You are on page 1of 12

Effects of Process-Control Agents on Mechanical Alloying of

Nanostructured Aluminum Alloys


L. SHAW, M. ZAWRAH, J. VILLEGAS, H. LUO, and D. MIRACLE
The effects of the process-control agents (PCAs) stearic acid and methanol on the mechanical alloying
(MA) of a nanostructured aluminum alloy (Al93Fe3Ti2Cr2) have been investigated. The dependency
of the powder-particle sizes, grain sizes, atomic-level strains, lattice parameters, formation of solid
solutions, and microstructural evolution of the aluminum alloy on the types of PCAs and their
concentrations have been studied using a variety of analytical instruments including X-ray diffraction
scanning electron microscopy, and transmission electron microscopy. The results clearly indicate that
prevention of excessive cold welding of Al particles can be achieved by the addition of a PCA at
the expense of reductions in the grain size, formation rate of solid solutions, and rate of microstructural
refinement, all of which are desired in MA of the Al alloy. Furthermore, a PCA that is more effective
in preventing excessive cold welding will also impose more hindrance to the MA process. These
phenomena have been discussed in the light of the adsorption of the PCA on the metal surface and
the lubricating function of the PCA.

I. INTRODUCTION

RECENT progress in the studies of bulk amorphous,


quasi-crystalline, and nanocrystalline alloys has resulted in
substantial improvements in the strength of Al-based
alloys.[110] Bulk Al alloys with tensile strengths as high as
800 to 1000 MPa and reasonable ductility (2 to 10 pct tensile
elongation) have been demonstrated recently.[8,9,10] Many of
these Al alloys derive their superior strengths from either a
microstructure of nanoscale fcc Al crystals distributed in
an amorphous matrix[7,9] or a microstructure containing a
mixture of nanoscale fcc Al particles and intermetallic aluminides.[10] However, in some cases, the superior strength
comes from a mixed microstructure of quasi-crystalline
nanoparticles distributed in a continuous fcc Al matrix.[5,1113]
This group of Al alloys also has excellent strength at elevated
temperatures, because of the unusual ability of quasi-crystalline phases in retaining their strength at high temperatures.[12,13]
The Al93Fe3Ti2Cr2 alloy is an example of the last group
of the Al alloys mentioned previously.[5] This alloy has a
mixed microstructure of quasi-crystalline (i.e., icosahedral
phase) nanoparticles distributed in a continuous fcc Al phase
in the gas-atomized condition, and this mixed microstructure
remains unchanged in the extruded bulk alloy.[5] It is shown
that the extruded Al93Fe3Ti2Cr2 alloy possesses an ultimate
tensile strength of 650 MPa at room temperature and 360
MPa at 300 C.[5] This high-temperature strength compares
very favorably against the current commercial Al alloys,
most of which lose their useful strengths at temperatures
above 250 C (typically becoming lower than 200 MPa).[14]

Recently, we have shown that the hardness of the mechanical


alloying (MA)processed Al93Fe3Ti2Cr2 alloy can be twice
the hardness of the 6061-Al alloy at the T6 condition and
higher than that of the cold-drawn 1045 steel.[15] In this case,
however, the high hardness of the Al93Fe3Ti2Cr2 alloy stems
from the formation of nanocrystalline supersaturated fcc
Al solid solutions. In spite of the high hardness obtained,
excessive cold welding of Al particles is observed in the
MA-processed Al93Fe3Ti2Cr2 alloy, and this excessive cold
welding has resulted in nonuniform microstructures and
properties.[15]
One of the effective methods to avoid excessive cold
welding is the application of surface-active substances, generally known as process-control agents (PCAs). Many
researchers have used PCAs to prevent excessive cold welding of Al particles during milling.[1625] The addition of a
PCA, however, could also alter the nature of the milling
product,[26] introduce contaminants,[27] and slow down the
MA process.[24] In spite of the important role and multiple
influences of PCAs in mechanical alloying of ductile materials, most of the studies in the open literature have been
devoted to the effect of PCAs on powder-particle sizes and
contamination.[24,27,28] Few studies on other effects of PCAs
exist in the open literature. Thus, the present study is carried
out with two objectives in mind. One is to conduct a systematic and comprehensive investigation of effects of PCAs on
the MA process, and the other is to optimize MA parameters
used in making the Al93Fe3Ti2Cr2 alloy based on the understanding developed in this study. The findings from this
investigation are presented as follows.
II. EXPERIMENTAL PROCEDURE

L. SHAW, Associate Professor, and J. VILLEGAS and H. LUO, Research


Assistants, are with the Department of Metallurgy and Materials Engineering, Institute of Materials Science, University of Connecticut, Storrs,
CT 06269. Contact e-mail: LShaw@mail.ims.uconn.edu M. ZAWRAH,
Senior Scientist, is with the Refractory and Ceramic Department, National
Research Center, Cairo, 12622-Dokki, Egypt. D. MIRACLE, Senior Scientist, is with the Air Force Research Laboratory, Materials and Manufacturing
Directorate, Wright-Patterson AFB, OH 45433.
Manuscript submitted February 28, 2002.
METALLURGICAL AND MATERIALS TRANSACTIONS A

Crystalline elemental powders were used to prepare Al


alloys with a nominal composition of Al93Fe3Ti2Cr2. The
aluminum powder used had a purity of 99.5 wt pct, with a
mean particle size of 70 m, while the corresponding values
for iron, chromium, and titanium powders used were 99.0,
98.5, and 99.5 pct and 50, 30, and 30 m, respectively. The
as-received aluminum powder was made via atomization,

U.S. GOVERNMENT WORK


NOT PROTECTED BY U.S. COPYRIGHT

VOLUME 34A, JANUARY 2003159

Table I. Milling Conditions for Al93Fe3Ti2Cr2 Alloy


PCA Used

Concentration
(Wt Pct)

None
Stearic acid
Methanol

0.5, 1.0, and 2.0


1.0, 1.5, and 2.0

Milling Duration
(h)
2 to 16
2 to 64
2 to 34

whereas the iron powder was manufactured via a vaporphase chemical decomposition process, the chromium powder via electrolytic plating and milling, and the titanium
powder via the hydride technique followed by a dehydride
process. The MA was performed in a Spex 8000 (SPEX
CertiPrep, Inc., Metuchen, NJ) laboratory mill using a stainless steel vial and balls, with a ball-to-powder weight ratio
of 5:1. Two different sizes of steel balls (1.0 and 0.5 cm
in diameter) were used simultaneously to provide uniform
milling for different sizes of powder particles. The vial
loaded with powders of the desired average composition
was sealed in a glove box under an argon atmosphere, and
milling was conducted in the stationary argon atmosphere.
Two types of PCAs, stearic acid (CH3(CH2)16COOH) and
methanol (CH3OH), were investigated. A summary of various milling conditions that were studied is presented in Table
I. The alloy without a PCA was also investigated, to provide
a baseline for comparison. The milling duration ranged from
2 to 64 hours, depending on the milling conditions used. A
fan was used to cool the vial during milling, and at the end
of each milling process, all the powder was taken out for
analyses of microstructures and powder characteristics. For
the alloys without the addition of a PCA, both large (1 to
5 mm) and small (0.2 to 0.5 mm) Al alloy particles were
generated. However, only the results of small particles were
used to compare with those obtained from the alloys with
the addition of a PCA.
Powders before and after milling were analyzed using Xray diffraction (XRD) methods with Cu K radiation. In
addition to monitoring the evolution of the crystal structure,
XRD was also used to measure the crystallite size and internal strains. This was accomplished through the use of the
Win-Crysize software,[29] which is based on the Warren
Averbach theory[30] and includes the consideration of instrumental broadening. The specific surface area (SSA) of
powder particles was quantified using nitrogen adsorption
based on the BraunauerEmmettTeller theory.[31] The particle morphology, microstructures, and distribution of elements within the particle were examined using a scanning
electron microscope (SEM) equipped with energy-dispersive
spectrometry (EDS). Analyses of the grain size, crystal structure, and chemical composition within the microstructure
were also performed using a PHILIPS* CM200 field-emis*PHILIPS is a trademark of Philips Electronic Instruments Corp., Mahwah, NJ.

sion gun transmission electron microscope (TEM). In the


TEM analysis, both bright- and dark-field image techniques,
coupled with selected-area diffraction (SAD), convergentbeam electron diffraction (CBED), and EDS with an electron
beam of 3.5 nm, were used to characterize powder particles.
160VOLUME 34A, JANUARY 2003

Fig. 1The SSA of Al93Fe3Ti2Cr2 alloys with 1 wt pct stearic acid as a


function of milling time.

Fig. 2The radii of equivalent spherical particles computed from the SSA
data of the Al alloys with 1 and 2 wt pct stearic acid.

III. RESULTS
Shown in Figure 1 is the SSA of Al alloy powders with
1 wt pct stearic acid as a function of milling time. It can be
seen that the SSA decreases substantially during the initial
four hours of milling, indicating an increase in particle size
due to cold welding. After 4 hours of milling, however, the
SSA increases rapidly and appears to approach a dynamic
balance between cold welding and fracturing beyond 16
hours of milling. Assuming that the powder particles formed
during milling are spheres, the radii of these equivalent
spherical particles can be calculated from the SSA and are
presented in Figure 2. Included in the figure are also the
radii of equivalent spherical particles computed from the
SSA data of the Al alloy powder with 2 wt pct stearic acid.
As seen in the figure, the particle size varies substantially
for the alloy powder with 1 wt pct stearic acid. In contrast,
the particle size of the alloy powder with 2 wt pct stearic
acid is insensitive to the milling time, indicating that 2 wt
pct stearic acid could prevent excessive cold welding even
at the early stage of milling.
The SEM images of powder morphology for several MAprocessed powders are shown in Figures 3(a) through (d),
which correspond to the data points a through d marked
in Figure 2. Note that most of the powder particles with 4
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 3SEM images of the Al alloy milled with 1 wt pct stearic acid for (a) 4 h, (b) 8 h, (c) 16 h, and (d ) 64 h.

hours of milling have sizes ranging from 100 to 500 m


(Figure 3(a)), whereas particle sizes range from 20 to 100
m, 5 to 70 m, and 3 to 40 m for powders milled for 8,
16, and 64 hours, respectively. Clearly, the trend of the
particle-size change with milling time observed in the SEM
is consistent with the measurement of the SSA. However,
the particle size observed in the SEM is substantially larger
than the size of equivalent spherical particles calculated from
the SSA data. The discrepancy comes from two sources,
one being the nonspherical shapes of powder particles, and
the other being the tortuous surface of particles. The former
source contributes to the discrepancy only at the early stage
of milling, while the latter source contributes at both early
and later stages. As can been seen in Figures 3(a) through
(d), particles become more or less spherical after prolonged
milling, especially with 64 hours of milling.
The XRD patterns of the Al alloy powder with 1 wt pct
stearic acid are shown in Figure 4. An XRD pattern of
the powder mixture without milling is also included for
comparison. It is noted that Fe, Cr, and Ti reflections are
hardly ascertainable even for the powder mixture without
milling, because Fe, Cr, and Ti concentrations are low and
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 4XRD patterns of the Al alloy with 1 wt pct stearic acid (SA) as
a function of milling time. For comparison, the XRD pattern of the starting
Al alloy powder (0 wt pct SA, 0 h) is also included.

some of their peaks overlap with Al peaks. Thus, the detection of the formation of Al-Fe-Cr-Ti alloys and other structural information could only rely on the measurement of
VOLUME 34A, JANUARY 2003161

Fig. 5XRD patterns of the Al alloys milled for 16 h but with different
amounts of PCA. Key: SAstearic acid; and MLmethanol.
Fig. 7The atomic level strain of fcc-Al in the Al alloy powder mixtures
without and with 1 wt pct stearic acid as a function of milling time.

Fig. 6Calculated crystallite sizes of fcc-Al in the Al alloy powder mixtures with and without stearic acid as a function of milling time.
Fig. 8The lattice parameter of fcc-Al in the Al alloy powder mixtures
without and with 1 wt pct stearic acid as a function of milling time.

changes in Al reflections. A close examination of Figure 4


reveals the following two interesting features.
(1) The peak position of Al reflections shifts to higher 2
angles as the milling time increases. This suggests the
formation of Al-based solid solutions and an increase
in the solute concentration as the milling time increases.
(2) All aluminum reflections exhibit broadening even after
only 2 hours of milling, indicating the grain-size reduction and, possibly, the introduction of internal strains.
Furthermore, the longer the milling time, the larger
the broadening.
Shown in Figure 5 are XRD patterns of Al alloys with
the same duration of milling, but different amounts of PCAs.
As can be seen from the figure, the powder mixture without
any PCA has the smallest integral area and the largest full
width at half maximum (FWHM) of the XRD reflections.
The powder mixtures with 1 wt pct stearic acid and 1 wt pct
methanol appear to have similar integral areas and FWHMs,
while the powder mixtures with 1.5 and 2.0 wt pct methanol
have the largest integral areas and the smallest FWHMs.
Based on the peak broadening, the crystallite size and
atomic-level strain have been estimated. Comparisons of the
crystallite size and atomic-level strains among the alloy
powders with different amounts of stearic acid are shown
in Figures 6 and 7. As seen from Figure 6, the calculated
162VOLUME 34A, JANUARY 2003

crystallite size decreases rapidly at the early stage of milling


and levels off at prolonged milling times. Furthermore, the
more PCA in the powder, the smaller the decrease in the
crystallite size. The crystallite size of the powder mixture
without a PCA and with 4 hours of milling is very close to
that of the powder mixture with 1 wt pct stearic acid and 8
hours of milling and to that of the powder mixture with 2
wt pct stearic acid and 16 hours of milling. Thus, the addition
of stearic acid, although preventing excessive cold welding,
decreases the grain refinement rate. For the atomic-level
strain, the addition of stearic acid slows down the strainincrease rate (Figure 7).
Using peak positions of Al(220), (311), and (222) in the
XRD patterns, the average lattice parameters of fcc Al calculated from the spacings of these three crystallographic planes
for each milling condition are shown in Figure 8. It is noted
that the lattice parameter decreases with increasing milling
time, suggesting the formation of fcc Albased solid solutions. The observed decrease in the lattice parameter is consistent with the possible formation of Al solid solutions with
Fe, Cr, and Ti solutes. The dissolution of Fe, Cr, and Ti
elements into fcc Al will decrease the lattice parameter of
fcc Al, because the atomic sizes of Fe and Cr elements are
smaller than that of Al, whereas the atomic size of Ti is
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 9The lattice parameter of fcc-Al in the Al alloy powder mixtures


with different amounts of PCA (wt pct). All powders have been milled for
16 h except the starting powder. Key: SAstearic acid; MLmethanol;
and Originalthe starting powder.

Fig. 10Crystallite sizes of fcc-Al in the Al alloy powder mixtures milled


for 16 h with different amounts of PCA (wt pct). Key: SAstearic acid;
and MLmethanol.

similar to that of Al.[32] It is further noted that the addition


of stearic acid has decreased the lattice-parameter reduction
rate. For example, to achieve the same lattice parameter for
the powder mixture without a PCA after 16 hours of milling,
64 hours of milling is required for the powder mixture with
1 wt pct stearic acid.
The types of PCAs and their concentrations also have
substantial effects on the particle size, crystallite size,
atomic-level strain, and lattice parameter. Shown in Figure
9 is a summary of the effect of different types of PCAs and
their concentrations on the lattice parameter of fcc Al. It
can be seen that the powder mixture having zero PCAs with
16 hours of milling has the largest reduction in the lattice
parameter among those listed. At the same milling time, the
powder mixture with 1 wt pct stearic acid has the secondlargest reduction in the lattice parameter. This is followed
by the powder mixture with 1 wt pct methanol. Finally, as
the concentration of methanol increases, the reduction in the
lattice parameter decreases.
The effect of different types of PCAs and their concentrations on the calculated crystallite size of fcc Al is shown in
Figure 10. The trend observed in the variation of the lattice
parameter with types of PCAs and their concentrations also
holds for the variation of the crystallite size, that is, the
powder mixture without a PCA has the largest reduction in
METALLURGICAL AND MATERIALS TRANSACTIONS A

the crystallite size, while the powder mixture with 2 wt


pct methanol has the smallest reduction. Furthermore, the
powder mixture with 1 wt pct stearic acid has a larger crystallite-size reduction than the powder mixture with 1 wt
pct methanol. These results clearly indicate that 1 wt pct
methanol imposes more hindrance than 1 wt pct stearic acid
to the crystallite-size reduction and the formation of Albased solid solutions. On the other hand, it has been observed
in this study that 1 wt pct methanol is sufficient in preventing
excessive cold welding of Al particles, whereas 1 wt pct
stearic acid is not.
The microstructural evolution of Al alloy powder particles
with 1 wt pct stearic acid as a function of milling time is
shown in Figure 11. The EDS technique was utilized in
these SEM analyses in order to identify the chemical composition of the small particles distributed within the large Al
particles (termed as the Al matrix hereafter). The compositions of the small particles so identified are marked in the
SEM images to facilitate the discussion. When the size of
a particle (or several particles) is too small to positively
identify its main composition, the composition at that particular location will be marked as an Al-based solid solution
using symbols such as Al(Ti), Al(Ti,Fe), Al(Fe), and Al(Ti,Fe,Cr) to indicate Al-Ti, Al-Ti-Fe, Al-Fe, and Al-Ti-Fe-Cr
solid solutions, respectively. The spatial resolution of the
EDS analysis in the present alloy system and the experimental condition employed is determined to be about 2 m.
At the early stage of milling (2 and 8 hours), a striated
structure, the typical microstructure observed in ball-milled
ductile-ductile material systems,[18,33] is observed. However,
as the milling time increases (e.g., 16 and 64 hours), the
striated structure disappears and is replaced by a microstructure of fine particles distributed in an Al matrix. These fine
particles are Cr-, Ti-, and Fe-rich particles and, in most
cases, contain Al solutes (to be discussed more in the following TEM analysis). Furthermore, Fe particles appear to dissolve and/or break up more quickly than Ti and Cr particles.
As summarized in Table II, at 8 hours of milling, Fe-rich
particles have the smallest size, while Cr-rich particles have
the largest size. This trend continues as the milling time
increases. In fact, at 16 hours of milling, it is extremely
difficult to find an Fe-rich particle larger than 0.25 m, and
this is true for Ti-rich particles when the milling time is
increased to 64 hours. Recall that the average sizes of the
starting Fe, Ti, and Cr particles are 50, 30, and 30 m,
respectively. Thus, it is certain that Fe particles have the
fastest breakup and/or dissolution rate, whereas Cr particles
have the slowest breakup and/or dissolution rate in the present Al alloy system.
Shown in Figure 12 are the backscattered electron images
of the microstructure of Al alloys with 16 hours of milling
and the addition of 1 or 2 wt pct methanol. Again, it is found
that the largest particles within the aluminum matrix are Crrich particles, and the smallest particles are Fe-rich particles.
Thus, the replacement of stearic acid by methanol does not
alter the order of the breakup and/or dissolution rate of Fe,
Ti, and Cr particles. When comparisons are made among
the alloys with 1 wt pct stearic acid (Figure 11(c)), 1 wt pct
methanol (Figure 12(a)), and 2 wt pct methanol (Figure
12(b)), it is noted that the alloy with 2 wt pct methanol
contains the largest number of fine particles, while the alloy
with 1 wt pct methanol has the second-largest number, and
VOLUME 34A, JANUARY 2003163

Fig. 11SEM backscatter electron images of the microstructure of Al alloys with 1 wt pct stearic acid milled for (a) 2 h, (b) 8 h, (c) 16 h, and (d ) 64 h.
The letters in the micrographs indicate the major composition of that particular location detected using EDS. The text provides more details.

Table II. The Size of Most of the Cr-, Ti-, and Fe-Rich
Particles within the Al Matrix
Milling Time
(h)

Size of Cr
Particles (m)

Size of Ti
Particles (m)

Size of Fe
Particles (m)

8
16
64

10
4
0.55

4
3
0.25*

1
0.25*
0.25*

*0.25 m is the smallest size of particles whose composition


can be qualitatively identified.

the alloy with 1 wt pct stearic acid has the lowest number.
Since the formation of solid solutions requires dissolution
of fine particles, these results suggest that the higher the
PCA concentration, the smaller the formation rate of solid
solutions. Furthermore, 1 wt pct methanol decreases the
solid-solution formation rate more than 1 wt pct stearic acid.
This result is in accordance with the effect of stearic acid
and methanol on the reduction in the lattice parameter and
crystallite size of fcc-Al (Figures 9 and 10).
To confirm the formation of nanosized grains and Albased solid solutions, TEM analyses coupled with SAD,
CBED, and EDS have been performed on Al alloys milled
for 16 hours without a PCA. The results are shown in Figures
164VOLUME 34A, JANUARY 2003

13 and 14. Figure 13(a) and (b) show the typical morphology
of nanosized Al grains and the associated SAD pattern,
which corresponds to fcc Al, respectively. Most of the grains
are elongated, reflecting the severe plastic deformation
which occurred during MA. The grain size ranges from
10 to 100 nm, depending on the location. The comparison
between the XRD and TEM analyses indicates that the grain
size estimated by XRD is at the lower end of the TEM
result. Figures 13(c) and (d) present a large polycrystalline
Cr particle embedded in the Al matrix and its associated
SAD pattern, respectively. The size of this elongated Cr
particle is 0.5 1.5 m. In general, Cr particles are relatively
large and easy to find in the TEM analysis. However, both
Ti and Fe particles are difficult to find. Shown in Figures
13(e) and (f) are an Fe particle embedded in the Al matrix
and its associated CBED pattern, respectively. The size of
this particle is 50 120 nm. Thus, the findings using the
TEM are consistent with those obtained from the SEM, that
is, Cr-rich particles are the largest in size and Fe-rich particles are the smallest. Using EDS, it is further found that all
the Cr, Fe, and Ti particles contain some Al elements (to be
discussed more in the following text).
The EDS analysis of TEM thin foils indicates that the
composition of the Al powder particle is nonuniform. The
nonuniformity manifests itself in two ways. One is the presMETALLURGICAL AND MATERIALS TRANSACTIONS A

IV. DISCUSSION

Fig. 12SEM backscatter electron images of the microstructure of Al


alloys milled for 16 h with (a) 1 wt pct methanol and (b) 2 wt pct methanol.
The letters in the micrographs indicate the major composition of that
particular location detected using EDS. The text provides more details.

ence of undissolved Cr, Fe, and Ti particles, and the other


is the nonuniform composition of the Al matrix, which may
contain almost zero alloying elements, only one element
(e.g., Fe, Ti, or Cr), two elements (e.g., Fe and Ti, Ti and
Cr, or Cr and Fe), or all three of the alloying elements,
depending on the location examined. Shown in Figures 14(a)
through (d) are several examples of the EDS analysis. Note
that these EDS spectra are obtained with an electron-beam
diameter of 3.5 nm. Figure 14(a) is an EDS spectrum taken
from an elongated Al grain, showing almost zero alloying
elements. Figure 14(b) is an EDS spectrum taken from
another elongated Al grain at a different location, showing
the formation of the Al-Fe solid solution. Figure 14(c) is
the EDS spectrum that contains both Ti and Fe elements,
showing the formation of Al-Ti-Fe solid solutions. Figure
14(d) is the EDS spectrum taken from the center of the Crrich particle shown in Figure 13(c), indicating the formation
of the Cr-based solid solution. Shown in Figure 14(e) is the
EDS spectrum acquired from the center of the Fe-rich particle shown in Figure 13(e), indicating the formation of the
Fe-based solid solution. Thus, the formation of Al-based
solid solutions found from the high-resolution transmission
electron microscopy analysis is in good agreement with the
XRD analysis (Figures 8 and 9).
METALLURGICAL AND MATERIALS TRANSACTIONS A

The present set of experiments clearly shows that the


addition of stearic acid and methanol can prevent excessive
cold welding of the Al alloy powder. This observation is
consistent with the results from other researchers.[16,24] In
the absence of the PCA, mechanical alloying of the Al
alloy powder is dominated by cold welding. As a result, the
particle size of the Al alloy powder can increase to as large
as 5 mm during milling. However, with the addition of 2
wt pct stearic acid or 1 wt pct methanol, it is found that the
powder-particle size changes little with milling time (Figure
2). This indicates that the rates of cold welding and fracturing
of Al powder particles with 2 wt pct stearic acid or 1 wt
pct methanol are in dynamic equilibrium. Thus, it can be
concluded that the PCA can reduce the cold-welding rate,
or increase the fracturing rate, or both. The reduction in the
cold-welding rate can be achieved by the adsorption of the
PCA on the metal surface, thereby impeding the clean metalto-metal contact necessary for cold welding. The increase
in the fracturing rate can result from the occlusion of the
PCA at the interface of two cold-welded particles, which,
in turn, can lead to a reduced interfacial strength and, thus,
a high fracturing rate. However, chemical analyses of milled
power particles[27] appear to suggest that the dominant mechanism for preventing excessive cold welding by the PCA is
due to the reduction of the cold-welding rate rather than
the increase in the fracturing rate. Chemical analyses have
indicated that the occlusion of the PCA in the molecular
form at the interface is not a major event during milling.[27]
Thus, the prevention of excessive cold welding is most likely
achieved via the adsorption of the PCA on the metal surface,
which impedes the clean metal-to-metal contact.
The prevention of excessive cold welding requires a certain amount of PCA. When the addition of the PCA is not
sufficient (e.g., 1 wt pct stearic acid), the particle size will
increase initially and then decrease as milling proceeds (Figures 2 and 3). Such a phenomenon has also been observed
by other researchers.[24] The decrease in the particle size
after prolonged milling in the present study (e.g., beyond 4
hours of milling for 1 wt pct stearic acid) is likely due to
work hardening and the formation of Al-based solid solutions, both of which can lead to reduced ductility and, thus,
increase the fracturing rate.
The most interesting result from the present study is that
the extent of plastic deformation of Al powder particles is
also affected by the addition of the PCA. Both the increase in
the lattice strain (Figure 7) and the crystallite-size reduction
(Figures 6 and 10) are indicative of the extent of plastic
deformation. Lattice strains can arise from the presence of
solutes in solid solutions, lattice distortion near grain boundaries, and the presence of dislocations. However, the main
contribution to lattice strains at the early stage of milling is
probably the dislocation density, because grain sizes are
large and no solid solutions are formed at the early stage of
milling. Thus, the present study shows that for a given
milling time, the addition of stearic acid decreases the extent
of plastic deformation, thereby leading to low dislocation
densities and lattice strains (Figure 7).
Detailed TEM studies of ball-milled powders[34] have
indicated that the sequence of the nanocrystallization process is as follows: (1) the generation of high-dislocationdensity regions, (2) the formation of subgrains with lowVOLUME 34A, JANUARY 2003165

Fig. 13TEM bright-field images and the associated diffraction patterns of Al alloys milled for 16 h with no PCA. (a) The bright-field image of typical
Al nanograins, (b) the associated SAD pattern of image (a), (c) one large polycrystalline Cr-rich particle embedded in the Al matrix, (d ) the associated
SAD pattern from the Cr-rich particle in image (c), (e) one small Fe-rich particle embedded in the Al matrix, ( f ) the associated CBED pattern from the
Fe-rich particle in image (e) with the zone axis of [111].

angle grain boundaries in these heavily deformed regions,


and (3) the further reduction in the size of subgrains, which
eventually become nanograins with complete random orientations. Another separate study[35] on warm extrusion of
aluminum has shown that the stable subgrain size of Al
166VOLUME 34A, JANUARY 2003

depends on the equilibrium dislocation density. All these


data suggest that the size of nanograins is proportional
to the dislocation density. Therefore, the crystallite-size
reduction can be used as an indication of the extent of
plastic deformation. As such, the crystallite-size reduction
METALLURGICAL AND MATERIALS TRANSACTIONS A

(a)

(b)

(c)

(d)

(e)
Fig. 14EDS spectra from TEM analyses. (a) An EDS spectrum taken from an elongated Al grain, (b) an EDS spectrum taken from another elongated
Al grain at a different location, (c) an EDS spectrum taken from an elongated Al grain at the location different from (a) and (b), (d ) the EDS spectrum
taken from the center of the Cr-rich particle shown in Fig. 13(c), and (e) the EDS spectrum acquired from the center of the Fe-rich particle shown in Fig. 13(e).

with and without the PCA also indicates that, at a given


milling time, the addition of the PCA reduces the extent
of plastic deformation (Figures 6 and 10). This result is
consistent with the trend shown by the lattice-strain variation discussed previously.
The reduced plastic deformation due to the presence of
the PCA is likely related to the lubricating effect of the
PCA. Ball milling of powder particles can be approximated
as a mini-forging process.[36] The degree of plastic deformation imparted to the powder particles in each ball-to-ball
collision or ball-to-wall collision is related to the kinetic
energy of the colliding balls, but is also affected by the
rigidity of the powder compact, since the mini forging in
ball milling takes place without lateral constraints. When
powder particles can slide laterally with ease during collision
METALLURGICAL AND MATERIALS TRANSACTIONS A

owing to the lubricating effect of the PCA, the kinetic energy


of the colliding balls will be consumed in overcoming the
friction force and making particles slide past one another.
As a result, powder particles have less plastic deformation
when the PCA act as lubricants. Furthermore, the more PCA,
the less plastic deformation the powder particle has, and,
thus, the less crystallite-size refinement and the smaller lattice strain.
Effects of the PCA on the lattice-parameter reduction and
microstructural evolution can also be explained based on
the adsorption of the PCA on the metal surface and the
lubricating function of the PCA. The present study clearly
shows that the refinement of Fe, Ti, and Cr particles and
the dissolution of these elements into fcc Al take place
during milling. The dissolution of the alloying elements into
VOLUME 34A, JANUARY 2003167

fcc Al requires a direct contact between the Al particle and


the alloying elements. Thus, the adsorption of the PCA on
the metal surface can impede the dissolution process. Furthermore, the more PCA, the slower the dissolution process,
and, thus, the less the reduction in the lattice parameter
(Figures 8, 9, and 12).
The present SEM study of the microstructure of Al powder
particles as a function of milling time suggests that the
refinement of Fe, Ti, and Cr particles takes place in the
following sequence: (1) flattening of Fe, Ti, and Cr particles
(Figure 11(a)), (2) breaking down of flattened particles into
fine equiaxed particles (Figure 11(b)), and (3) dissolution
of the fine particles into the Al matrix (Figure 11(d)). Thus,
the refinement of Fe, Ti, and Cr particles within the Al
matrix is related to the degree of plastic deformation as
well as the dissolution rate. The former is affected by the
lubricating effect of the PCA, whereas the latter can be
altered by the adsorption of the PCA on the metal surface.
As discussed previously, the addition of the PCA reduces
the extent of plastic deformation in each collision and the
dissolution rate of the alloying elements into the Al matrix.
As a result of these reductions, it is expected that the more
PCA added, the more alloying particles that will be retained
in the Al matrix. This expectation is in good agreement with
the experimental observation (Figure 12).
The effect of 1 wt pct methanol on the MA process is
different from that of 1 wt pct stearic acid. Essentially, 1 wt
pct methanol results in smaller particle sizes, larger crystallite sizes, and larger lattice parameters than with 1 wt pct
stearic acid. These results indicate that 1 wt pct methanol
is more efficient in preventing excessive cold welding, yet
at the same time imposes more hindrance to the grain-size
refinement and the formation of solid solutions than does 1
wt pct stearic acid. These observations are consistent with
the analysis carried out previously, that is, the prevention
of excessive cold welding can be achieved by adding a PCA
at the expense of grain-size refinement and solid-solution
formation. Using a PCA that is more effective in preventing
excessive cold welding will also slow down the MA process more.
The reason for 1 wt pct methanol and 1 wt pct stearic
acid to have different impacts is not clearly understood yet.
However, several possible sources for the different effects
may be considered. First, the difference may stem from the
physical properties of the two organic compounds. Stearic
acid has a melting point of 67 C to 69 C and a boiling
point of 183 C to 184 C, while the corresponding values
for methanol are 98 C and 64.6 C, respectively.[24] Since
the average temperature of the vial during milling is about
50 C (measured by pressing a thermocouple against the
surface of the vial as soon as the Spex mill is turned off
after running for several hours), methanol will be in a liquid
state, while stearic acid will be in a solid state. The different
physical states of the two organic compounds may result in
different effects on the MA process observed.
The two organic compounds also have different phasetransformation behaviors at the collision site. Many studies
have suggested that the temperature spike at the collision
site could be 180 C,[37] 350 C,[38] and 500 C[39] higher than
the average temperature, depending on the milling conditions
and powder characteristics. As a result of this temperature
spike (e.g., an 180 C increase), methanol at the collision site
168VOLUME 34A, JANUARY 2003

Fig. 15Schematic of one of the possible adsorption models for (a) methanol and (b) stearic acid on the Al alloy surface. It is assumed that hydrogen
elements have been dissociated from the -OH and -COOH groups in methanol and stearic acid, respectively, in the adsorption process, because the
chemical analysis of the gas phase during ball milling indicates that the
gas phase is mainly composed of hydrogen for these two types of PCA.[27]

will undergo a liquid-to-gas phase transformation, whereas


stearic acid will experience a solid-to-gas transformation.
Such different phase transformations may have influence on
the MA process.
Finally, another possible source for the different effects
of the two organic compounds may be due to different
absorptions of molecules on the metal surface. Assuming
that the absorption of methanol on the metal surface takes
a form as shown in Figure 15(a), the coverage of the metal
surface by 1 wt pct methanol can then be computed using
the SSA data. Taking 1 kg of Al alloy powder before milling
as an example, the total surface area is 400 m2. Based on
the length of the CH bond and the HCH bond angle,[40]
it can be assumed that each methanol molecule takes up a
surface area of about 0.2 0.2 nm2 if the adsorption takes
the form as shown in Figure 15. Further, if a monolayer
adsorption is assumed, 1 wt pct methanol, which contains
0.3125 moles of molecules in this case, will take up a surface
area of 7,524 m2, which is about 18 times larger than the
available powder surface area of 400 m2. A similar calculation can also be performed for 1 wt pct stearic acid if it is
assumed that molecules of the adsorbed stearic acid stand
up as shown in Figure 15. The total surface area that 1 wt
pct stearic acid can cover is 1,907 m2 if each stearic acid
molecule takes up a surface area of about 0.3 0.3 nm2.
Note that 1 wt pct stearic acid is also sufficient to provide
full coverage of the metal surface area. Thus, the difference
is that 1 wt pct methanol provides more molecular layers
in covering the surface of powder particles, whereas the
molecule of stearic acid is larger than that of methanol.
Thus, this analysis seems to suggest that multiple molecular
layers are more effective in preventing excessive cold welding and impose more hindrance in the crystallite-size refinement and the formation of solid solutions than are large
molecules with fewer molecular layers.
V. CONCLUDING REMARKS
In this study, the effects of the PCAs, stearic acid and
methanol, on the MA of Al93Fe3Ti2Cr2 alloy powder have
been investigated. Based on this study, the following conclusions can be offered.
METALLURGICAL AND MATERIALS TRANSACTIONS A

1. The addition of stearic acid and methanol can effectively


prevent excessive cold welding of Al particles during
ball milling.
2. The addition of stearic acid and methanol decreases the
rate of the grain-size refinement, the lattice-stain increment, and the formation of Al-based solid solutions. Furthermore, the higher the concentration of a PCA, the
more hindrance to the grain-size refinement, the latticestain increment, and the formation of Al-based solid
solutions.
3. The microstructural evolution of Al alloy powder particles is affected by the addition of stearic acid and methanol. For a given milling condition, the higher the
concentration of the PCA, the more Fe, Ti, and Cr particles remain undissolved.
4. The effects of stearic acid and methanol on the prevention
of excessive cold welding, the grain-size refinement, the
lattice-strain increment, and the formation of Al-based
solid solutions are most likely achieved via the adsorption
on the metal surface and the lubricating function provided
by these molecules. The former prevents the clean metalto-metal contact necessary for cold welding, whereas the
latter reduces the degree of plastic deformation of powder
particles during each impact, which, in turn, slows down
the grain-size refinement and the formation of solid
solutions.
5. To completely suppress excessive cold welding, a certain
amount of the PCA is needed. In the case of stearic acid,
a 2 wt pct addition is sufficient, whereas a 1 wt pct
addition is not. For methanol, 1 wt pct is sufficient.
6. A PCA that is more effective in preventing excessive
cold welding will also impose more hindrance to the
grain-size refinement, the lattice-strain increment, and
the formation of Al-based solid solutions.
The implication of the present study for making Al-based
nanostructured materials via the MA approach is that the
selection of PCAs and their quantities should be intelligently
judged. If too little PCA is used, excessive cold welding
will occur and the resulting product will have a nonuniform
microstructure and properties. However, if too much PCA
is used, the ball-milling time has to be extended substantially,
because the PCA impedes the grain-size refinement, the
formation of solid solutions, and the homogenization of
microstructures, all of which are normally desired in MA.
The selection of the PCA quantity should also include the
consideration of its molecular formula, because different
PCAs have different efficiencies. Finally, the addition of the
PCA will alter the composition of MA-processed products
slightly,[27] which should also be taken into account for the
precise control of the alloy composition.

ACKNOWLEDGMENTS
The first author (LS) is grateful to the University of Connecticut for granting a sabbatical leave to conduct research
at the Air Force Research Laboratory. The second author
(MZ) thanks the Academy of Scientific Research and Technology, Egypt for providing a fellowship for him to conduct
this research at the University of Connecticut. The partial
support from the Laboratory Directors Fund of the Air Force
Research Laboratory and the University of Connecticut
METALLURGICAL AND MATERIALS TRANSACTIONS A

Research Foundation are also greatly appreciated. The


authors also appreciate the help provided by Dr. Robert
Wheeler, UES, Inc., in initiating the TEM experiment.

REFERENCES
1. Y. He, S.J. Poon, and G.J. Shiflet: Science, 1988, vol. 241, pp. 1640-42.
2. A. Inoue, K. Ohtera, A.-P. Tsai, and T. Masumoto: Jpn. J. Appl. Phys.,
1988, vol. 27 (3), pp. L280-L282.
3. A. Inoue, K. Ohtera, and T. Masumoto: Jpn. J. Appl. Phys., 1988,
vol. 27 (5), pp. L736-L739.
4. G.J. Shiflet, Y. He, and S.J. Poon: J. Appl. Phys., 1988, vol. 64 (12),
pp. 6863-65.
5. A. Inoue and H. Kimura: Mater. Sci. Eng., 2000, vol. A286 (1), pp.
1-10.
6. A. Inoue and T. Masumoto: Mater. Sci. Eng., 1991, vol. A133, pp. 6-9.
7. H. Chen, Y. He, G. Shiflet, and S.J. Poon: Scripta Metall. Mater.,
1991, vol. 25, pp. 1421-24.
8. A. Inoue: Mater. Sci. Eng., 1994, vols. A179A180, pp. 57-61.
9. Z.C. Zhong, X.Y. Jiang, and A.L. Greer: Mater. Sci. Eng., 1997, vols.
A226A228, pp. 531-35.
10. J.Q. Guo and K. Ohtera: Acta Mater., 1998, vol. 46 (11), pp. 3829-38.
11. A. Inoue: Mater. Sci. Forum, 1997, vols. 235238, pp. 873-80.
12. A. Inoue, H.M. Kimura, K. Sasamori, and T. Masumoto: Mater. Trans.,
JIM, 1995, vol. 36 (1), pp. 6-15.
13. A. Inoue, H. Kimura, K. Sasamori, and T. Masumoto: Mater. Trans.,
JIM, 1996, vol. 37 (6), pp. 1287-92.
14. D.L. Erich: Development of A Mechanically Alloyed Aluminum
Alloy for 450650 F Service, AFML-TR-79-4210, US Dept. of
Commerce, National Technical Information Service (NTIS), Springfield, VA, 1980.
15. M. Zawrah and L. Shaw: Mater. Sci. Eng., 2002, in press.
16. J.S. Benjamin and M.J. Bomford: Metall. Trans. A, 1977, vol. 8A,
pp. 1301-05.
17. R.F. Singer, W.C. Oliver, and W.D. Nix: Metall. Trans. A, 1980, vol.
11A, pp. 1895-1901.
18. P.S. Gilman and W.D. Nix: Metall. Trans. A, 1981, vol. 12A, pp.
813-24.
19. J. Eckert, J.C. Holzer, C.E. Krill, and W.L. Johnson: J. Mater. Res.,
1992, vol. 7 (7), pp. 1751-61.
20. S. Srinivasan, S.R. Chen, and R.B. Schwarz: Mater. Sci. Eng., 1992,
vol. A153, pp. 691-95.
21. S. Srinivasan, P.B. Desch, and R.B. Schwarz: Scripta Metall. Mater.,
1991, vol. 25, pp. 2513-16.
22. Z. Peng, C. Suryanarayana, and F.H. Froes: Scripta Metall. Mater.,
1992, vol. 27, pp. 475-80.
23. W.E. Frazier and M.J. Koczak: Scripta Metall., 1987, vol. 21, pp.
129-34.
24. L. Lu and M.O. Lai: Mechanical Alloying, Kluwer Academic Publishers, Norwell, MA, 1998.
25. S. Enzo, R. Frattini, G. Mulas, and F. Delogu: Mater. Sci. Forum,
1998, vols. 269272, pp. 391-96.
26. A.P. Radlinski, A. Calka, B.W. Ninham, and W.A. Kaczmarek: Mater.
Sci. Eng., 1991, vol. A134, pp. 1346-49.
27. A. Arias: Chemical Reactions of Metal Powders with Organic and
Inorganic Liquids during Ball Milling, NASA TN D-8015, US
National Aeronautics and Space Administration, National Technical
Information Service (NTIS), Springfield, VA, 1975.
28. A.M. Harris, G.B. Schaffer, and N.W. Page: in Mechanical Alloying for
Structural Applications, J.J. deBarbadillo, F.H. Froes, and R. Schwarz,
eds., ASM INTERNATIONAL, Materials Park, OH, 1993, pp. 15-19.
29. WIN-CRYSIZE software, version 3.05, Sigma-C Corp., Munich,
Germany.
30. B.E. Warren: Progress in Metal Physics, Pergamon Press, London,
1957, vol. 8.
31. S. Brunauer, P.H. Emmett, and E. Teller: J. Am. Chem. Soc., 1938,
vol. 60, p. 309.
32. P. Haasen: Physical Metallurgy, 3rd Ed., Cambridge University Press,
Cambridge, Great Britain, 1996, p. 142.
33. J.S. Benjamin and T.E. Volin: Metall. Trans., 1974, vol. 5, pp. 1929-34.
34. H.J. Fecht, E. Hellstern, Z. Fu, and W.L. Johnson: Metall. Trans. A,
1990, vol. 21A, pp. 2333-37.
35. H.J. McQueen and J.J. Jonas: in Treatise on Materials Science and
VOLUME 34A, JANUARY 2003169

Technology, vol. 6, Plastic Deformation of Materials, R.J. Arsenault,


ed., Academic Press, New York, NY, 1975, pp. 393-493.
36. D.R. Maurice and T.H. Courtney: Metall. Trans. A, 1990, vol. 21A,
pp. 289-303 (1990).
37. G.B. Schaffer and P.G. McCormick: Metall. Trans. A, 1992, vol. 23A,
pp. 1285-90.

170VOLUME 34A, JANUARY 2003

38. R.M. Davis, B. McDermott, and C.C. Koch: Metall. Trans. A, 1988,
vol. 19A, pp. 2867-74.
39. R.B. Schwarz and C.C. Koch: Appl. Phys. Lett., 1986, vol. 49 (3),
pp. 146-48.
40. Handbook of Chemistry and Physics, D.R. Lide, ed., CRC Press, Boca
Raton, FL, 1990.

METALLURGICAL AND MATERIALS TRANSACTIONS A

You might also like