research results that employ MA as the principal pro-
cessing method and the primary benets it provides for
different types of materials.
Figure 11. Applicability of various classications of comminution
machines to different material types.
Nanoparticles from Mechanical Attrition 7
5.1. Solid-State Amorphization
A solid alloy with a noncrystalline atomic arrangement
is called an amorphous (metallic) alloy. The synthesis of
an amorphous phase in the Ni-Nb system by MA starting
from blended elemental powders of Ni and Nb in 1983
[22] has given rise to increased research activity in this
area. Amorphous phases have been synthesized by MA
from blended elemental powder mixtures, pre-alloyed
powders and/or intermetallics, mixtures of intermetallics,
or mixtures of intermetallics and elemental powders.
The effect of process variables on amorphization
behavior has been studied in several alloy systems. The
most important variables that have been studied are
milling energy and milling temperature. Increased milling
energy, which is achieved by higher ball-to-powder weight
ratio and increased speed of rotation, is expected to intro-
duce more strain and increase the defect concentration
in the powder, thereby leading more readily to amor-
phization. On the other hand, higher milling energies can
also produce more heat, resulting in crystallization of the
amorphous phase. A balance between these two effects
will determine the nature of the nal product phase.
Eckert et al. [33] reported that in Ni-Zr system milled in
a planetary ball mill at a low intensity did not produce any
amorphous phase, because of the lack of energy. How-
ever, when the intensity was increased, amorphous phase
formation was observed in a wide composition range of
3083% Ni. At even higher intensities, amorphous phase
formation was observed between 66 and 75% Ni. These
observations suggest that with increasing milling energy,
the heat generated is also high, which crystallizes the
amorphous phases. It appears that the maximum amor-
phization range is observed at intermediate values of
milling intensity; too low an intensity does not provide
enough energy to amorphize, while at very high intensi-
ties, the amorphous phase formed would crystallize.
There have been conicting results on the effect of
milling temperature on the nature of the phase formed.
Koch et al. [34] summarized the results of varying the
temperature of the mill on the kinetics of amorphization
in intermetallics. They concluded that generally a lower
milling temperature accelerated the amorphization pro-
cess. Since a nanostructured material can easily be pro-
duced at lower milling temperatures, the increased grain
boundary area drives the crystal-to-amorphous transfor-
mation. For example, the time required for amorphiza-
tion in NiTi was 2 h at 190

C, 13 h at 60

C, and 18 h
at 220

C [34].
Mechanical alloying introduces contamination into the
milled powder and thus alters the constitution and stabil-
ity of powder products. Generally, the presence of addi-
tional elements favors amorphization. Contamination has
also been found to affect the stability of the amorphous
phases formed. Formation of a crystalline phase has been
reported on milling of Ti-Al powders after an amorphous
phase had been formed [35]. Lattice measurements sug-
gested that this FCC crystalline phase is due to the
increased nitrogen contamination of the milled powder,
and the crystalline phase has been identied as TiN.
5.2. Metals, Alloys, and Intermetallics
Metals are routinely milled to the nanoparticle size range,
with typical crystallite sizes in the 110-nm range. The
average grain size of a NiAl intermetallic compound was
as low as 5 nm after 100 h of milling [36] and as low as
4.7 nm for Fe-C alloys milled in a horizontal ball milling
machine [37].
Kim et al. [30] studied the effect of xed and cyclic
operation on the formation of nanocrystalline Fe-Co
alloy powders produced by mechanical alloying in a plan-
etary ball mill. The rst method was conventional milling
with xed velocity (1300 rpm). The second method
involved an operation cycle characterized by a time inter-
val of 4 min at 1300 rpm followed by 1 min of oper-
ation at 900 rpm. In both cases, the milling time was
varied from 1 to 100 h. It is generally known that the
MA process reaches a steady state where the particles
have homogeneous size and shape. The steady state was
achieved after 30 h of milling for xed operation and
15 h of milling for a cyclic operation. The advantage of
the cyclic operation is due to the fact that the periodical
changes of the milling velocity in cyclic operation break
the balance of deformation, fracture, and welding in the
process and maximize the effect of fracture. The steady-
state crystallite grain size obtained was in the range of
1015 nm. Interestingly, during the rst hour of milling
the grain size of Fe-Co increased and then decreased to
the nal 1015 nm after achieving steady state.
Raghu et al. [38] studied the differences between
nanocrystalline copper-tungsten alloys milled in argon
and air. They determined that the particle size initially
decreases and then increases in powders milled in air
(Fig. 12). The crystallite sizes of copper and tungsten
decrease continuously with milling time in all milling
atmospheres (Fig. 13). Crystallite size levels off at a value
of 10 nm for copper and 15 nm for tungsten after 8 h of
milling. It was concluded that whereas fracture is dom-
inant in milling in argon, adequate welding for the MA
process appears to be available during milling in air.
Eckert and Borner [36] studied nanostructure forma-
tion of ball-milled NiAl intermetallic compounds in a
planetary ball mill with hardened steel balls and vial at a
rotation speed of 180 rpm. Figure 14 illustrates that the
average grain size and the rms strain scale with composi-
tion of the material. Whereas the grain size is reduced to
only 12 nm for Ni
46
Al
54
, it decreased to about 5 nm for
Ni-rich compounds. Blending Ni
46
Al
54
with elemental Ni
to give an overall composition of Ni
50
Al
50
and subsequent
mechanical alloying reduced the grain size from 12 nm
after 100 h to about 9 nm after 300 h of total milling time
8 De Castro and Mitchell
Figure 12. Particle size of Cu-5vol%W powders milled in different
atmospheres. Reprinted with permission from [38], T. Raghu et al.,
Mater. Sci. Eng. A 304306, 438 (2001). 2001, Elsevier Science.
(Fig. 15). On the other hand, the grain size of 100-h ball-
milled Ni
60
Al
40
increased upon blending with elemental
Al and further mechanical alloying. This demonstrates
that the ultimate grain size is coupled with the com-
position of the material. Table 3 summarizes grain size
results obtained by milling different metals, alloys, and
intermetallics.
Figure 13. Average grain size of (a) copper and (b) tungsten powders
milled in argon and air. Reprinted with permission from [38], T. Raghu
et al., Mater. Sci. Eng. A 304306, 438 (2001). 2001, Eslevier Science.
Figure 14. Average grain size (lled symbols) and atomic-level strain
(open symbols) for Ni
x
Al
100x
powders after 100 h of milling as a func-
tion of composition. Reprinted with permission from [36], J. Eckert and
I. Borner, Mater. Sci. Eng. A 239240, 619 (1997). 1997, Elsevier
Science.
Figure 15. Average grain size obtained for Ni
56
Al
54
and Ni
60
Al
40
pow-
ders after 100 h of ball milling and after addition of elemental Ni or
Al and subsequent mechanical alloying as a function of total milling
time. For comparison the average grain size for Ni
50
Al
50
obtained after
100 h of continuous ball milling is also shown. Reprinted with permis-
sion from [36], J. Eckert and I. Borner, Mater. Sci. Eng. A 239240, 619
(1997). 1997, Elsevier Science.
Table 3. Typical grain sizes for metals, alloys, and intermetallics for
MA.
Milling Milling
Composition technique Grain size (nm) time (h) Reference
Fe-Co powders Rotary ball mill 1015 30 [30]
Fe Vibratory mill 20 4 [32]
NiAl Vibratory mill 12 100 [36]
Ni silicides Vibratory mill 1017 30 [1]
Fe-C Horizontal ball 4.7 500 [37]
mill
Fe
3
Al Vibratory mill 12.6 100 [76]
Nanoparticles from Mechanical Attrition 9
5.3. Ceramics
Recent studies show that highly exothermic reactions, for
example, the formation of TiC [39], can be initiated by
high-energy ball milling. This process is referred to as
a mechanically induced self-propagating reaction (MSR).
In the studies of Xinkun et al. [39], raw powders of Ti
and C were mixed in a mole ratio of 1:1 and milled.
TiC particles with 20-nm crystallites were fabricated just
after reaction at 120 min. It took 10 to 20 h for the Ti
and C powders to react completely. The reaction of Ti
and C powders was investigated by measuring the tem-
perature of the vial outer wall. Figure 16 shows the rela-
tion between the temperature of the outer wall and the
milling time and indicates that the temperature of the vial
increased slowly as the milling time was extended. When
the milling time reached 115 min, the temperature of the
vial increased abruptly and reached its maximum (from
319 to 332 K). The temperature peak was associated with
the exothermic reaction of Ti and C. After the peak, the
vial temperature decreased gradually during subsequent
milling.
Traditionally, the raw materials used in mechanical
alloying include at least one ductile material used as a
host or binder for the other ingredients. However, Davis
and Koch [40] mechanically alloyed a brittle Si-Ge sys-
tem to obtain a solid solution Si(Ge) and observed a lat-
tice change with the increased duration of MA. Recently,
Zhang and Tam [41] studied the mechanical alloying of
an all-ceramic-phase component, TiC + TiN. Two com-
positions were explored, 70% TiC + 30% TiN and 50%
TiC + 50% TiN (by weight). MA was conducted for up
to 80 h in a Fritsch planetary ball mill at a ball-to-powder
ratio of 20:1. Figure 17 shows a plot of the lattice param-
eter as a function of milling time for 70% TiC + 30%
TiN milled with steel balls. A similar plot was obtained
for 50% TiC + 50% TiN milled with WC balls. The two
compositions show a similar trend in lattice parameter
change, irrespective of milling media, and therefore it is
indicated that solid solutions occur during the MA pro-
cess. It was also determined that the rate of change of the
Figure 16. Temperature of outer vial wall as a function of milling time
for milling of Ti and C powders. The sharp increase in wall temperature
at 115 min is attributed to formation of TiC. Reprinted with permission
from [39], Z. Xinkun et al., Mater. Sci. Eng. C 16, 103 (2001). 2001,
Elsevier Science.
Figure 17. Lattice parameter as a function of MA time for 70wt%
TiC + 30wt% TiN as determined from XRD patterns. Reprinted with
permission from [41], S. Zhang and S. C. Tam, J. Mater. Processing 67,
12, (1997). 1997, Elsevier Science.
lattice parameter with MA time is almost the same for
the two compositions; this shows that in MA, the reac-
tion rate is independent of the concentration of the
reactant as long as the latter is not depleted, which
is different from a chemical reaction, where the reac-
tion rate is proportional to the concentration of reac-
tants. The reduction of crystallite size was very rapid in
the beginning, and after 40 h it was in the range of 20
30 nm (Fig. 18). In the studies of Indris et al. [42], the
system (1 x)Li
2
O:xB
2
O
3
reacts either at high temper-
atures or because of mechanochemical treatment. Indris
determined that the reaction product is not determined
by the overall composition of the mixture, but by the con-
ditions at the particle interfaces.
Indris et al. [42] also studied the oxide ceramics Li
2
O,
LiNbO
3
, LiBO
2
, B
2
O
3
, and TiO
3
as monophase materi-
als. Nanostructures of these oxides were prepared in a
Figure 18. Reduction in average grain size as a function of milling time
for 50wt% TiC + 50wt% TiN. Reprinted with permission from [41],
S. Zhang and S. C. Tam, J. Mater. Processing 67, 112, (1997). 1997,
Elsevier Science.
10 De Castro and Mitchell
Figure 19. Average grain size of Li
2
O, LiNbO
3
, and B
2
O
3
powders for
various milling times. Reprinted with permission from [42], S. Indris
et al., J. Mater. Synth. Processing 8, 245 (2000). 2000, Plenum.
high-energy ball mill Spex 8000 with alumina milling vials
with a ball-to-powder ratio of 2:1 to produce at least 2 g
of powder. Figure 19 shows the average grain size for
LiNbO
3
, Li
2
O, and B
2
O
3
. Although the kinetics of grain
size reduction is different, all samples show a similar sat-
uration value for a nal grain size of about 23 nm.
Kong et al. [43] prepared Bi
4
Ti
3
O
12
ceramics by com-
bining Bi
2
O
3
and TiO
2
powders. Although the particle
size was reduced to 100200 nm after 3 h, no reaction
between the powders took place. After 9 h of milling,
peaks for Bi
4
Ti
3
O
12
appeared in the XRD pattern, indi-
cating reaction between the starting powders. After 15 h,
the reaction was completed, with a further reduction in
particle size. Kong et al. also studied lead zirconium
titanate (PZT) ceramics [44, 45] because of its excel-
lent dielectric, piezoelectric, and electro-optic proper-
ties. PZT powders were formed in a planetary ball mill
from PbO, ZrO
2
, and TiO
2
. The grain size from XRD
ranged between 35 nm (4 h of milling) and 26 nm (24 h
of milling). The particle size, as determined with SEM,
ranged from 120 nm to 65 nm.
5.4. Composites
Mechanical alloying (MA) has been successfully used to
synthesize a number of commercially important alloys
and composites [4651]. The mechanically alloyed pow-
ders have attributes such as ne powder size and exi-
bility of alloying choices. In particular, MA has unique
advantages in producing metal matrix composite pow-
ders. The low-temperature solid-state process eliminates
reactions between the reinforcement particles and the
matrix. It is commonly known that for particulate rein-
forcement of metal matrix composites, the mechanical
properties are inuenced signicantly by the quantity and
the distribution mode of the reinforcements and by the
nature of the interfaces between the reinforcements and
the matrix. Finer size reinforcements produce higher spe-
cic mechanical properties. However, they also encounter
signicant agglomeration problems for that size range,
resulting in poor dispersion and subsequently less than
optimal mechanical properties.
The main problem with the ne-grained structure is
its instability at high temperatures. Because of the large
excess of free energy, signicant grain growth has been
observed in several nanocrystalline materials [52, 53].
Senkov et al. [53] studied grain growth of TiAl
3
/Ti
5
Si
3
composites produced by mechanical alloying and hot
isostatic pressing (HIP). Blends of elemental and pre-
alloyed powders were used for the MA. Intimate mixing
of the elements and powder amorphization occurred dur-
ing the MA. The average grain size increased from 140
to 540 nm when the HIP temperature was increased from
850

C to 1100

C.
Hwang and Nishimura [54] demonstrated that mag-
nesium matrix nanocomposites could be mechanochem-
ically synthesized by milling elemental precursors, of
which two of the elemental powders (normally, a
metal and an element such as C) form the secondary
particles. Nanocomposites have also been successfully
mechanochemically synthesized by Fan et al. [55] with
Nb-Si-Ti-C mixtures, Zhou et al. [56] with Ni-Al-Ti-C mix-
tures, and Takahashi and Hashimoto [57] in the Cu-M-C
system (where M = Zr, Hf, Nb, Ta, or Ti). Hwang and
Nishimura [54] used a mixture of Mg, Ti, and C powder
to synthesize nanometer-sized TiC particles embedded
in a nanocrystalline Mg matrix by mechanochemical
reaction. Powder compositions corresponding to Mg-
xTiC (where x = 515 volume %) were milled. The
XRD pattern shows the development of the Mg-15%TiC
nanocomposite during milling (Fig. 20). After 3 h of
milling, the sample still contained the three forms of
elemental powders, but with further milling (6 h), Ti
peaks were reduced and TiC peaks appeared. After 24 h
of milling, the powder showed only the Mg and TiC
Figure 20. Evolution of Mg-Ti-C nanocomposite during milling and
after heat treatment. Reprinted with permission from [54], S. Hwang
and C. Nishimura, Scripta Mater. 44, 2457 (2001). 2001 Elsevier
Science.
Nanoparticles from Mechanical Attrition 11
Figure 21. Average grain size of MmM
5
phase in MmM
5
-Mg composite
prepared by 20 h of milling as a function of Mg content. Reprinted
with permission from [58], M. Zhu et al., Mater. Sci. Eng. A 286, 130
(2000). 2000, Elsevier Science.
peaks, indicating that Ti and C reacted during milling
to form TiC. The grain sizes of TiC particles and Mg
grains of as-milled samples were approximately 5.6 and
43.1 nm, respectively, and those of heat-treated samples
were approximately 7.1 and 62.3 nm, respectively. Mg-
Ti-C nanocomposites exhibited remarkably high ductility
while retaining compressive strength similar to that of
Mg-TiC nanocomposite.
Zhu et al. [58] studied the effect of Mg content
on the microstructure of mechanical alloyed
MmNi
3.5
(CoAlMn)
1.5
-Mg (abbreviated as MmM
5
-Mg).
Figure 21 shows that the grain size of the MmM
5
phase
increases with Mg content in the composite. The authors
explained this observation by the consideration that Mg
absorbed part of the energy imparted to the MmM
5
phase. MmM
5
particles were gradually embedded in
Mg as milling proceeded. The larger the amount of
Mg in the composite, the less energy was consumed in
rening the grain size of MmM
5
and thus the grain size
of MmM
5
increased with increasing Mg content.
Lee and Munir [59] studied the synthesis of dense
TiB
2
-TiN nanocrystalline composites through mechanical
and eld activation. Powder mixtures were mechanically
activated through ball milling. The powders were blended
in a stoichiometric ratio according to the reaction
Ti +0.5B +0.5BN 0.5TiB
2
+0.5TiN
They were mixed in a Shaker mixer (Glenmill) for 24 h
and ball milled in a planetary mill. Figure 22 shows the
renement of the grain size and the increase in strain as
a function of milling times for Ti. The small grain size of
unmilled Ti powder (71111 nm) may be related to the
particular preparation method of this powder, namely,
that it was made from sponge Ti. As shown in Figure 22,
the strain value changed very little in early stages of
milling, but after 2 h of milling, the strain increased
markedly. Within the rst 6 h of milling, the decrease
in crystallite size for Ti was relatively linear with milling
Figure 22. Average grain size and strain for Ti as a function of milling
time, for the WilliamsonHall (WH) and HalderWagner (HW) meth-
ods. Reprinted with permission from [59], J. W. Lee and Z. A. Munir,
J. Am. Ceram. Soc. 84, 1209 (2001). 2001, The American Ceramic
Society.
time. Figure 22 also shows that the minimum crystal-
lite size of Ti that can be obtained before the onset of
a reaction in the mill is in the range of 16.733.3 nm.
Figure 23a shows the results of crystallite size analyses
on milled powders (10 h) that were subsequently reacted
in the spark plasma synthesis (SPS) apparatus for 1 or
12 min, and Figure 23b shows the results on powders
that had reacted during 12 h of milling and subsequently
reacted in the SPS for 5 to 12 min. It is interesting to
note that the crystallite size of the product phases after 1
min of SPS treatment is roughly the same whether they
are formed in the SPS or previously during milling. How-
ever, when the treatment is longer (12 min), the size of
those formed in the SPS is about one-half that of those
formed during milling.
5.5. Polymers
Unlike the materials previously described here, there
is little work on nanoparticle polymers by mechanical
alloying. We include here recent progress toward the
formation of polymer nanoparticles and nanocrystalline
polymer particles.
High-energy milling of polymeric materials subjects
the blend components to a complex deformation eld
in which shear, multiaxial extension, fracture, and cold
welding proceed concurrently. This technique has been
successfully used to prepare blends of thermoplastics
[6065] and to cause allotropism in semicrystalline poly-
mers [66, 67]. Smith et al. have demonstrated that
nanostructured polymer blends composed of poly(methyl
methacrylate) (PMMA) and either poly(ethylene-alt-
propylene) (PEP) or polyisopropene (PI) can be
12 De Castro and Mitchell
Figure 23. Average grain size of TiN and TiB
2
formed in spark plasma
sintering (SPS). (a) Reactant powders milled for 10 h (no product for-
mation). (b) Reactant powders milled for 12 h (with powder formation).
Reprinted with permission from [59], J. W. Lee and Z. A. Munir, J. Am.
Ceram. Soc. 84, 1209 (2001). 2001, The American Ceramic Society.
produced by cryogenic mechanical alloying [16]. In cryo-
genic mechanic alloying the vial is sealed under argon
(-10 ppm O
2
) and placed in a custom-designed nylon
sleeve that allows peripheral circulation of liquid nitro-
gen. The temperature is maintained at 180

C or below
[68].
In the PEP/PMMA blend pressed at room tempera-
ture and cryomilled for 1 h, the PMMA particles mea-
sured 0.23 m in size and possessed irregular shapes.
As milling time increased, the PMMA particles were
seen to decrease in size; the particles were about 1 m
across after 3 h of cryomilling and 0.5 m across in
the blend cryomilled for 5 h. After 8 h of milling, the
outlining of PMMA by PEP appears to be incomplete
but demonstrates that the PMMA domain is reduced to
300 nm. These results indicate that PEP can be highly
dispersed within PMMA by mechanical alloying. In the
PI/PMMA blends pressed at room temperature, the PI
ows around and delineates the PMMA particles, which
decrease in size with increasing milling time. After 1 h of
cryomilling, the PMMA particles measure less that 2 m
across and subsequently decrease to less than 300 nm
across in the blend milled for 10 h. They also deter-
mined that in PEP/PMMA, temperature had little effect
on blend morphology; however, for the PI/PMMA blends
milled for 10 h and consolidated at 125

C, the PI formed
isolated domains (measuring up to 30 nm across), as
well as interconnected regions of PMMA. At 200

C, the
PI/PMMA blend retained a semicontinuous nanoscale PI
network in which the smallest PI features were about
5 nm across and PMMA-rich inclusions were as small as
10 nm across.
Smith et al. also studied cryogenic mechanical alloy-
ing as an alternative strategy for the recycling of tires
by examining the degree of dispersion achieved in CMA
blends composed of PMMA or PET and tire [69]. After
5 h of milling, PMMA/tire blends with 25 and 10 wt%
tire formed dispersions of the tire within a continu-
ous PMMA matrix due to CMA, and they were as
small as 300 nm and as large as 3 m across. Ternary
PI/tire/PMMA blends with total rubber concentration of
either 10 or 25-wt% revealed that the rubber is highly dis-
persed in PMMA, forming domains as small as 100 nm.
Ball milling has also been used to simply blend poly-
mers, or amorphitize them, in a way similar to the way
metals are amorphitized [7072]. Milling overcomes some
problems associated with conventional blending meth-
ods such as thermal degradation due to excessive heat-
ing in the melting process, or the difculty in removing
the polymer from the solvent if the solution method is
used. These authors also studied the partial amorphiza-
tion of semicrystalline poly(ether-ether ketone) (PEEK)
by means of a ball-milling technique. The milling device
used was a Fritsch (Pulverisette 6) centrifugal ball mill
working with a rotation speed of about 500 rpm. The
sample mass milled was 2 g, with a product-to-ball ratio
of 1:30. PEEK samples of 10 mg were analyzed by clas-
sical differential scanning calorimetry. The analysis of
XRD results revealed that ball milling leads to a partial
amorphization of PEEK (Fig. 24), therefore concluding
Figure 24. Room-temperature XRD patterns for PEEK samples. (a)
Unmilled. (b) After 24 h of milling. Reprinted with permission from
[71], J. Font et al., Mater. Res. Bull., 36, 1665 (2001). 2001, Elsevier
Science.
Nanoparticles from Mechanical Attrition 13
that in the case of semicrystalline PEEK, a polymer with
a higher degree of amorphization was obtained as a result
of this mechanical treatment.
6. CONTAMINATION
A major concern in the processing of nanoparticles by
MA is the nature and amount of impurities that contam-
inate the milled powder. Contamination can arise from
several sources, including
impurities in starting powders,
vials and grinding media,
milling atmosphere, and
control agents added to the powders.
During MA the powder particles become trapped
between the grinding medium and undergo severe plastic
deformation; fresh surfaces are created because of the
fracture of the powder particles. Collisions occur between
the grinding medium and the vial and among the grinding
balls. All of these effects can cause deterioration of the
grinding medium and vial, resulting in the incorporation
of these impurities into the powder. The extent of con-
tamination increases with increasing milling energy and
milling time.
Various attempts have been made to minimize powder
contamination during MA. One way of minimizing the
contamination from the grinding medium and the con-
tainer is to use a material for the container and grinding
medium that is the same as the powder being milled.
For example, one could use copper milling balls and a
copper vial for milling copper and copper alloy powders.
However, milling effectiveness may be limited in such
cases. In general, to minimize contamination from the
container and grinding medium, the container and grind-
ing medium should be harder/stronger than the powder
being milled. Nonetheless, contamination is difcult to
completely avoid in MA.
The problem of milling atmosphere is serious and has
been found to be a major cause of contamination. It has
been observed that if the container is not properly sealed,
the atmosphere surrounding the container, usually air,
leaks into the container and contaminates the powder.
This is particularly problematic for oxidation-sensitive
materials, such as pure metals. De Castro and Mitchell
studied contamination caused by high-energy milling with
a SPEX 8000 mixer/mill [73]. Aluminum with a purity of
99.9% or better and an initial particle size in the range
of 50100 m was milled in stainless steel and tungsten
carbide vials with milling media of the same composition
as the vial. A ball-to-powder ratio of 10:1 was used in all
cases. Ethanol was added as surfactant during milling and
subsequently removed. Figures 25 and 26 show contami-
nation as a function of milling time in stainless steel and
tungsten carbide, respectively. For the stainless steel vials,
Figure 25. Contamination of aluminum powders milled in stainless
steel vials.
contamination is reported as the sum of the atomic per-
centages of Fe, V, and Cr, as determined from X-ray u-
orescence (XRF). Vanadium and chromium are common
alloying elements in stainless steel. For milling in WC,
contamination is reported as percentage WC as deter-
mined by XRF. As can be seen, contamination increases
dramatically with milling time, especially in the WC vial.
The objective of current studies is to reduce the contam-
ination caused by the milling media/vials and at the same
time retain, or further reduce in size, the desired nanos-
tructure of the powders.
7. CHARACTERIZATION OF NANOPARTICLES
The powders obtained after MA must be characterized
for their size, shape, surface area, phase constitution,
and microstructural features. Additionally, one could also
characterize the transformation behavior of the mechan-
ically alloyed powders in annealing and other treatments.
The measurement of crystallite size and lattice strain in
mechanically alloyed powders is very important, since the
phase constitution and transformation appear to be crit-
ically dependent upon them. We attempt here to give a
very brief overview of the common techniques used to
determine both particle and crystallite size, in an attempt
to help differentiate between nanocrystals and nanopar-
ticles, both of which can result from mechanical attrition.
Figure 26. Contamination of aluminum powders milled in tungsten car-
bide vials.
14 De Castro and Mitchell
Those wishing to obtain more detailed information on
these well-established analytical techniques are directed
to their respective literature sources.
The crystalline size or grain size in nanoparticles can
be determined with several techniques that rely upon the
peak width in X-ray diffraction patterns. In the Hall
Williamson method [74, 75], crystalline size, D, is given
by
pcos 0 = 2: sin 0 +
0.94\
D
(3)
where p is the peak full width at half-maximum
(FWHM), 0 is the diffraction angle, \ is the wavelength
of the X-ray, and : is the effective strain associated with
mechanical alloying and ultrane grain size. The crys-
talline size D can be obtained by extrapolating sin 0 to
zero to eliminate the strain term.
The most widely used equation to determine grain size
from XRD data is the Scherrer equation,
1
o
=
K\
pcos 0
(4)
where K is a constant of order 1 dependent on parti-
cle shape, \ is the wavelength of the X-ray radiation, 0
is the diffraction angle, and p is the FWHM, after cor-
rections concerning instrument broadening. It should be
noted that for both the HallWilliamson and Scherrer
methods of determining grain size, one or several peaks
from the XRD pattern can be selected for analysis, and
the grain size reported for each, or as an average of mul-
tiple peak determinations. In either case, the grain size
determination from these methods gives a bulk-average
value.
The size and shape of powder particles may be deter-
mined accurately with direct methods of either scanning
electron microscopy (SEM) for relatively coarse powders
or transmission electron microscopy (TEM) for ne pow-
ders. In contrast to XRD, which gives a bulk-average
value for grain size, observations of particle size via
SEM/TEM give only values of an isolated sample area
and, as such, may not be representative of the entire sam-
ple, vis--vis particle size distribution. As outlined previ-
ously, the values of particle size should not be confused
with crystalline size, as determined from XRD data.
8. SUMMARY AND CONCLUSIONS
Mechanical alloying is a simple and useful processing
technique that is now being employed in the production
of nanocrystals and/or nanoparticles from all material
classes. Although a variety of mechanical alloying devices
exist, the high-energy ball mill is typically used to pro-
duce particles in the nanoscale size range. Particle size
reduction is effected over time in the high-energy ball
mill, as is a reduction in crystallite grain size, both of
which reach minimum values at extended milling times.
The distinction between nanoparticles and nanocrystals
in the current MA literature is not clear. Grain sizes, as
determined with standard X-ray diffraction techniques,
are typically reported as a function of milling time; how-
ever, the actual particle sizes that result from milling,
while assumed to be larger than the reported grain sizes
for polycrystalline materials, often go unreported. More-
over, what are often reported as nanocrystals may not be
such in the technical denition of the term (-10-nm grain
size), especially for polymeric materials, where grain
sizes in excess of 100 nm are still being obtained. Par-
ticle agglomeration, where nanoparticles stick together
because of attractive forces, is also a serious issue at long
milling times. Finally, contamination from milling media
(e.g., stainless steel vials and balls) is a serious problem
that has not yet been thoroughly investigated. Despite
these difculties, MA is more widely used than ever and
continues to be applied to the formation of nanoparti-
cles and nanocrystalline structures in an ever-increasing
variety of metals, ceramics, and polymers.
REFERENCES
1. M. K. Datta, S. K. Pabi, and B. S. Murty, Mater. Sci. Eng. A284, 219
(2000).
2. H. Gleiter, Prog. Mater. Sci. 33, 223 (1989).
3. C. B. Murray, S. Sun, H. Doyle and T. Bitley, MRS Bull. 985 (2001).
4. J. Karch, R. Birringer, and H. Gleiter, Nature 300, 556 (1987).
5. R. W. Siegel, H. Hahn, S. Ramasamy, L. Zongquan, L. Ting and
R. Gronsky, J. Phys. C5, 681 (1988).
6. C. C. Koch, Nanostruct. Mater. 2, 109 (1993).
7. C. C. Koch, Proceedings of Nanophases and Nanocrystalline Struc-
tures, TMS Symposium, 1992.
8. R. W. Siegel, in Materials Science and Technology (R. W. Cahn,
P. Haasen, and E. S. Kramer, Eds.), p. 583. VCH, Weinheim, 1991.
9. P. S. Gilman and J. S. Benjamin, Annu. Rev. Mater. Sci. 13, 279
(1983).
10. R. F. Singer, in Powder Metallurgy of Superalloys (G. H.
Gessinger, Ed.), pp. 213294. Butterworth, London, 1984.
11. R. Sundaresen and F. H. Froes, J. Metals 39, 213 (1987).
12. A. W. Weeber and H. Bakker, Physica, B 153, 93 (1988).
13. C. C. Koch, Annu. Rev. Mater. Sci. 19, 121 (1989).
14. G. B. Schaffer and P. G. McCormick, Mater. Forum 16, 91 (1992).
15. H. Bakker, G. F. Zhou, and H. Yang, Prog. Mater. Sci. 39, 159
(1995).
16. A. P. Smith, H. Ade, C. M. Balik, C. C. Koch, S. D. Smith and
R. J. Spontak, Macromolecules 33, 2595 (2000).
17. Z. He and T. H. Courtney, Mater. Sci. Eng., A 315, 166 (2001).
18. J. S. Benjamin, Sci. Am. 234, 40 (1976).
19. J. S. Benjamin, E. Artz, and L. Schultz, Eds. DMG informationge-
sellschaft: Oberursel, pp. 318 (1989).
20. J. S. Benjamin, Metal Powder Rep. 45, 122 (1990).
21. A. E. Ermakov, E. E. Yurchikov, and V. A. Barinov, Phys. Metal.
Metallorg. 52, 50 (1981).
22. C. C. Koch, O. B. Cavin, C. G. MaKamey and J. O. Scarbrough,
Appl. Phys. Lett. 43, 1017 (1983).
23. C. C. Koch, in Materials Science and Technology (R.W. Cahn,
Ed.) pp. 193245. VCH Verlagsgesellschaft GmbH, Weinheim,
Germany, 1991.
Nanoparticles from Mechanical Attrition 15
24. C. Suryanarayana, Bibliography on Mechanical Alloying and
Milling. Cambridge International Science Publishing, Cambridge,
U.K., 1995.
25. C. Suryanarayana, Metals Mater. 2, 195 (1996).
26. M. O. Lai and L. Lu, in Mechanical Alloying. Kluwer Academic,
Boston, MA, 1998.
27. B. S. Murty, Int. Mater. Rev. 43, 101 (1998).
28. E. Kuhn, Powder Metallurgy, ASM Handbook, Vol. 7, ASM
International: Materials Park, OH. pp. 5670. 1984.
29. E. Artz, G. Dehm, P. Gumbsch, O. Kraft and D. Weiss, Prog. Mater.
Sci. 46, 282 (2001).
30. Y. D. Kim, J. Y. Chung, J. Kim and H. Jeon, Mater. Sci. Eng., A
291, 17 (2000).
31. C. C. Harris, Trans. Soc. Mining Eng. 17 (1967).
32. I. Borner and J. Eckert, Mater. Sci. Eng., A 226228, 541 (1997).
33. J. Eckert, L. Shhultz, E. Helestern and K. Urban, J. Appl. Phys. 64,
3224 (1988).
34. C. C. Koch, D. Pathak, and K. Yamada, Mechanical Alloying for
Structural Applications (J. J. deBarbadillo, Ed.), pp. 205212. ASM
International: Materials Park, OH, 1993.
35. C. Suryanarayana, Intermetallics 3, 153 (1995).
36. J. Eckert and I. Borner, Mater. Sci. Eng., A 239240, 619 (1997).
37. M. Umemoto, Z. G. Lui, K. Masuyama and K. Rsuchiya, Scripta
Mater. 44, 1741 (2001).
38. T. Raghu, R. Sundaresan, P. Ramakrishnan and T. R. Mohan, Mater.
Sci. Eng., A 304306, 438 (2001).
39. Z. Xinkun, Z. Kunyu, C. Baochang, L. Qiushi, Z. Xinquin, C. Tieli
and S. Yuunsheng, Mater. Sci. Eng., C 16, 103 (2001).
40. R. M. Davis and C. C. Koch, Scr. Metall. 21, 305 (1987).
41. S. Zhang and S. C. Tam, J. Mater. Processing 67, 112 (1997).
42. S. Indris, D. Bork, and P. Heitjans, J. Mater. Synth. Processing 8, 245
(2000).
43. L.B., Kong, J. Ma, W. Zhu and O. K. Tan, Mater. Lett, 51, 108
(2001).
44. L. B. Kong, T. S. Zhang, and J. Ma, J. Mater. Res. 16, 1636 (2001).
45. L. B. Kong, J. Ma, H. T. Huang, W. Zhu and O. K. Tan, Mater. Lett.
50, 129 (2001).
46. L. Lu and M. O. Lai, Mechanical Alloying. Kluwer Academic
Publishers, Boston, MA, 1998.
47. J. S. Rawers, G. Seavens, P. Govier, C. Dogan and R. Doan, Metall.
Mater. Trans. A 27, 3126 (1996).
48. R. Angers, M. R. Krishnadev, R. Tremblay, J. F. Corriveau and
D. Dube, Mater. Sci. Eng., A 262, 9 (1999).
49. K. A. Khor and Y. Li, Mater. Sci. Eng., A 256, 271 (1998).
50. S. M. Zhu and K. Iwasaki, Mater. Sci. Eng., A 270, 170 (1999).
51. F. Y. C. Boey, Z. Yuan, and K. A. Khor, Mater. Sci. Eng., A 252,
276 (1998).
52. T. R. Malow and C. C. Koch, Mater. Sci. Forum 225227, 594
(1996).
53. O. N. Senkov, M. Cavusoglu, and F. H. Froes, Mater. Sci. Eng., A
300, 85 (2001).
54. S. Hwang and C. Nishimura, Scripta Mater. 44, 2457 (2001).
55. G. J. Fan, M. X. Quan, Z. Q. Hu, J. Eckert and L. Schultz, Scripta
Mater. 41, 1147 (1999).
56. L. Z. Zhou, J. T. Guo, and G. J. Fan, Mater. Sci. Eng., A 249, 103
(1998).
57. T. Takashi and Y. Hashimoto, Mater. Sci. Forum 8890, 175 (1992).
58. M. Zhu, W. H. Zhu, Y. Gao, X. Z. Che and J. H. Ahn, Mater. Sci.
Eng., A 286, 130 (2000).
59. J. W. Lee and Z. A. Munir, J. Am. Ceram. Soc. 84, 1209 (2001).
60. J. Pan and W. J. Shaw, Microstruct. Sci. 20, 351 (1993).
61. J. Pan and W. J. Shaw, Microstruct. Sci. 21, 95 (1994).
62. W. J. Shaw, J. Pan, and M. A. Growler, in Proceedings of the 2nd
Conference on Structural Applied Mechanics in Alloying, 1993.
ASM International, Materials Park, OH.
63. H. L. Castricum H. Yang, H. Bakker and J. K. Van Deursen, Mater.
Sci. Forum 235 (1997).
64. M. P. Farrel, R. G. Kander, and A. O. Aning, J. Mater. Synth. Proc.
4, 1996 (1996).
65. J. Font, J. Muntasell, and E. Cesari, Mater. Res. Bull. 34, 157 (1999).
66. T. Ishida, Mater. Sci. Lett. 13, 623 (1994).
67. C. M. Balik, C. Bai, and C. C. Koch, Mater. Res. Soc. Symp. Proc.
461, 39 (1997).
68. A. P. Smith, H. Ade, C. M. Balik, C. C. Koch, S. D. Smith and
R. J. Spontak, Macromolecules 33, 2595 (2000).
69. A. P. Smith, J. S. Shay, R. J. Spontak, C. M. Balik, H. Ade, S. D.
Smith and C. C. Koch, Polymer 42, 4453 (2001).
70. J. Font, J. Muntasell, and E. Cesari, Mater. Res. Bull. 34, 157 (1999).
71. J. Font, J. Muntasell, and E. Cesari, Mater. Res. Bull. 36, 1665
(2001).
72. C. Bay, R. J. Spontak, C. C. Koch, C. K. Shaw and C. M. Balik,
Polymer 41, 7147 (2000).
73. C. DeCastro and B. Mitchell, unpublished observations.
74. B. D. Cullity, Elements of X-ray Diffraction, p. 356 Addison-
Wesley, Reading, MA, 1978.
75. G. K. Williamson and W. H. Hall, Acta Metall. 1, 22 (1953).
76. S. M. Zhu, M. Tamura, K. Sakamoto and K. Iwasaki, Mater. Sci.
Eng., A 292, 83 (2000).

Vous aimerez peut-être aussi