You are on page 1of 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 the materials milled in mechanical attrition devices are


highly crystalline, such that the crystallite (grain) size
Unlike nanoparticles produced from “bottom-up” pro-
cesses such as self-assembly and templated synthesis, after milling is often between 1 and 10 nm in diame-
nanoparticles from mechanical attrition are produced by ter. Such materials are termed “nanocrystalline” [2]. The
a “top-down” process. Such nanoparticles are formed in sizes of the nanocrystals and the nanoparticles may or
a mechanical device, generically referred to as a “mill,” may not be the same. In some of the nanostructured
in which energy is imparted to a course-grained mate- materials literature, particularly that involving bottom-up
rial to effect a reduction in particle size. Under cer- processes, the term “nanocrystal” is reserved for crys-
tain conditions, the resulting particulate powders can talline particles with low concentrations of defects, such
exhibit nanostructural characteristics on at least two lev- as are found in single crystals, whereas “nanoparticles”
els. First, the particles themselves, which normally pos- are those nanoscale particles that contain gross internal
sess a distribution of sizes, can be “nanoparticles” if their grain boundaries, fractures, or internal disorder, whether
average characteristic dimension (diameter for spheri- the crystals they contain are nanocrystalline or not [3].
cal particles) is less than 100 nm [1]. Second, many of However, we will see that because of the large amount

Synthesis, Functionalization and Surface Treatment of Nanoparticles


ISBN: 1-58883-009-8 / $35.00 Edited by M.-I. Baraton
All rights of reproduction in any form reserved. 1 Copyright © 2002 by American Scientific Publishers
2 De Castro and Mitchell

of strain imparted to particles during the milling pro- Nickel Company (INCO) developed the process around
cess, it is virtually impossible to obtain defect-free crys- 1966. The technique was the result of an intense research
tals via mechanical attrition. As a result, we will adhere effort to produce nickel-based superalloys for gas turbine
to the more general definitions; that is, nanocrystals are applications. Benjamin has summarized the historical ori-
1–10 nm in diameter, and nanoparticles are less than gins of the process and the background work that led to
100 nm in diameter. Thus, it is possible to have both the development of MA [18–20]. To summarize, they pro-
nanocrystalline nanoparticles and coarser particles that duced a nickel-chromium-aluminum-titanium alloy first
contain nanocrystals. Both types of materials produced produced in a small high-speed shaker mill and later in a
by mechanical attrition will be addressed in this chapter. one-gallon stirred ball mill, starting the birth of MA as a
The importance of nanoparticles lies in their inher- method to produce oxide dispersion-strengthened (ODS)
ently large surface-to-volume ratio relative to that of alloys on an industrial scale. The process developed by
larger particles. These high surface areas can potentially Benjamin was initially referred to as “milling/mixing” and
improve catalytic processes and interfacially driven phe- later termed “mechanical alloying” by Ewan C. Mac-
nomena such as wetting and adhesion. Nanoparticles Queen, a patent attorney for INCO. The formation of
have the potential for use in structural and device appli- an amorphous phase by mechanical grinding of a Y-Co
cations in which enhanced mechanical and physical char- intermetallic compound in 1981 [21] and in the Ni-Nb
acteristics are required. As for the internal structure of system by ball milling of blended elemental mixtures
the nanoparticles, it has been found that nanocrystalline in 1983 [22] established MA as a potential nonequilib-
materials have comparative advantages over their micro- rium processing technique. Beginning in the mid-1990s, a
crystalline counterparts in hardness, fracture toughness, number of investigations were carried out to synthesize a
and low temperature ductility [4, 5]. As new methods variety of stable and meta-stable phases, including super-
for surface modification and postattrition processing of saturated solid solutions, crystalline and quasi-crystalline
nanoparticles are developed, the potential applications intermediate phases, and amorphous alloys [23–27]. Since
for them continue to grow. then, mechanical alloying has been applied to virtually all
The early work on the production of nanostructured material classes, including metals, ceramics, and polymers.
materials by mechanical attrition has been reviewed pre-
viously [6–15]. However, significant information contin-
ues to become available in this rapidly evolving field, 3. PRINCIPLES OF MILLING
particularly with regard to the range of material classes The fundamental principle of size reduction in mechan-
to which it is applied. We attempt here to provide an ical attrition devices lies in the energy imparted to
updated review of the subject, as mechanical attrition is the sample during impacts between the milling media.
used to form nanoparticles in new materials such as poly- The impact process is shown in Figure 1. This model
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
(flaking); agglomeration; solid-state blending (incomplete
alloying); modifying, changing, or altering properties of
a material (density, flowability, 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
Figure 1. Model of impact event at a time of maximum impacting force,
of homogeneous materials from blended elemental pow- showing the formation of a microcompact. Reprinted with permission
der mixtures. John Benjamin and his colleagues at the from [28], E. Kuhn, ASM Handbook, Vol. 7, Materials Park, OH, 1984.
Paul D. Merica Research Laboratory of the International © 1984, ASM.
Nanoparticles from Mechanical Attrition 3

Table 1. Selected Young’s moduli for isotropic solids at room


temperature.

Material E (×10−6 psi) E (10−10 N/m2 )

Aluminum 10 7
-Iron 30 21
Copper 16 11
Silicon 16 11
Tungsten 59 41
Titanium carbide (TiC) 45 31
Alumina (Al2 O3 ) 58 40
Magnesia 45 31
Hard rubber (ebonite) 06 04
Polystyrene (atactic) 04 03
Polyethylene (branched) 003 002

where c is the length of the crack, E 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 fineness 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:
Figure 2. Process of trapping an incremental volume of powder
between two balls in a randomly agitated charge of balls and pow- • Increasing resistance to fracture.
der. (a–c) Trapping and compaction of particles. (d) Agglomeration. (e)
Release of agglomerate by elastic energy. Reprinted with permission
• Increasing cohesion between particles, with decreas-
from [28], E. Kuhn, ASM Handbook, Vol. 7, Materials Park, OH, 1984. ing particle size causing agglomeration.
© 1984, ASM. • Excessive clearance between impacting surfaces.
• Coating of the grinding medium by fine particles that
represents the moment of collision, during which par- cushion the microbed of particles from impact.
ticles are trapped between two colliding balls within a • Surface roughness of the grinding medium.
space occupied by a dense cloud, dispersion, or mass • Bridging of large particles to protect smaller parti-
of powder particles [28]. Figure 2 shows the process of cles in the microbed.
trapping an incremental volume between two balls. Com- • Increasing apparent viscosity as particle size
paction begins with a powder mass that is characterized decreases.
by large spaces between particles compared with the par-
ticle size. The first stage of compaction starts with the Table 2. Selected surface energies of solid materials.
rearrangement and restacking of particles. Particles slide
past one another with a minimum of deformation and Material Surface energy,  (J/m2 )
fracture, producing some fine, irregularly shaped parti- Au 140
cles. The second stage of compaction involves elastic and Cu 1.43–1.70
plastic deformation of particles [29]. Cold welding may Ag 1.14–1.20
Ni 190
occur between particles in metallic systems during this
NaCl 030
stage. The third stage of compaction, involving particle Al2 O3 0905
fracture, results in further deformation and/or fragmen- MgO 1.00–1.20
tation of the particles. TiC 119
For brittle materials, particle fracture is well described LiF 034
CaF2 045
by Griffith theory. According to the theory, the stress, F ,
BaF2 028
at which crack propagation leading to catastrophic failure CaCO3 023
(fracture) occurs in the particle is approximated by Si 124
 Poly(tetrafluoroethylene) 00185
E Poly(ethylene) 0033
F ≈ (1) Poly(styrene) 0034
c
4 De Castro and Mitchell

Figure 5. SPEX 8000D dual mixer/mill.

They differ in their capacity, efficiency of milling, and


additional arrangements for heat transfer and particle
Figure 3. Change of average particle size as a function of milling time
for MA Fe-20at.%Co powders by (a) fixed r.p.m. operation and (b)
removal. A brief description of the different mills avail-
cyclic operation. Reprinted with permission from [30], Y. D. Kim et al., able for MA can be found below.
Mater. Sci. Eng. A 291, 17 (2000). © 2000, Elsevier Science.
4.1. SPEX Shaker Mills
• Decreasing internal friction of slurry as particle size Shaker mills such as SPEX (Fig. 5), which mill about
decreases. 10–20 g of powder at a time, are most commonly used
The grinding limit is clearly demonstrated in the studies for laboratory investigations and for alloy screening pur-
of Kim et al. [30]; after milling of Fe-Co powders for 30 h, poses. The common variety of the mill has one vial (see
the MA process reached a steady state where the parti- Fig. 6), containing the sample and the grinding media,
cles have become homogenized in size and shape (Fig. 3). which is secured in the clamp and swung energetically
This work also clearly distinguishes between particle size, back and forth several thousands times a minute. The
which approaches a limit of 10 m with a planetary ball back-and-forth shaking motion is combined with lateral
mill, and crystallite size, which approaches 10 nm (Fig. 4). movements of the ends of the vial. With each swing of
Figure 4 also illustrates the strain buildup in the particles the vial the milling media, typically hard, spherical objects
as milling time increases and grain sizes decrease. 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
4. ATTRITION DEVICES 1200 rpm) of the clamp motion, the ball velocities are
Different types of milling equipment are available high (on the order of 5 m/s), and consequently the force
for mechanical alloying and nanoparticle formation. of the ball’s 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 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 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 Figure 7. SEM of Bi4 Ti3 O12 milled for different times: (a) 3, (b) 9, (c)
both act on the vial contents, consisting of material to be 15, and (d) 20 h. Reprinted with permission from [43], L. B. Kong et al.,
Mater. Lett. 51, 108 (2001). © 2001, Elsevier Science.
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 reduction appears to take place by interparticle collisions
causes the grinding balls to run down the inside of the and ball sliding. A motor rotates the impellers, which in
vial—the friction effect—followed by the material being turn agitate the steel balls in the drum.
ground. Grinding balls lift off and travel freely through Attritors (Fig. 8) are the mills in which large quanti-
the inner chamber of the vial and collide against the ties of powder (from about 0.5 to 40 kg) can be milled
opposing inside wall—the impact effect. at a time. Attritors of different sizes and capacities are
Even though the disk and the vial rotation speeds could available. The grinding tanks or containers are avail-
not be independently controlled in the earlier versions of able in stainless steel or stainless steel coated with alu-
this device, it is possible to do so in modern versions. In a mina, silicon carbide, silicon nitride, zirconia, rubber,
single mill there can be either two (Pulverisette 5 or 7) or or polyurethane. A variety of grinding media are also
four (Pulverisette 5) milling stations. Recently, a single- available: glass, flint stones, stealite ceramic, mullite, sil-
version mill was also developed (Pulverisette 6). Grinding icon carbide, silicon nitride, Sialon, alumina, zirconium
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-filled 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; Figure 8. Attrition ball mill. Reprinted with permission from [28],
between the balls and the container wall; and between E. Kuhn, ASM Handbook, Vol. 7, Materials Park, OH, 1984. © 1984,
the balls, the agitator shaft, and the impellers. Some size 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 final 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 Figure 10. Typical size capabilities of common classes of size reduc-
tion equipment. Reprinted with permission from [28], E. Kuhn, ASM
shaker mill provides the largest input and therefore leads
Handbook, Vol. 7, Materials Park, OH, 1984. © 1984, ASM.
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
5. MILLED MATERIALS
during the collision. After 32 h of milling the grain size
achieved was 40 nm at 90 rpm, 31 nm at 180 rpm, and Though traditionally limited to metallic materials, in par-
20 nm at 360 rpm (Fig. 9). ticular FCC metals, mechanical alloying is now being uti-
The different classifications of mills and the typical lized to form nanoparticles and nanocrystalline particles
range of particle sizes they produce are summarized in from a variety of materials. We report here on recent
Figure 10. Notice that vibratory mills are the typical class research results that employ MA as the principal pro-
of mills used to produce nanoparticles. Figure 11 shows cessing method and the primary benefits it provides for
how the different mill types can be used to attrit materials different types of 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, Figure 11. Applicability of various classifications of comminution
Mater. Sci. Eng. A 226–228, 541 (1997). © 1997, Elsevier Science. machines to different material types.
Nanoparticles from Mechanical Attrition 7

5.1. Solid-State Amorphization phase had been formed [35]. Lattice measurements sug-
gested that this FCC crystalline phase is due to the
A solid alloy with a noncrystalline atomic arrangement
increased nitrogen contamination of the milled powder,
is called an amorphous (metallic) alloy. The synthesis of
and the crystalline phase has been identified as TiN.
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 5.2. Metals, Alloys, and Intermetallics
area. Amorphous phases have been synthesized by MA Metals are routinely milled to the nanoparticle size range,
from blended elemental powder mixtures, pre-alloyed with typical crystallite sizes in the 1–10-nm range. The
powders and/or intermetallics, mixtures of intermetallics, average grain size of a NiAl intermetallic compound was
or mixtures of intermetallics and elemental powders. as low as 5 nm after 100 h of milling [36] and as low as
The effect of process variables on amorphization 4.7 nm for Fe-C alloys milled in a horizontal ball milling
behavior has been studied in several alloy systems. The machine [37].
most important variables that have been studied are Kim et al. [30] studied the effect of fixed and cyclic
milling energy and milling temperature. Increased milling operation on the formation of nanocrystalline Fe-Co
energy, which is achieved by higher ball-to-powder weight alloy powders produced by mechanical alloying in a plan-
ratio and increased speed of rotation, is expected to intro- etary ball mill. The first method was conventional milling
duce more strain and increase the defect concentration with fixed velocity (1300 rpm). The second method
in the powder, thereby leading more readily to amor- involved an operation cycle characterized by a time inter-
phization. On the other hand, higher milling energies can val of 4 min at 1300 rpm followed by 1 min of oper-
also produce more heat, resulting in crystallization of the ation at 900 rpm. In both cases, the milling time was
amorphous phase. A balance between these two effects varied from 1 to 100 h. It is generally known that the
will determine the nature of the final product phase. MA process reaches a steady state where the particles
Eckert et al. [33] reported that in Ni-Zr system milled in have homogeneous size and shape. The steady state was
a planetary ball mill at a low intensity did not produce any achieved after 30 h of milling for fixed operation and
amorphous phase, because of the lack of energy. How- 15 h of milling for a cyclic operation. The advantage of
ever, when the intensity was increased, amorphous phase the cyclic operation is due to the fact that the periodical
formation was observed in a wide composition range of changes of the milling velocity in cyclic operation break
30–83% Ni. At even higher intensities, amorphous phase the balance of deformation, fracture, and welding in the
formation was observed between 66 and 75% Ni. These process and maximize the effect of fracture. The steady-
observations suggest that with increasing milling energy, state crystallite grain size obtained was in the range of
the heat generated is also high, which crystallizes the 10–15 nm. Interestingly, during the first hour of milling
amorphous phases. It appears that the maximum amor- the grain size of Fe-Co increased and then decreased to
phization range is observed at intermediate values of the final 10–15 nm after achieving steady state.
milling intensity; too low an intensity does not provide Raghu et al. [38] studied the differences between
enough energy to amorphize, while at very high intensi- nanocrystalline copper-tungsten alloys milled in argon
ties, the amorphous phase formed would crystallize. and air. They determined that the particle size initially
There have been conflicting results on the effect of decreases and then increases in powders milled in air
milling temperature on the nature of the phase formed. (Fig. 12). The crystallite sizes of copper and tungsten
Koch et al. [34] summarized the results of varying the decrease continuously with milling time in all milling
temperature of the mill on the kinetics of amorphization atmospheres (Fig. 13). Crystallite size levels off at a value
in intermetallics. They concluded that generally a lower of 10 nm for copper and 15 nm for tungsten after 8 h of
milling temperature accelerated the amorphization pro- milling. It was concluded that whereas fracture is dom-
cess. Since a nanostructured material can easily be pro- inant in milling in argon, adequate welding for the MA
duced at lower milling temperatures, the increased grain process appears to be available during milling in air.
boundary area drives the crystal-to-amorphous transfor- Eckert and Borner [36] studied nanostructure forma-
mation. For example, the time required for amorphiza- tion of ball-milled NiAl intermetallic compounds in a
tion in NiTi was 2 h at −190  C, 13 h at 60  C, and 18 h planetary ball mill with hardened steel balls and vial at a
at 220  C [34]. rotation speed of 180 rpm. Figure 14 illustrates that the
Mechanical alloying introduces contamination into the average grain size and the rms strain scale with composi-
milled powder and thus alters the constitution and stabil- tion of the material. Whereas the grain size is reduced to
ity of powder products. Generally, the presence of addi- only 12 nm for Ni46 Al54 , it decreased to about 5 nm for
tional elements favors amorphization. Contamination has Ni-rich compounds. Blending Ni46 Al54 with elemental Ni
also been found to affect the stability of the amorphous to give an overall composition of Ni50 Al50 and subsequent
phases formed. Formation of a crystalline phase has been mechanical alloying reduced the grain size from 12 nm
reported on milling of Ti-Al powders after an amorphous after 100 h to about 9 nm after 300 h of total milling time
8 De Castro and Mitchell

Figure 14. Average grain size (filled symbols) and atomic-level strain
Figure 12. Particle size of Cu-5vol%W powders milled in different (open symbols) for Nix Al100−x powders after 100 h of milling as a func-
atmospheres. Reprinted with permission from [38], T. Raghu et al., tion of composition. Reprinted with permission from [36], J. Eckert and
Mater. Sci. Eng. A 304–306, 438 (2001). © 2001, Elsevier Science. I. Borner, Mater. Sci. Eng. A 239–240, 619 (1997). © 1997, Elsevier
Science.

(Fig. 15). On the other hand, the grain size of 100-h ball-
milled Ni60 Al40 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 15. Average grain size obtained for Ni56 Al54 and Ni60 Al40 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 Ni50 Al50 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 239–240, 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 10–15 30 [30]


Fe Vibratory mill 20 4 [32]
NiAl Vibratory mill 12 100 [36]
Ni silicides Vibratory mill 10–17 30 [1]
Fe-C Horizontal ball 4.7 500 [37]
Figure 13. Average grain size of (a) copper and (b) tungsten powders
mill
milled in argon and air. Reprinted with permission from [38], T. Raghu
Fe3 Al Vibratory mill 12.6 100 [76]
et al., Mater. Sci. Eng. A 304–306, 438 (2001). © 2001, Eslevier Science.
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
Figure 17. Lattice parameter as a function of MA time for 70wt%
the milling time reached 115 min, the temperature of the TiC + 30wt% TiN as determined from XRD patterns. Reprinted with
vial increased abruptly and reached its maximum (from permission from [41], S. Zhang and S. C. Tam, J. Mater. Processing 67,
319 to 332 K). The temperature peak was associated with 12, (1997). © 1997, Elsevier Science.
the exothermic reaction of Ti and C. After the peak, the
vial temperature decreased gradually during subsequent lattice parameter with MA time is almost the same for
milling. the two compositions; this shows that in MA, the reac-
Traditionally, the raw materials used in mechanical tion rate is independent of the “concentration” of the
alloying include at least one ductile material used as a “reactant” as long as the latter is not depleted, which
host or binder for the other ingredients. However, Davis is different from a chemical reaction, where the reac-
and Koch [40] mechanically alloyed a brittle Si-Ge sys- tion rate is proportional to the concentration of reac-
tem to obtain a solid solution Si(Ge) and observed a lat- tants. The reduction of crystallite size was very rapid in
tice change with the increased duration of MA. Recently, the beginning, and after 40 h it was in the range of 20–
Zhang and Tam [41] studied the mechanical alloying of 30 nm (Fig. 18). In the studies of Indris et al. [42], the
an all-ceramic-phase component, TiC + TiN. Two com- system 1 − x
Li2 O:xB2 O3 reacts either at high temper-
positions were explored, 70% TiC + 30% TiN and 50% atures or because of mechanochemical treatment. Indris
TiC + 50% TiN (by weight). MA was conducted for up determined that the reaction product is not determined
to 80 h in a Fritsch planetary ball mill at a ball-to-powder by the overall composition of the mixture, but by the con-
ratio of 20:1. Figure 17 shows a plot of the lattice param- ditions at the particle interfaces.
eter as a function of milling time for 70% TiC + 30% Indris et al. [42] also studied the oxide ceramics Li2 O,
TiN milled with steel balls. A similar plot was obtained LiNbO3 , LiBO2 , B2 O3 , and TiO3 as monophase materi-
for 50% TiC + 50% TiN milled with WC balls. The two als. Nanostructures of these oxides were prepared in a
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 Figure 18. Reduction in average grain size as a function of milling time
at 115 min is attributed to formation of TiC. Reprinted with permission for 50wt% TiC + 50wt% TiN. Reprinted with permission from [41],
from [39], Z. Xinkun et al., Mater. Sci. Eng. C 16, 103 (2001). © 2001, S. Zhang and S. C. Tam, J. Mater. Processing 67, 112, (1997). © 1997,
Elsevier Science. Elsevier Science.
10 De Castro and Mitchell

the matrix. Finer size reinforcements produce higher spe-


cific mechanical properties. However, they also encounter
significant agglomeration problems for that size range,
resulting in poor dispersion and subsequently less than
optimal mechanical properties.
The main problem with the fine-grained structure is
its instability at high temperatures. Because of the large
excess of free energy, significant grain growth has been
observed in several nanocrystalline materials [52, 53].
Senkov et al. [53] studied grain growth of TiAl3 /Ti5 Si3
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
Figure 19. Average grain size of Li2 O, LiNbO3 , and B2 O3 powders for to 540 nm when the HIP temperature was increased from
various milling times. Reprinted with permission from [42], S. Indris 850  C to 1100  C.
et al., J. Mater. Synth. Processing 8, 245 (2000). © 2000, Plenum. Hwang and Nishimura [54] demonstrated that mag-
nesium matrix nanocomposites could be mechanochem-
high-energy ball mill Spex 8000 with alumina milling vials ically synthesized by milling elemental precursors, of
with a ball-to-powder ratio of 2:1 to produce at least 2 g which two of the elemental powders (normally, a
of powder. Figure 19 shows the average grain size for metal and an element such as C) form the secondary
LiNbO3 , Li2 O, and B2 O3 . Although the kinetics of grain particles. Nanocomposites have also been successfully
size reduction is different, all samples show a similar sat- mechanochemically synthesized by Fan et al. [55] with
uration value for a final grain size of about 23 nm. Nb-Si-Ti-C mixtures, Zhou et al. [56] with Ni-Al-Ti-C mix-
Kong et al. [43] prepared Bi4 Ti3 O12 ceramics by com- tures, and Takahashi and Hashimoto [57] in the Cu-M-C
bining Bi2 O3 and TiO2 powders. Although the particle system (where M = Zr, Hf, Nb, Ta, or Ti). Hwang and
size was reduced to 100–200 nm after 3 h, no reaction Nishimura [54] used a mixture of Mg, Ti, and C powder
between the powders took place. After 9 h of milling, to synthesize nanometer-sized TiC particles embedded
peaks for Bi4 Ti3 O12 appeared in the XRD pattern, indi- in a nanocrystalline Mg matrix by mechanochemical
cating reaction between the starting powders. After 15 h, reaction. Powder compositions corresponding to Mg-
the reaction was completed, with a further reduction in xTiC (where x = 5–15 volume %) were milled. The
particle size. Kong et al. also studied lead zirconium XRD pattern shows the development of the Mg-15%TiC
titanate (PZT) ceramics [44, 45] because of its excel- nanocomposite during milling (Fig. 20). After 3 h of
lent dielectric, piezoelectric, and electro-optic proper- milling, the sample still contained the three forms of
ties. PZT powders were formed in a planetary ball mill elemental powders, but with further milling (6 h), Ti
from PbO, ZrO2 , and TiO2 . The grain size from XRD peaks were reduced and TiC peaks appeared. After 24 h
ranged between 35 nm (4 h of milling) and 26 nm (24 h of milling, the powder showed only the Mg and TiC
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 [46–51]. The mechanically alloyed pow-
ders have attributes such as fine powder size and flexi-
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
Figure 20. Evolution of Mg-Ti-C nanocomposite during milling and
properties are influenced significantly by the quantity and after heat treatment. Reprinted with permission from [54], S. Hwang
the distribution mode of the reinforcements and by the and C. Nishimura, Scripta Mater. 44, 2457 (2001). © 2001 Elsevier
nature of the interfaces between the reinforcements and Science.
Nanoparticles from Mechanical Attrition 11

Figure 21. Average grain size of MmM5 phase in MmM5 -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 Figure 22. Average grain size and strain for Ti as a function of milling
grains of as-milled samples were approximately 5.6 and time, for the Williamson–Hall (WH) and Halder–Wagner (HW) meth-
ods. Reprinted with permission from [59], J. W. Lee and Z. A. Munir,
43.1 nm, respectively, and those of heat-treated samples
J. Am. Ceram. Soc. 84, 1209 (2001). © 2001, The American Ceramic
were approximately 7.1 and 62.3 nm, respectively. Mg- Society.
Ti-C nanocomposites exhibited remarkably high ductility
while retaining compressive strength similar to that of time. Figure 22 also shows that the minimum crystal-
Mg-TiC nanocomposite. lite size of Ti that can be obtained before the onset of
Zhu et al. [58] studied the effect of Mg content a reaction in the mill is in the range of 16.7–33.3 nm.
on the microstructure of mechanical alloyed Figure 23a shows the results of crystallite size analyses
MmNi35 (CoAlMn)15 -Mg (abbreviated as MmM5 -Mg). on milled powders (10 h) that were subsequently reacted
Figure 21 shows that the grain size of the MmM5 phase in the spark plasma synthesis (SPS) apparatus for 1 or
increases with Mg content in the composite. The authors 12 min, and Figure 23b shows the results on powders
explained this observation by the consideration that Mg that had reacted during 12 h of milling and subsequently
absorbed part of the energy imparted to the MmM5 reacted in the SPS for 5 to 12 min. It is interesting to
phase. MmM5 particles were gradually embedded in note that the crystallite size of the product phases after 1
Mg as milling proceeded. The larger the amount of min of SPS treatment is roughly the same whether they
Mg in the composite, the less energy was consumed in are formed in the SPS or previously during milling. How-
refining the grain size of MmM5 and thus the grain size ever, when the treatment is longer (12 min), the size of
of MmM5 increased with increasing Mg content. those formed in the SPS is about one-half that of those
Lee and Munir [59] studied the synthesis of dense formed during milling.
TiB2 -TiN nanocrystalline composites through mechanical
and field activation. Powder mixtures were mechanically 5.5. Polymers
activated through ball milling. The powders were blended
in a stoichiometric ratio according to the reaction Unlike the materials previously described here, there
is little work on nanoparticle polymers by mechanical
Ti + 05B + 05BN → 05TiB2 + 05TiN alloying. We include here recent progress toward the
formation of polymer nanoparticles and nanocrystalline
They were mixed in a Shaker mixer (Glenmill) for 24 h polymer particles.
and ball milled in a planetary mill. Figure 22 shows the High-energy milling of polymeric materials subjects
refinement of the grain size and the increase in strain as the blend components to a complex deformation field
a function of milling times for Ti. The small grain size of in which shear, multiaxial extension, fracture, and cold
unmilled Ti powder (71–111 nm) may be related to the welding proceed concurrently. This technique has been
particular preparation method of this powder, namely, successfully used to prepare blends of thermoplastics
that it was made from sponge Ti. As shown in Figure 22, [60–65] and to cause allotropism in semicrystalline poly-
the strain value changed very little in early stages of mers [66, 67]. Smith et al. have demonstrated that
milling, but after 2 h of milling, the strain increased nanostructured polymer blends composed of poly(methyl
markedly. Within the first 6 h of milling, the decrease methacrylate) (PMMA) and either poly(ethylene-alt-
in crystallite size for Ti was relatively linear with milling propylene) (PEP) or polyisopropene (PI) can be
12 De Castro and Mitchell

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 [70–72]. 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 difficulty in removing
the polymer from the solvent if the solution method is
used. These authors also studied the partial amorphiza-
Figure 23. Average grain size of TiN and TiB2 formed in spark plasma
sintering (SPS). (a) Reactant powders milled for 10 h (no product for-
tion of semicrystalline poly(ether-ether ketone) (PEEK)
mation). (b) Reactant powders milled for 12 h (with powder formation). by means of a ball-milling technique. The milling device
Reprinted with permission from [59], J. W. Lee and Z. A. Munir, J. Am. used was a Fritsch (Pulverisette 6) centrifugal ball mill
Ceram. Soc. 84, 1209 (2001). © 2001, The American Ceramic Society. working with a rotation speed of about 500 rpm. The
sample mass milled was 2 g, with a product-to-ball ratio
produced by cryogenic mechanical alloying [16]. In cryo- of 1:30. PEEK samples of 10 mg were analyzed by clas-
genic mechanic alloying the vial is sealed under argon sical differential scanning calorimetry. The analysis of
(<10 ppm O2 ) and placed in a custom-designed nylon XRD results revealed that ball milling leads to a partial
sleeve that allows peripheral circulation of liquid nitro- amorphization of PEEK (Fig. 24), therefore concluding
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 0.2–3 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
flows around and delineates the PMMA particles, which
Figure 24. Room-temperature XRD patterns for PEEK samples. (a)
decrease in size with increasing milling time. After 1 h of Unmilled. (b) After 24 h of milling. Reprinted with permission from
cryomilling, the PMMA particles measure less that 2 m [71], J. Font et al., Mater. Res. Bull., 36, 1665 (2001). © 2001, Elsevier
across and subsequently decrease to less than 300 nm 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, Figure 25. Contamination of aluminum powders milled in stainless
• vials and grinding media, steel vials.
• milling atmosphere, and
• control agents added to the powders. contamination is reported as the sum of the atomic per-
centages of Fe, V, and Cr, as determined from X-ray flu-
During MA the powder particles become trapped
orescence (XRF). Vanadium and chromium are common
between the grinding medium and undergo severe plastic
alloying elements in stainless steel. For milling in WC,
deformation; fresh surfaces are created because of the
contamination is reported as percentage WC as deter-
fracture of the powder particles. Collisions occur between
mined by XRF. As can be seen, contamination increases
the grinding medium and the vial and among the grinding
dramatically with milling time, especially in the WC vial.
balls. All of these effects can cause deterioration of the
The objective of current studies is to reduce the contam-
grinding medium and vial, resulting in the incorporation
ination caused by the milling media/vials and at the same
of these impurities into the powder. The extent of con-
time retain, or further reduce in size, the desired nanos-
tamination increases with increasing milling energy and tructure of the powders.
milling time.
Various attempts have been made to minimize powder
contamination during MA. One way of minimizing the 7. CHARACTERIZATION OF NANOPARTICLES
contamination from the grinding medium and the con- The powders obtained after MA must be characterized
tainer is to use a material for the container and grinding for their size, shape, surface area, phase constitution,
medium that is the same as the powder being milled. and microstructural features. Additionally, one could also
For example, one could use copper milling balls and a characterize the transformation behavior of the mechan-
copper vial for milling copper and copper alloy powders. ically alloyed powders in annealing and other treatments.
However, milling effectiveness may be limited in such The measurement of crystallite size and lattice strain in
cases. In general, to minimize contamination from the mechanically alloyed powders is very important, since the
container and grinding medium, the container and grind- phase constitution and transformation appear to be crit-
ing medium should be harder/stronger than the powder ically dependent upon them. We attempt here to give a
being milled. Nonetheless, contamination is difficult to very brief overview of the common techniques used to
completely avoid in MA. determine both particle and crystallite size, in an attempt
The problem of milling atmosphere is serious and has to help differentiate between nanocrystals and nanopar-
been found to be a major cause of contamination. It has ticles, both of which can result from mechanical attrition.
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 50–100 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 Figure 26. Contamination of aluminum powders milled in tungsten car-
tungsten carbide, respectively. For the stainless steel vials, bide vials.
14 De Castro and Mitchell

Those wishing to obtain more detailed information on which reach minimum values at extended milling times.
these well-established analytical techniques are directed The distinction between nanoparticles and nanocrystals
to their respective literature sources. in the current MA literature is not clear. Grain sizes, as
The crystalline size or grain size in nanoparticles can determined with standard X-ray diffraction techniques,
be determined with several techniques that rely upon the are typically reported as a function of milling time; how-
peak width in X-ray diffraction patterns. In the Hall– ever, the actual particle sizes that result from milling,
Williamson method [74, 75], crystalline size, D, is given while assumed to be larger than the reported grain sizes
by for polycrystalline materials, often go unreported. More-
over, what are often reported as nanocrystals may not be
094 such in the technical definition of the term (<10-nm grain
cos  = 2 sin  + (3)
D size), especially for polymeric materials, where grain
where is the peak full width at half-maximum sizes in excess of 100 nm are still being obtained. Par-
(FWHM),  is the diffraction angle,  is the wavelength ticle agglomeration, where nanoparticles stick together
of the X-ray, and  is the effective strain associated with because of attractive forces, is also a serious issue at long
mechanical alloying and ultrafine grain size. The crys- milling times. Finally, contamination from milling media
talline size D can be obtained by extrapolating sin  to (e.g., stainless steel vials and balls) is a serious problem
zero to eliminate the strain term. that has not yet been thoroughly investigated. Despite
The most widely used equation to determine grain size these difficulties, MA is more widely used than ever and
from XRD data is the Scherrer equation, continues to be applied to the formation of nanoparti-
cles and nanocrystalline structures in an ever-increasing
K variety of metals, ceramics, and polymers.
Lo = (4)
cos 

where K is a constant of order 1 dependent on parti-


cle shape,  is the wavelength of the X-ray radiation,  REFERENCES
is the diffraction angle, and is the FWHM, after cor-
1. M. K. Datta, S. K. Pabi, and B. S. Murty, Mater. Sci. Eng. A284, 219
rections concerning instrument broadening. It should be (2000).
noted that for both the Hall–Williamson and Scherrer 2. H. Gleiter, Prog. Mater. Sci. 33, 223 (1989).
methods of determining grain size, one or several peaks 3. C. B. Murray, S. Sun, H. Doyle and T. Bitley, MRS Bull. 985 (2001).
from the XRD pattern can be selected for analysis, and 4. J. Karch, R. Birringer, and H. Gleiter, Nature 300, 556 (1987).
the grain size reported for each, or as an average of mul- 5. R. W. Siegel, H. Hahn, S. Ramasamy, L. Zongquan, L. Ting and
R. Gronsky, J. Phys. C5, 681 (1988).
tiple peak determinations. In either case, the grain size 6. C. C. Koch, Nanostruct. Mater. 2, 109 (1993).
determination from these methods gives a bulk-average 7. C. C. Koch, “Proceedings of Nanophases and Nanocrystalline Struc-
value. tures,” TMS Symposium, 1992.
The size and shape of powder particles may be deter- 8. R. W. Siegel, in “Materials Science and Technology” (R. W. Cahn,
mined accurately with direct methods of either scanning 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
electron microscopy (SEM) for relatively coarse powders (1983).
or transmission electron microscopy (TEM) for fine pow- 10. R. F. Singer, in “Powder Metallurgy of Superalloys” (G. H.
ders. In contrast to XRD, which gives a bulk-average Gessinger, Ed.), pp. 213–294. Butterworth, London, 1984.
value for grain size, observations of particle size via 11. R. Sundaresen and F. H. Froes, J. Metals 39, 213 (1987).
SEM/TEM give only values of an isolated sample area 12. A. W. Weeber and H. Bakker, Physica, B 153, 93 (1988).
13. C. C. Koch, Annu. Rev. Mater. Sci. 19, 121 (1989).
and, as such, may not be representative of the entire sam-
14. G. B. Schaffer and P. G. McCormick, Mater. Forum 16, 91 (1992).
ple, vis-à-vis particle size distribution. As outlined previ- 15. H. Bakker, G. F. Zhou, and H. Yang, Prog. Mater. Sci. 39, 159
ously, the values of particle size should not be confused (1995).
with crystalline size, as determined from XRD data. 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).
8. SUMMARY AND CONCLUSIONS 18. J. S. Benjamin, Sci. Am. 234, 40 (1976).
19. J. S. Benjamin, E. Artz, and L. Schultz, Eds. DMG informationge-
Mechanical alloying is a simple and useful processing sellschaft: Oberursel, pp. 3–18 (1989).
technique that is now being employed in the production 20. J. S. Benjamin, Metal Powder Rep. 45, 122 (1990).
of nanocrystals and/or nanoparticles from all material 21. A. E. Ermakov, E. E. Yurchikov, and V. A. Barinov, Phys. Metal.
classes. Although a variety of mechanical alloying devices Metallorg. 52, 50 (1981).
22. C. C. Koch, O. B. Cavin, C. G. MaKamey and J. O. Scarbrough,
exist, the high-energy ball mill is typically used to pro-
Appl. Phys. Lett. 43, 1017 (1983).
duce particles in the nanoscale size range. Particle size 23. C. C. Koch, in “Materials Science and Technology” (R.W. Cahn,
reduction is effected over time in the high-energy ball Ed.) pp. 193–245. VCH Verlagsgesellschaft GmbH, Weinheim,
mill, as is a reduction in crystallite grain size, both of Germany, 1991.
Nanoparticles from Mechanical Attrition 15

24. C. Suryanarayana, “Bibliography on Mechanical Alloying and 50. S. M. Zhu and K. Iwasaki, Mater. Sci. Eng., A 270, 170 (1999).
Milling.” Cambridge International Science Publishing, Cambridge, 51. F. Y. C. Boey, Z. Yuan, and K. A. Khor, Mater. Sci. Eng., A 252,
U.K., 1995. 276 (1998).
25. C. Suryanarayana, Metals Mater. 2, 195 (1996). 52. T. R. Malow and C. C. Koch, Mater. Sci. Forum 225–227, 594
26. M. O. Lai and L. Lu, in “Mechanical Alloying.” Kluwer Academic, (1996).
Boston, MA, 1998. 53. O. N. Senkov, M. Cavusoglu, and F. H. Froes, Mater. Sci. Eng., A
27. B. S. Murty, Int. Mater. Rev. 43, 101 (1998). 300, 85 (2001).
28. E. Kuhn, “Powder Metallurgy,” ASM Handbook, Vol. 7, ASM 54. S. Hwang and C. Nishimura, Scripta Mater. 44, 2457 (2001).
International: Materials Park, OH. pp. 56–70. 1984. 55. G. J. Fan, M. X. Quan, Z. Q. Hu, J. Eckert and L. Schultz, Scripta
29. E. Artz, G. Dehm, P. Gumbsch, O. Kraft and D. Weiss, Prog. Mater. Mater. 41, 1147 (1999).
Sci. 46, 282 (2001). 56. L. Z. Zhou, J. T. Guo, and G. J. Fan, Mater. Sci. Eng., A 249, 103
30. Y. D. Kim, J. Y. Chung, J. Kim and H. Jeon, Mater. Sci. Eng., A (1998).
291, 17 (2000). 57. T. Takashi and Y. Hashimoto, Mater. Sci. Forum 88–90, 175 (1992).
31. C. C. Harris, Trans. Soc. Mining Eng. 17 (1967). 58. M. Zhu, W. H. Zhu, Y. Gao, X. Z. Che and J. H. Ahn, Mater. Sci.
32. I. Borner and J. Eckert, Mater. Sci. Eng., A 226–228, 541 (1997). Eng., A 286, 130 (2000).
33. J. Eckert, L. Shhultz, E. Helestern and K. Urban, J. Appl. Phys. 64, 59. J. W. Lee and Z. A. Munir, J. Am. Ceram. Soc. 84, 1209 (2001).
3224 (1988). 60. J. Pan and W. J. Shaw, Microstruct. Sci. 20, 351 (1993).
34. C. C. Koch, D. Pathak, and K. Yamada, Mechanical Alloying for 61. J. Pan and W. J. Shaw, Microstruct. Sci. 21, 95 (1994).
Structural Applications (J. J. deBarbadillo, Ed.), pp. 205–212. ASM 62. W. J. Shaw, J. Pan, and M. A. Growler, in “Proceedings of the 2nd
International: Materials Park, OH, 1993. Conference on Structural Applied Mechanics in Alloying,” 1993.
35. C. Suryanarayana, Intermetallics 3, 153 (1995). ASM International, Materials Park, OH.
36. J. Eckert and I. Borner, Mater. Sci. Eng., A 239–240, 619 (1997). 63. H. L. Castricum H. Yang, H. Bakker and J. K. Van Deursen, Mater.
37. M. Umemoto, Z. G. Lui, K. Masuyama and K. Rsuchiya, Scripta Sci. Forum 235 (1997).
Mater. 44, 1741 (2001). 64. M. P. Farrel, R. G. Kander, and A. O. Aning, J. Mater. Synth. Proc.
38. T. Raghu, R. Sundaresan, P. Ramakrishnan and T. R. Mohan, Mater. 4, 1996 (1996).
Sci. Eng., A 304–306, 438 (2001). 65. J. Font, J. Muntasell, and E. Cesari, Mater. Res. Bull. 34, 157 (1999).
39. Z. Xinkun, Z. Kunyu, C. Baochang, L. Qiushi, Z. Xinquin, C. Tieli 66. T. Ishida, Mater. Sci. Lett. 13, 623 (1994).
and S. Yuunsheng, Mater. Sci. Eng., C 16, 103 (2001). 67. C. M. Balik, C. Bai, and C. C. Koch, Mater. Res. Soc. Symp. Proc.
40. R. M. Davis and C. C. Koch, Scr. Metall. 21, 305 (1987). 461, 39 (1997).
41. S. Zhang and S. C. Tam, J. Mater. Processing 67, 112 (1997). 68. A. P. Smith, H. Ade, C. M. Balik, C. C. Koch, S. D. Smith and
42. S. Indris, D. Bork, and P. Heitjans, J. Mater. Synth. Processing 8, 245 R. J. Spontak, Macromolecules 33, 2595 (2000).
(2000). 69. A. P. Smith, J. S. Shay, R. J. Spontak, C. M. Balik, H. Ade, S. D.
43. L.B., Kong, J. Ma, W. Zhu and O. K. Tan, Mater. Lett, 51, 108 Smith and C. C. Koch, Polymer 42, 4453 (2001).
(2001). 70. J. Font, J. Muntasell, and E. Cesari, Mater. Res. Bull. 34, 157 (1999).
44. L. B. Kong, T. S. Zhang, and J. Ma, J. Mater. Res. 16, 1636 (2001). 71. J. Font, J. Muntasell, and E. Cesari, Mater. Res. Bull. 36, 1665
45. L. B. Kong, J. Ma, H. T. Huang, W. Zhu and O. K. Tan, Mater. Lett. (2001).
50, 129 (2001). 72. C. Bay, R. J. Spontak, C. C. Koch, C. K. Shaw and C. M. Balik,
46. L. Lu and M. O. Lai, “Mechanical Alloying.” Kluwer Academic Polymer 41, 7147 (2000).
Publishers, Boston, MA, 1998. 73. C. DeCastro and B. Mitchell, unpublished observations.
47. J. S. Rawers, G. Seavens, P. Govier, C. Dogan and R. Doan, Metall. 74. B. D. Cullity, “Elements of X-ray Diffraction,” p. 356 Addison-
Mater. Trans. A 27, 3126 (1996). Wesley, Reading, MA, 1978.
48. R. Angers, M. R. Krishnadev, R. Tremblay, J. F. Corriveau and 75. G. K. Williamson and W. H. Hall, Acta Metall. 1, 22 (1953).
D. Dube, Mater. Sci. Eng., A 262, 9 (1999). 76. S. M. Zhu, M. Tamura, K. Sakamoto and K. Iwasaki, Mater. Sci.
49. K. A. Khor and Y. Li, Mater. Sci. Eng., A 256, 271 (1998). Eng., A 292, 83 (2000).

You might also like