Sie sind auf Seite 1von 60

Progress in Materials Science 51 (2006) 160

www.elsevier.com/locate/pmatsci

Synthesis and mechanical behavior of


nanostructured materials via cryomilling
D.B. Witkin
a

a,*

, E.J. Lavernia

Department of Chemical Engineering and Materials Science, University of California,


Irvine, CA 92697-2575, United States
Department of Chemical Engineering and Materials Science, University of California,
Davis, CA 95616, United States

Abstract
Cryomilling, the mechanical attrition of powders within a cryogenic medium, is a method
of strengthening materials through grain size renement and the dispersion of ne, nanometerscale particles. The technique was developed as a means to decrease both the size of these particles and their spacing within a metallic matrix to increase threshold creep stress and intermediate temperature performance. More recent work has been concerned with increasing the
strength of lightweight structural materials. In this overview paper, the available literature
is reviewed that covers the microstructural evolution during cryomilling, consolidation and
processing, the thermal stability of the microstructure, and mechanical properties of consolidated materials. The properties of cryomilled materials are compared to those results for powders and consolidated materials generated by mechanical alloying, milling at ambient
temperatures and other means to produce ne grained materials.
 2005 Elsevier Ltd. All rights reserved.

Corresponding author. Tel.: +1 949 8243426; fax: +1 949 8242541.


E-mail addresses: dwitkin@uci.edu (D.B. Witkin), lavernia@ucdavis.edu (E.J. Lavernia).

0079-6425/$ - see front matter  2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pmatsci.2005.04.004

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Contents
1.

2.

3.

4.

Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1. Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2. Apparatus and processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3. Description of consolidated cryomilled Al . . . . . . . . . . . . . . . . . . . . .
Structural evolution of cryomilled materials . . . . . . . . . . . . . . . . . . . . . . . .
2.1. Formation of nanocrystalline structure . . . . . . . . . . . . . . . . . . . . . . .
2.2. Thermal stability of cryomilled powders. . . . . . . . . . . . . . . . . . . . . . .
2.3. Grain size characteristics and distributions . . . . . . . . . . . . . . . . . . . . .
2.4. Primary consolidation of cryomilled powder . . . . . . . . . . . . . . . . . . . .
2.5. Processing of primary consolidated forms. . . . . . . . . . . . . . . . . . . . . .
Mechanical properties of cryomilled materials . . . . . . . . . . . . . . . . . . . . . . .
3.1. Overview of mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2. Room temperature deformation: strengthening mechanisms of cryomilled Al .
3.3. Room temperature mechanical properties: nature of deformation . . . . . .
3.3.1. Inhomogeneous flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4. Room temperature mechanical properties: compression testing of Al 5083.
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2
3
4
5
6
6
13
16
17
23
35
35
36
44
44
52
55
57
57

1. Introduction
The increase in strength of crystalline materials as their grain size decreases has
prompted signicant research into producing bulk nanocrystalline materials for
structural applications. This research eort has comprised a number of techniques
and approaches to achieving a bulk material with nanometer-scale grains. One such
approach begins with the production of nanocrystalline powder followed by consolidation into a bulk form. The initial step of producing nanocrystalline powders can
take one of two forms. In the rst, a change of physical state, such as gas condensation or rapid solidication, leads to particles which themselves are nanometric. In the
second, the powder particles themselves may be on a micron scale, but the
microstructure of the particles is nanocrystalline, having a grain size that falls in
the 10100 nm range. Such particles may be produced by severe plastic deformation
techniques, such as ball milling.
Cryomilling is a mechanical attrition technique in which powders are milled in a
slurry formed with milling balls and a cryogenic liquid. It is important to maintain a
distinction between this process, in which the powder charge is in intimate contact
with the cryogenic liquid, and deformation occurring at cryogenic temperatures by
chilling a milling vessel from without (e.g., see [1]); the term cryomilling has been
used for both. As a process relying on severe plastic deformation, the formation
of nanostructures within the powder particles can be compared to other ball milling
methods, as well as to microstructural renement of bulk materials, attained via
equal channel angular pressing (ECAP), friction stir processing, and cold working.

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

As a powder metallurgy technique, it requires a consolidation step, and changes in


microstructure and properties during consolidation can be compared with those
for consolidated or compacted nanoparticles and traditional ball-milled powders.
In this review, the cryomilling process and its products are considered, including
the formation of nanostructured powders and their thermal behavior and stability.
The microstructural changes that occur during consolidation and other secondary
processing are also considered. Finally, the mechanical properties of the bulk forms
will be discussed, and the implications for deformation of ne-grained materials will
be evaluated. The discussion focuses on metals with particular attention paid to Al
alloys and their potential as high-strength structural materials. In each case, the
properties and performance of cryomilled materials will be compared to various
nanostructured analogues as described above. The focus on structural materials necessarily ignores the potential use of cryomilled materials in other applications. For
example, there have been research eorts to use cryomilled powders as feedstocks
for spray-processed coatings [24].
1.1. Motivation
The cryomilling technique was originally developed at Exxon Research and Engineering [5]. The rst description of cryomilling in the issued U.S. Patent [5] was for
an yttriated iron alloy, and indicated that mechanically alloying this alloy in liquid
nitrogen allowed much shorter milling times to reach the nest particle sizes and
smaller recrystallized grain sizes when compared to mechanical alloying performed
in air or in Ar. Cryomilling was rst described in the literature for a composite
AlAl2O3 [6]. The goal of the research at that time can be understood in context
of the development of dispersion-strengthened alloys. Dispersion strengthening
had been attempted to improve creep resistance, and had been advanced by mechanical alloying of metallic and non-metallic constituents [79]. The threshold stress for
creep resistance and high-temperature strength were increased when dispersoid particle size and interparticle spacing were reduced. Mechanical alloying of Al was aided
by the addition of an organic process control agent (PCA), which was necessary to
mediate the welding of particles [9,10]. PCAs in common use in milling of Al have
typically included methanol, stearic acid, and paran compounds. The presence
of the organic material led to dispersion strengthening by oxides and carbides derived from the organic compound. In a well-known review published contemporaneously with the rst report of cryomilling, Gleiter indicated that eorts to produce
nanocrystalline powders of f.c.c. metals had not been successful in part because of
in-situ sintering of the powder into millimeter-sized agglomerates [11]. This discrepancy is important in considering both ball milling as a technique and the science and
engineering aspects of nanocrystalline materials. If ones ambition is to interrogate
the fundamental properties of an elemental material in a particular lattice system,
e.g., face-centered cubic, as its grain size is reduced below 100 nm, then the presence
of contamination, whether metal particles derived from milling debris, or secondary
phases and particles from process control agents, clearly compromises the integrity of
the material. On the other hand, if one is trying to produce an engineering material

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

with a particular combination of strength, density and creep resistance, then the
addition of secondary particles may be desirable if not indeed practical and
necessary.
Nevertheless, there has been an acknowledgement of microstructural renement
as a desirable characteristic. Several papers on the microstructural characterization
of cryomilled NiAl published in the early 1990s [1215], for example, do not discuss
the grain size or microstructure with respect to nanocrystalline or nanostructured character. In the development of this alloy, the goal was enhancing strength
by dispersion for a specic temperature range, not general strengthening from grain
renement. On the other hand, recent papers on bulk cryomilled Al explicitly refer to
alloy characterization and development in terms of grain size renement and
strengthening.
1.2. Apparatus and processing
As described by Luton et al. [6], the rst cryomilling experiments were undertaken
in an eort to reduce the size and spacing of the dispersoid particles. The Exxon
cryomilling apparatus consisted of a 10 l Szegvari attritor mill that was modied
to allow continuous ow of liquid nitrogen into the mill. The milling environment
was a circulating slurry consisting of the milling balls, powder and owing liquid
nitrogen. Thermocouple probes were used to monitor the temperature of the mill
and the level of the liquid nitrogen. The rst powder charges described by the Exxon
group were 99.5% purity Al combined with 3%, 7% and 15% 50 nm particles of alumina [6]. Soon after, results for cryomilled NiAl intermetallic were reported, which
indicated superior high-temperature mechanical performance relative to NiAl and
NiAl composites processed by other methods [14]. The microstructure and mechanical properties, with a focus on creep between 750 and 1150 C, were reported extensively during the 1990s [1222].
The basic design of a cryomilling apparatus, pictured schematically in Fig. 1a, has
not changed in recent years, regardless of the scale of the mill. The milling apparatus
typically includes thermocouples to monitor the attritor and ensure that a constant
level of liquid nitrogen and thus a constant milling environment is maintained. The
liquid nitrogen ows into the vessel continuously; evaporated nitrogen is exhausted
through a blower equipped with a particle lter to contain any powder particles entrained by the gas ow. Changes in temperature immediately above the level of the
slurry are used to open and close valves controlling liquid nitrogen ow. Attritors in
current use include 10 l cryomills at University of California, Davis that are typically
used with 1 kg powder charges (Fig. 1b) and a large commercial attritor with a
capacity of 35 kg of powder operated under the auspices of Boeings Rocketdyne
division.
Upon the conclusion of a cryomilling run, the powder is collected in a slurry with
liquid nitrogen by means of a gravity draining valve. The powder can be transferred
under the liquid nitrogen cover to a glove box, where the nitrogen is allowed to evaporate. The next processing step for the powder is a degassing step to remove volatile
contaminants, such as water vapor or residual PCA. In processing materials for

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Fig. 1. Schematic design of a typical cryomilling attritor mill (a). Photograph of a mill in use with a 500 g
powder charge (b).

structural applications, the milled powder has typically been transferred to Al cans
tted with hollow stems and valves for hot vacuum degassing. Once the degassing
cycle is completed, the can is then sealed at the stem to preserve the vacuum. Beginning with work at Exxon [6], the most common primary consolidation route has
been Hot Isostatic Pressing (HIPping), often followed by extrusion of the HIP billet.
Work on cryomilled NiAl compared extrusion of canned powders with HIPped powders [13]. Ongoing research activities include alternative consolidation methods, such
as Cold Isostatic Pressing (CIPping) and secondary processing and forming
methods, including dierent types of extrusion, open-die forging, hot pressing, and
rolling.
1.3. Description of consolidated cryomilled Al
Processing steps in Exxons cryomilled AlAl2O3 composites included HIPping
and extrusion, followed by cold swaging to a diameter appropriate for mechanical
testing. The results of this testing will be discussed subsequently; for introductory
purposes, the microstructure of the materials will be described presently. Luton
et al. [6] reported that the microstructure of the extruded material consisted of three
constituents. The rst was ne grained aluminum with a grain size ranging from 50
to 300 nm. The second was the alumina particles, which had retained their size
(50 nm) during the milling and consolidation process. The nal constituent was described as a reasonably uniform distribution of exceedingly ne dispersoids in the
size range 210 nm [6]. These dispersoids were subsequently revealed using
high-resolution transmission electron microscopy (HRTEM) as platelets roughly
1015 nm wide and only a few atomic layers thick [23]. Computed images based

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

on HRTEM and convergent beam diraction indicated that the dispersoids were
essentially a single atomic layer of nitrogen in tetrahedral positions in the fcc Al lattice or a single layer of oxygen atoms in the octahedral positions of the lattice, both
typically forming on [1 1 1] planes in Al. The formation of these platelets was surmised to come from the incorporation of nitrogen and oxygen during the fracturing
and rewelding process during milling. Subsequent study of cryomilled NiAl indicated
that the AlN phase formed during milling, or at room temperature or below, and
that the platelets were not a result of elevated temperature processing during consolidation or extrusion [12]. The spacing of these ne dispersoids was estimated to be
80 nm [6].
Given the general microstructural characteristics, it is worthwhile to compare the
cryomilled Al composite to examples of the oxide dispersion strengthened materials
which prompted development of the cryomilling technique. Mechanically alloyed Al
that was hot pressed and extruded and contained dierent concentrations of carbon
and oxygen based on the weight percent PCA used in milling had a grain size of
roughly 500 nm and contained two types of alumina dispersoids [10]. The rst were
equiaxed particles with a diameter of 30 nm that originated from oxide lms coating
the premilled Al powder particles, while the second were aky oxide particles
roughly 10 nm thick and 100 nm wide. The second type of particle was surmised
to have formed as a lm on the milled powder which was subsequently broken up
during extrusion. The microstructure of a mechanically alloyed Al that was cold
compacted and then vacuum hot pressed was described as having sub-grains with
diameters of approximately 300 nm [24]. This material contained 30 nm oxide particles with a spacing of 60 nm, and Al4C3 particles less than 10 nm in size separated by
42 nm. The cryomilled AlAl2O3 composite is thus very similar in microstructure to
the mechanically alloyed Al. In assessing the properties of cryomilled alloys, comparison to their mechanically alloyed analogues is reasonable in evaluating their performance and understanding their behavior. For background, a recent review has
provided a comprehensive treatment of mechanical milling and alloying [25].

2. Structural evolution of cryomilled materials


2.1. Formation of nanocrystalline structure
As nanocrystalline materials have attracted increased attention, the mechanisms
by which the nanostructure arises have been investigated more extensively. A basic
empirical description of the development of nanostructures during mechanical milling was presented by Fecht [26]. This was a three-stage process of grain renement,
starting with the localization of deformation into shear bands with high dislocation
density, which is followed by annihilation and recombination of dislocations, forming nanometer-scale sub-grains. This sub-grain structure extends throughout the
sample during continued milling. The nal stage is the transformation of sub-grain
boundary structure to randomly oriented high-angle grain boundaries [26]. A schematic depiction of the process at the microscopic level is given in Fig. 2, which is

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Fig. 2. Schematic representation of grain renement mechanism during ball milling, adapted from Table
III, in [27].

based on the work of Xun et al. [27]. The relationship between the development of
nanocrystalline grain dimensions and concomitant changes in powder particle morphology was recently suggested [27]. In this model, based on experimental evidence
for a cryomilled Zn22Al eutectoid alloy, the attening of the powder particles from
their original spherical shape was matched by an increasing dislocation density of the
pre-milled coarse grained structure. These attened particles were fragmented with
increasing milling time, as the microscopic processes of sub-boundary formation
take place. Cold welding of particles into roughly equiaxed, but not spherical
agglomerates corresponded to increasing grain boundary angles between sub-grains
and the formation of randomly oriented nanoscale grains [27].
The minimum grain size achievable by ball milling, dmin, has been related to several physical properties of elemental metals. In the case of melting temperature, there
is an inverse relation to dmin for higher melting fcc metals, but for hcp and bcc systems as well as fcc elements with melting temperatures above that of Pd (1828 K),
dmin is constant [28,29]. The constant portion of the fcc, bcc and hcp families plot
as three separate parallel curves with dmin values ordered as fcc < bcc < hcp [28]. If
the value of dmin is normalized by the magnitude of the respective Burgers vector,
however, all three systems, including some alloys, can be represented by an equation
of the form [30]
d min
A expcT m ;
b

where A and c are constants. For fcc metals, an inverse relationship was also found
between dmin and bulk modulus, and a direct relationship between dmin and the critical equilibrium distance between two edge dislocations [29]. More recently, Mohamed [30,31] was able to support this empirical model with a demonstration of
the relationship between dmin and the activation energy for recovery, which in turn
can be related to melting temperature and bulk modulus. These ndings are summarized in Table 1. One consequence of these correlations especially pertinent to a discussion of cryomilling is the prediction that milling at lower temperatures will not
reduce dmin appreciably in metals with lower self-diusion activation energies, such
as Al. In apparent contradiction to this conclusion is the behavior exhibited by Zn
milled at cryogenic temperatures (but not within the cryogenic medium), in which
crystallites from 2 to 6 nm were observed after short milling times [1], which would

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Table 1
Minimum grain size for dierent milling experiments
Material

Minimum grain size


(nm)

Inverse dependence
of dmin/b

Cryomilling

Mechanical
milling

Tm (K)

26 2 [34]
25.5 [47]

22 2 [29]
25 [42]

933

Ni

16 5 [43]

12 3 [29]

Fe

21 [45]

8.1 [42]
9.4 4.5 [46]

Al

B (GPa)

Direct dependence
of dmin
Qd (kJ/mol)

Lc (nm)

75

143

10.4

1726

177

289

4.5

1808

170

281

4.4

necessitate renement of both the empirical model [26] as well as the fundamental
model [30]. The appearance of ne grains was attributed to nucleation via dynamic
recrystallization, and their proportion decreased on continued milling while the overall average grain size of the population decreased. These ne grains may thus have
been a metastable condition, and Zn can be seen to represent an outlier when dmin
is normalized by b and plotted against melting temperature or activation energy
for self-diusion [31].
Another factor to consider in the grain renement process is microstructural inhomogeneity. The empirical model cited previously [26] describes deformation localized
into shear bands that ultimately comprise the entire sample. It would not be surprising, therefore, to nd microstructural features that manifest dierent relative degrees
of deformation, and not simple lognormal distributions of randomly oriented equiaxed grains. In cryomilled Al7.5Mg, two distinctive components of the microstructure were observed after cryomilling eight hours: a random distribution of equiaxed
grains between 10 and 30 nm diameter, and less frequently, elongated grains measuring 100200 nm long by 30 nm wide [32,33]. Similar microstructure was observed in
cryomilled commercial purity Al [34]. TEM investigation of cryomilled Al7.5Mg
0.3Sc revealed that the mean width of the elongated regions in cryomilled powder
was essentially equal to the average grain size of equiaxed grains [35]. These elongated grains were compared to the lamellar structures that arise during cold rolling
of wrought Al to high strains [36,37], with continued deformation taking place within the individual lamellae, producing sub-grains that are roughly equiaxed [34]. This
microstructure is thus consistent with the phenomenological model [26], but the
appearance of high-angle grain boundaries separating elongated regions as observed
in cold worked Al is not necessarily anticipated by the ball milling model. A similar
microstructure has been noted in AlMg processed by ECAP [38]. The implication
for the cryomilled AlMg, though not stated, is that these microstructures are evidence of insucient milling time. Elongated grains of similar dimensions were also
reported as a transient feature in cryomilled Inconel 625, present in samples milled
for four and six hours, but not in powder milled eight hours [39]. It should be noted,
however, that variations in cryomilling conditions, including the ball to charge mass
ratio and the size of the mill, can signicantly inuence the energy of the milling envi-

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

ronment. Given that, comparing two sets of data this way, beyond noting the similar
microstructures, is at best inconclusive and likely inadvisable. At present, the lamellar features observed in the cryomilled powders may represent an intermediate or
transient stage in structural evolution not explicit in the oft-cited empirical model
developed for high-energy ball milling performed at ambient temperature [26]. Like
the aforementioned early appearance of nanometer crystallites in Zn during milling
at cryogenic temperatures [1], the lamellar structures indicate that renements in the
model are necessary to assess the role of milling or deformation temperature.
Dierences in both theory and experiment related to the structural evolution and
minimum grain size may be examined in light of comparison of grain size evolution
with milling time and dmin for cryomilled and mechanically milled materials. The list
of systems for comparison is not extensive, as a result of the relatively limited number of systems of cryomilled powders which have received extensive study that overlap studies of mechanically milled or alloyed powders. The relative rate of grain size
decrease as a function of milling time for various elements and alloys is shown in Fig.
3. As stated in the cryomilling patent [5], one perceived benet of the cryomilling
process is quicker achievement of nal or minimum grain size. As seen in Fig. 3, this
claim may be true in the case of Al, but the cryomilling data in this case are for Al
7.5Mg [40]. The dierences in milling are not limited to milling medium and temperature, however, as the ball milled data in Fig. 3 [29,41,42] are from SPEX laboratory
mills, which operate at much higher frequency than planetary mills. The dierences
in processing thus include the mechanical energy of the mill, the mass of the powder
charge and the ball to charge ratio.
On the other hand, dmin shows good agreement between the two types of milling.
Reported minimum grain sizes for dierent milling experiments are given in Table 1.

80
Cryo Al-7.5Mg [40]

70

Cryo Inconel 625 [39]


Al [42]

Grain Size (nm)

60

Al [29]
AlRu [41]

50

Fe [42]
Ni [29]

40
30
20
10
0
0.1

10

100

1000

Milling Time (hours)


Fig. 3. Grain size as function of milling time for cryomilled and ball milled powders.

10

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

The grain size of mechanically milled Ni was 12 3 nm [29], while for cryomilled Ni
it was 22 4 nm via XRD measurement and 16 5 nm using TEM images [43].
Cryomilled Fe had a grain size of approximately 21 nm [44,45], while mechanically
milled Fe has been reported with a minimum grain size of 8.1 nm [42] and
9.4 4.5 nm [46]. For Al, the average grain size obtained after cryomilling was reported as 26 2 nm [34] and 25.5 nm [47], while for mechanical milling it was
22 2 nm [29] and approximately 25 nm [42]. While the cryomilling results described
here are all for powders prepared using a Szegvari attritor operated under similar
conditions (albeit with milling times ranging from 8 h for Al [34] to 25 h for Fe
[44,45]), the powders for mechanical milling were milled under various conditions
for widely varying times, depending on the reference. Much longer cryomilling runs
on the intermetallic NiAl (up to 170 h) indicated that the continuous uptake of N
and O and reaction with Al led to the loss of the B2 crystal structure and its replacement by Ni in solid solution with Al [20]. The nal grain size was reported to be between 6 and 25 nm, which is consistent with the metallic systems of Al and Ni.
Another aspect of the microstructure to consider is the presence of the Al-bearing
dispersoids, such as AlN. While Al2O3 or carbide particles may also be present, as in
mechanically alloyed materials, the formation of AlN is the result of milling in the
liquid nitrogen environment. Examples of high-resolution TEM of dispersoids in extruded cryomilled Al10Ti2Cu are shown in Fig. 4 [48]. In Fig. 4a, a single dispersoid with a length of approximately 10 nm is pictured, while in Fig. 4b, several
dispersoids are shown clustered near a grain boundary. There have been some dierences reported in the size and nature of the dispersoids. Reported values for the dispersoid particle spacing come from materials consolidated and processed by dierent
means, and widely varying temperatures, so they likely do not represent the as-milled
distribution for each of the systems. While the original cryomilled AlAl2O3 composite were platelets about 10 nm and separated by 80 nm [6], in NiAl the dispersoids
ranged in size from 5 to 100 nm [12], and were distributed inhomogeneously after
consolidation [13,14]. For the cryomilled eutectic system Zn22%Al, an average dispersoid size of 14 nm was calculated, reecting a distribution range of particle diameter from less than 5 nm up to 65 nm diameter [49]. While this was not a normal
distribution for particle size, the mean and median were roughly coincident. For
cryomilled Zn22Al, the dispersoid separation distance was estimated to be about
130 nm.
The formation of the AlN or Aloxynitride phases has not been investigated to a
great extent. The original HRTEM on the cryomilled AlAl2O3 indicated that N-rich
platelets with a thickness of only several atomic layers form on the family of [1 1 1]
planes [23]. The formation of Ni3N during cryomilling of Ni has been discussed in
more detail [43,50]. Of particular note is the nding that Ni3N led to anisotropic
residual strain in fcc Ni, where higher strain was measured in the (2 0 0) plane than
in other lattice planes [50]. In addition, the relative amount of microstrain measured
in the (2 0 0) diraction peak correlated to the N content of the powder, which was in
turn a function of milling time. The uptake of N by the octahedral sites leading to the
residual strain in the lattice could thus be controlled by selection of milling conditions [50]. At the same time, the thermodynamic stability of Ni3N is considerably less

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

11

Fig. 4. High-Resolution TEM images of AlN dispersoids: (a) single dispersoid, and (b) multiple
dispersoids adjacent to grain boundary. Micrographs provided in advance of publication [48] by Prof. R.S.
Mishra.

than that of AlN (Table 2). As will be discussed in the next section, decomposition of
the Ni3N was proposed as one cause of grain growth in annealed cryomilled Ni [43].
The relative stability of AlN in comparison likely contributes to the greater thermal
stability stability of cryomilled Al relative to Ni at similar homologous temperatures
(Fig. 5a). Also shown in Table 2 are thermodynamic data for Cr and Fe nitrides.

12

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Table 2
Molar heats of formation and free energies of metalnitride compounds [142]
Compound

DH298 (kJ)

DG300 (kJ)

AlN
CrN
Cr2N
Fe4N
Ni3N

318.6
123.1
114.7
10.9
0.8

287.1
96.7
93.8
4.2
27.2

350
Fe
Fe2.6Al

Average Grain Size (nm)

300

Fe10Al

Ni

250

Al [47]
Al-7.5Mg [40]

200

Al-7.5Mg [32]
Inconel 625 [39]

150
100
50
0
0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0.7

0.8

Homologous Temperature (T/Tm)


350
Cryo Fe [44]
Cryo Fe2.6Al [44]

300

Grain Size (nm)

Cryo Fe10Al [44]


Fe [46]

250

Fe [60]
Fe IGC [61]

200

Fe milled [61]

150
100
50
0
0

0.1

0.2

0.3

0.4

0.5

0.6

Homologous Temperature

Fig. 5. (a) Grain growth data for isothermal annealing experiments of various cryomilled alloys; (b)
Comparison of grain growth data for cryomilled Fe and FeAl alloys and mechanically milled Fe.

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

13

Values for the former are also more favorable than those for Ni, suggesting an
important role in the cryomilling of NiCr alloys (e.g., Inconels). Iron is typically
present as a contaminant from milling tools, but cryomilling in the presence of a
hard secondary phase, as has been performed for AlAlN [5154] and NiAlN
[55], led to a direct relationship between the volume of the added AlN and the Fe
contamination in both. The relatively low stability of FeN (in this case Fe2N)
was held responsible for the thermal desorption of N during heating of the cryomilled AlAlN powder [52].
2.2. Thermal stability of cryomilled powders
Unlike nanoparticles, which have applications in the particulate form [56,57],
many investigations of nanocrystalline materials produced via cryomilling or
mechanical milling have at least implicitly acknowledged their use as structural materials. This requires the consolidation of the powders, which has often been evaluated
a priori by isothermal annealing experiments to assess the thermal stability of the
microstructure. The annealing experiments are often accompanied by thermal analysis such as dierential scanning calorimetry (DSC), to determine, for example, the
kinetics of grain growth, recovery or recrystallization. If these can be identied,
along with the grain sizes at the respective annealing temperatures, grain growth
and accompanying loss of strength might be minimized during consolidation steps.
In Fig. 5a, grain growth measurements published for several cryomilled metals
and alloys are shown graphically as a function of homologous melting temperature,
including Ni [43], the Ni-based superalloy Inconel 625 [58], Fe and FeAl alloys
[44,45], Al [47], and Al7.5Mg [40]. For consistency, only data for annealing times
of one hour are plotted except for Al, where data for 30 min are plotted. The discrepancy for the Al data is ameliorated by noting that for annealing temperatures below
0.80Tm the grain growth did not vary appreciably for annealing times of 30 and
100 min [47]. The time-independence of grain growth during annealing was also true
for Al7.5Mg and for Ni below annealing temperatures of 0.62Tm [43]. On the basis
of the available data for cryomilled alloys presented in Fig. 5a, cryomilled elemental
Al would appear to be more thermally stable than Ni or Fe. In the case of Ni, this
might be attributed to the decomposition of Ni3N above 500 C [43]. In general,
however, the formation of a stable oxide layer on the pre-milled Al powders probably adds to resistance to grain growth by incorporation into the mechanically milled
microstructure, as was previously described for mechanically milled Al [10]. The two
FeAl alloys exhibit the same limited increase in grain size as the Al and the Al
7.5Mg, which may also be attributed to the higher stability of AlN compared to
Fenitrides (Table 2). This is further indicated by the role of adding Al to a spray
atomized M50 high-speed steel, where there was no additional stability imparted
by the addition of 5 wt.% Al prior to milling [59]. In this case, the thermal stability
may have resulted from the formation of various carbides and oxides in the absence
of AlN and alumina. In fact, the grain size of cryomilled M50 steel was less without
the Al addition while the grain growth trend with annealing temperature was essentially the same as the cryomilled FeAl [45]. The importance of the dispersoids

14

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

cannot, however, be dismissed in stabilizing the grain size. For example, when NiAl
was cryomilled in liquid Ar, no AlN or Al(ON) could form, and the grain growth
during isothermal annealing led to grain sizes roughly an order of magnitude larger
than when the NiAl was milled in liquid nitrogen [20]. As shown in Fig. 5a, Inconel
625 departs from the usual trend, as annealing at 900 C for one hour actually led to
a smaller grain size than at 700 and 800 C. This was attributed to the precipitation
at elevated temperature of two niobium-bearing phases, Ni3Nb and NbC, and their
stabilizing eect on the grain size [58].
The cryomilled Fe and FeAl alloys are plotted with data for mechanically milled
Fe [46,60,61] and Fe produced by inert gas condensation [61] in Fig. 5b, indicating
that grain growth appears to initiate in powders from both milling environments at a
similar temperature. The data for Fe from Moelle and Fecht [60] are presented with
the caveat that they are for annealing treatments of approximately 2400 s, not 1 h.
The nature of the grain growth has prompted some inquiries based on kinetics
analysis of thermal data. Investigation of both mechanically milled Fe [46] and cryomilled Al [47] has indicated two dierent values of activation energy for grain growth
depending on the annealing time and temperature. In both cases, lower temperature
data gave a better t to the parabolic grain growth relationship given by
1=n

D1=n  D0 kt;

where D0 and D are the grain size initially and at time t, respectively, and k is an
Arrhenius-type rate constant that includes the activation energy for grain growth,
Qgg. The denominator of the exponent, n, would be 0.5 in the ideal case where the
only driving force on the grain boundary results from its curvature. Deviations from
this value, which are the norm [62], are attributed to pinning of the grain boundaries,
which may be another factor contributing or impeding grain growth. The higher
temperature data were tted to a model that accounts for pinning forces on a migrating grain boundary. Both relied on Burkes equation:


D0  D
Dm  D0
ln
kt;
3
Dm
Dm  D
where the additional term Dm refers to the maximum grain size resulting from the
pinning force. Malow and Koch [46] found that the activation energy for grain
growth in mechanically milled Fe was almost double at 0.40.45Tm than it was
for the lower temperatures. The higher value of 248 kJ/mol corresponded closely
to values obtained for polycrystalline Fe and lattice self diusion of Fe. Because this
value is also close to the values for diusion of solute elements present due to contamination during milling (e.g., Mn, Co, Cr or Ni), solute-dependence of the grain
growth was suggested. At lower temperatures and shorter annealing times, the calculated activation energy was signicantly lower than for the grain boundary diusion of Fe, another obvious basis for comparison considering the results of previous
work on nanocrystalline metals reported by the authors [46]. There is, however, no
clear-cut explanation for the lower activation energy value or the existence of two
grain growth mechanisms. The cryomilled Al followed a similar trend, with higher
temperature annealing treatments, in this case from 0.78 to 0.83Tm, leading to an

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

15

activation energy in reasonable agreement with, albeit lower than, the lattice self diffusion activation energy [47]. In the cryomilled material, the presence of dispersoids
derived from milling in liquid nitrogen provides a pinning force that would not be
present in the mechanically milled Fe. As will be discussed in more detail in considering the microstructural evolution during consolidation and extrusion, the grain
sizes observed in these materials tend to deviate from those predicted by particle pinning due to the Zener theory. In the context of grain growth during isothermal
annealing, a considerable increase in grain size from approximately 60 nm to nearly
300 nm on annealing 300 s at 773 and 873 K, respectively (0.83 and 0.94Tm) was correlated to a large exothermic DSC signal at roughly 850 K and attributed to unpinning of grain boundaries [47]. The lower temperature value, 79 kJ/mol, was closer to
the grain boundary self diusion energy for Al (84 kJ/mol) than in the case of Fe. A
dierent scenario was encountered in cryomilled Ni, as a single activation energy was
derived over the entire temperature range investigated, 0.450.62Tm [43], leading to
an activation energy value that was in close agreement with the grain boundary self
diusion energy for Ni. As the authors point out, this agreement may be more fortuitous than truly indicative of the underlying mechanism, as the grain boundary diffusion activation energy may well be lower for ne grained materials, which have
larger grain boundary volumes, than coarse grained conventional materials. If so,
the calculated value would reect a combination of some lower value for grain
boundary self diusion and an additional source, such as unpinning. In this case,
these two purported mechanisms could not be separated over the range of temperatures investigated, as was the case for Fe and Al.
To summarize, the isothermal grain growth studies on cryomilled powders suggest
several trends, none of which appear to be universal. First, cryomilled Fe and
mechanically milled Fe exhibit similar grain growth behavior as a function of annealing temperature. Cryomilled Ni is similar to Fe in this respect. Cryomilled Al exhibits greater stability to higher homologous temperatures. The addition of Al as an
alloying element to Fe stabilizes the grain size such that the curves for Al and Fe
Al are similar. These results are consistent with the relative stability of the respective
nitrides (Table 2) and enhanced grain size stability owing to greater nitride stability.
A similar comparison for cryomilled Al and mechanically milled Al could not be
made based on a lack of overlap in the literature with respect to annealing conditions
and grain size determination (i.e., conrmation of XRD grain size by direct measurement using TEM). The relative lack of data for nanocrystalline Al processed by
mechanical milling may be better appreciated by recalling the necessity of adding
a process control agent to limit cold welding. The behavior of milled Al would thus
be aected by the presence of wear debris from the mill (which is less of a problem
for milling Fe with steel tools) as well as organic contaminants. In addition, the easily oxidized Al makes handling of a pure specimen dicult to achieve. Comparison of dierent impurity elements arising during cryomilling and conventional
milling is given in Table 3. Assessment of activation energies for grain growth leads
to an incomplete picture. While grain growth in cryomilled Al and mechanically milled Fe is controlled by temperature-dependent mechanisms, in cryomilled
Ni a single growth mechanism was determined, albeit over a narrower range of

16

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Table 3
Concentrations of impurities during milling

Al
Al
Al
Al
Ni
Ni
NiAl

Mill

Milling time (h)

Cryo [47]
Spex [24]
Spex [29]
Szegvari [42]
Cryo [43]
Spex [29]
Cryo

16
3
24
600
15
40
16 [12]
170 [20]

0.26
1.14

0.038

0.03
0.176

O
0.85
0.60

<2.0
0.51

1.28
7.6

Fe

0.25

0.19

5.87
9.3

0.15
0.11
0.08
<0.1
0.087
1.6
0.07
0.40

temperatures. The calculated activation energies provide some clarication of these


mechanisms, with varying levels of certainty. For example, for mechanically milled
Fe, an activation energy that is consistent with solute drag on migrating boundaries
is also consistent with a phenomenological model for that grain growth. On the other
hand, an activation energy for grain growth in cryomilled Ni that is very close to the
value for grain boundary self diusion of Ni would appear to be contrary to a more
detailed understanding of the presence of dispersoid particles and the role that they
may play in impeding grain boundary movement.
2.3. Grain size characteristics and distributions
The results for grain growth during isothermal annealing were based on average
grain sizes that were consistent with both XRD analysis and TEM observation. In
reality, the grain size of nanocrystalline materials is often more accurately described
by a distribution of grain sizes [63]. In the preceding discussion, such distributions
were presented in some cases [46,47]. Changes in the grain size distributions have
also been noted as milling progresses [1,34], and grain size distributions arising as
the result of consolidation of nanoscale powders have been mentioned, for example,
in Cu formed by compaction of inert gas condensed powders [64] and consolidated
cryomilled Fe10Al [45]. An inhomogeneous or polydisperse grain size distribution
may in fact be benecial to consolidated ultra-ne grained or nanostructured materials, contributing to improved fracture toughness or ductility without detracting
from high strength [6569]. The origins of such a heterogeneous microstructure have
been explored for cryomilled Al7.5Mg [32].
Noting the inhomogeneity of the as-milled microstructure may be a necessary
condition to understanding the microstructural evolution during annealing. Cryomilled Al7.5Mg powders were annealed at 150 and 250 C, corresponding to
roughly 0.50 and 0.64Tm [32]. DSC analysis of the as-milled powder indicated a release of energy due to recovery between temperatures of 100 and 230 C, so the selected annealing temperatures represent two dierent levels of recovery. The increase
in average grain size during annealing was not sucient to account for the total enthalpy associated with recovery. Even the recovery process itself could not account
for all the stored energy, despite the high dislocation density found in the severely

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

17

deformed as-milled powder. The balance of the energy release was attributed to the
relaxation of lattice strain and the reduction of the volume of non-equilibrium grain
boundaries. Perhaps more striking, however, were the microstructures resulting from
the annealing treatments. While the as-milled microstructures were dominantly equiaxed nanometer sized grains with some elongated grains of nanometer width and
containing nanometer scale sub-grains, the annealed powders contained clusters of
grains that were 200500 nm in diameter within the nanometer scale matrix, resulting in a bimodal, or duplex, overall microstructure. Also noted in powder annealed
at 250 C were coarse grained regions that had appeared to retain an elongated texture. These were described as recrystallized regions, but the bulk of the sample had
retained its as-milled microstructure, with an average grain size of 42 nm. The
authors suggested that the presence of a highly deformed, heterogeneous microstructure in the as-milled state would lead to regions of the sample that would recover and
recrystallize dierently, leading to a bimodal microstructure that had retained, in a
fashion, its heterogeneity. As will be discussed, this situation arises during the thermomechanical consolidation and processing of nanostructured powders, but this
study on Al7.5Mg serves to establish that duplex microstructures can arise during
thermal treatment of the powder in the absence of stresses that arise during consolidation. In addition, a kinetic analysis of DSC measurements showed that two different activation energies were associated with each of the two exothermic peaks,
with a lower value again calculated for the lower temperature peak (recovery) than
for the higher temperature peak. The activation energy calculated for the lower temperature peak was approximately 122 kJ/mol, which is comparable to the value for
Mg diusion in Al (136 kJ/mol), suggesting that the recovery process is controlled by
the diusion of Mg [32]. Such a scenario would be consistent with a grain rotation
and coalescence mechanism for grain growth.
2.4. Primary consolidation of cryomilled powder
In the early work, cryomilled Al was HIPped at approximately 0.84Tm [6], but the
as-HIPped microstructure was not described. Consolidation of cryomilled yttriated
NiAl was accomplished by both extrusion of canned powders and by HIPping
[13,14]. These early studies did not present descriptions of the microstructure of
the as-milled powders that would allow comparison to the primary consolidated
microstructure. Al cryomilled in the presence of added AlN particles (430 vol.%)
was consolidated by a combination of cold and hot pressing [53,54]. In this case,
coarser grains were noted in the consolidated material with 20.6 vol.% Al; coarse
grains were free of both the added AlN particles, which measured approximately
32 nm, as well as any dispersoids that might have formed during milling [53]. Cryomilled Fe and FeAl were consolidated by both warm compaction and HIPping. The
density of HIPped samples, which were essentially full dense, greatly exceeded that of
the compacted samples, while uniaxial compression only resulted in a maximum density of 85% within the temperature range of the apparatus [44].
The microstructure of Fe10Al after consolidation by warm compaction at
0.46Tm is similar to that reported for statically annealed cryomilled Al powders.

18

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Four regions were identied, including equiaxed grains less than 10 nm in diameter,
elongated grains measuring 100 nm long, 100 nm equiaxed grains, and large recrystallized regions with a grain size of 1 lm [45]. The rst two of these regions were considered unchanged from milling during consolidation, with the elongated grains
indicative of incomplete milling, while the two larger grain size regions were attributed to grain growth. For cryomilled Fe2.6Al, HIPping at 0.66Tm led to clustered
large grains (micron scale) within a ne grained matrix, which appears to be best described as sub-micron [45], a microstructure that closely resembles the annealed samples of cryomilled Al7.5Mg powders previously discussed. Chemical analysis of the
dierent regions of the Fe10Al compact using energy dispersive spectroscopy (EDS)
in the TEM also indicated partitioning of Al according to microstructure, with
18.8 wt.% Al in the nest grained regions, 16.6 wt.% in the elongated regions,
2.7 wt.% in the 100 nm equiaxed grains, and pure Fe in the largest grains. The nest
grain size region also contained signicant oxygen, 8.0 wt.% compared to 2 wt.% for
the overall material, but oxides were detected in selected area diraction patterns
from all regions except the Fe-rich large grained regions. This partitioning of Al after
thermomechanical consolidation was not addressed, other than to indicate that the
higher thermal stability aorded by the Al (Fig. 5) suppressed grain growth where Al
was present. This does not answer whether the Al was rejected from some regions
and concentrated in others, leading to the heterogeneities, or if the chemical heterogeneity was a function of incomplete milling and alloying. Chemical partitioning in
this material was subsequently investigated using Atom Probe Field Microscopy [70],
as discussed below.
The evolution of microstructure during the HIPping cycle has not been well documented. The as-HIPped microstructure has not been described with nearly the same
detail as the as-milled powders. Like the milling process itself, where the evolution of
particle size and morphology is paralleled by grain size renement, the consolidation
process also takes place at two length scales. At the microstructural level, the static
annealing studies of the milled powder were undertaken in part to evaluate favorable
temperature ranges for the consolidation of milled powders. These studies did not
include the possible grain growth that could arise due to the application of both
temperature and stress. On the other hand, modeling of the HIP process typically
focuses on the densication of the powders, and not the transformations taking place
within the powders [71]. The densication has been described as a three-step process,
beginning with loose packing of powder in the HIP can, elimination of connected
porosity due to the growth of necks at contact points between adjacent particles,
and reduction of the size of individual pores [72]. The growth of necks and lling
in of pores at each stage is accomplished by a combination of plastic deformation
of the powder particles, power-law creep, and mass diusion to the remaining free
surfaces. In the case of cryomilled powers, the powders cannot be considered either
spherical or monodisperse in size, so the rate equations developed for idealized powder populations are not likely to be accurate [72]. The qualitative model is nevertheless useful in one important respect, because where plastic ow of the consolidating
materials yields to diusive mass transport as a mechanism for densication, interparticle voids will be lled by material that no longer represents the as-milled condi-

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

19

tion. The result is that the HIP material resembles a two-phase composite material.
This result has been presented for cryomilled Al10Ti2Cu, an alloy designed to include a stable intermetallic phase (Al3Ti) in addition to the ne-grained fcc Al. The
HIPping process led to the formation of halos of an Al-rich phase surrounding the
powder particles that contained Al, Ti and Cu [73].
The two-component microstructure should be distinguished from the growth of
coarse grains within the cryomilled microstructure. The cluster of coarse grains in
annealed cryomilled Al7.5Mg powder [32] obviously results from grain growth or
recrystallization within the milled grains. For the coarse-grained cluster observed
in HIPped Fe2.6Al [45], the absence of additional information precludes denitive
understanding of the nature of the coarse grained region. Supporting information is
shown in Fig. 6 for cryomilled Al 5083 that was HIPped at a temperature of 300 C
(0.67Tm) and a pressure of 172 MPa. In Fig. 6a, an optical micrograph of the asHIPped material clearly shows contrast between the lighter colored halos surrounding the milled particles. A TEM image of a shared particle boundary (Fig. 6b) shows
the dierence in grain size between the particles themselves, for which the grain size
averages less than 100 nm, and the interparticle material, where the grains are
roughly 13 lm [74]. The presence of a number of dislocations, visible in the largest
grain in the interparticle region is possibly due to quenching of the sample at the conclusion of the HIP cycle. Based on the optical image in Fig. 6a, these coarse-grained
regions represent about 1020% of the sample, which is consistent with the results for
Al10Ti2Cu [73].
The grain size distributions of four dierent HIP billets of cryomilled Al 5083,
based on examination of TEM images, are presented in Fig. 7 along with their
respective optical microstructures. These billets are from three dierent cryomilling
runs. The HIP samples in Fig. 7a and b (H-1 and H-2) were taken from the same
batch of Al 5083 cryomilled in a laboratory-scale attritor (approximately 10 l capacity) with a 1 kg powder charge and a 32:1 ball-to-charge ratio. These two billets were
HIPped at 170 MPa and at temperatures of 300 and 350 C, respectively. The sample
in Fig. 7c (H-3) was also from laboratory-scale cryomill, but in a smaller attritor
(approximately 5 l capacity) and a 500 g powder charge at a 20:1 ball-to-charge ratio
and was HIPped at 300 C and a pressure of 186 MPa. The fourth HIP billet (H-4)
was produced from a commercial-scale attritor and a 20 kg powder charge. This
powder was charged into a single large can and HIPped according to a proprietary
HIP cycle at an approximate temperature of 300 C. Prior to HIPping, all canned
powders were vacuum degassed at approximately 400 C to a vacuum of roughly
106 Torr. The grain size distributions in Fig. 7 only plot the grain size of the
ne-grained regions,1 and indicate that the microstructure in the cryomilled region
of all four is qualitatively similar. The optical images accompanying each sample,
however, show dierences between them. In two of the samples, shown in Fig. 7e
and f, the prior particle boundaries are still intact. In the second of these, unclosed
1

TEM investigation of the samples was sucient to qualitatively assess the grain size of these coarsegrained regions but their size relative to the transparent region of individual foils made statistical grain-size
counting dubious compared to the ne-grained regions.

20

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Fig. 6. Micrographs of as-HIPped Al 5083: (a) optical micrograph showing intact milled particles
surrounded by halos of Al and (b) TEM image showing the interface of ultra-ne grained regions of
several particles and the coarse grained region between them.

pores remain at the center of the interparticle regions. In comparison, the HIP samples in Fig. 7g and h show evidence of much greater deformation of the particles
themselves. While the light-colored regions are still apparent in similar proportions
to material shown in Fig. 7e and f, individual powder particles are no longer intact
and the network of coarse grained regions is irregular.

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

21

Fig. 7. (a)(d): Grain size distributions of as-HIPped and extruded cryomilled Al 5083 and (e)(h): optical
images of the samples for each of the grain size distributions.

The grain size of the as-HIPped material may be considered in comparison to the
grain size of the as-milled powders. The grain size of as-milled powders was previously found to be approximately 25 nm for Al and AlMg [40,47], based on XRD
and TEM measurements. The same batch of cryomilled powder was used for HIP
materials whose grain size distributions are shown in Fig. 7a and b. XRD data from
this powder batch, analyzed using the same method as for the previous AlMg powder [40], had a grain size of 22 nm. The annealing data in [40] suggested grain growth
at the degassing temperature to about 65 nm. The histograms in Fig. 7a and b indicate that the modal average of the grain size was between 50 and 100 nm, suggesting

22

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Fig. 7 (continued)

that degassing alone might have been responsible for grain growth, although the
mean grain sizes based on these distributions were 113 76 nm and 85 49 nm,
respectively. While the powders for the other two HIP samples were not characterized, the mean grain sizes of H-3 and of the commercial scale billet H-4 were
129 75 nm and 139 107 nm, respectively. In these two samples the largest grain
size fraction was between 50 and 100 nm. The occurrence of larger grains (>200 nm)
shifts the mean average away from the modal average grain size, but the population
of grains between 50 and 100 nm is consistent with the ndings of isothermal annealing experiments.

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

23

Fig. 7 (continued)

2.5. Processing of primary consolidated forms


Powder metallurgy products must be fabricated in such a way as to break up prior
particle boundaries, homogenize the material, and close remaining pores to eliminate
possible stress concentration points that can serve to initiate cracks or lead to failure
by interface debonding. As illustrated in Fig. 7, HIPping may not be sucient to
achieve these goals, necessitating additional processing. The rst reported cryomilling of AlAl2O3 used a combination of HIP and extrusion [6], while an initial eort
to cryomill NiAl with 0.5 wt.% Y2O3 used HIPping or extrusion of canned powders
at three dierent reduction ratios [13,14]. Both extruded and HIPped products

24

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Fig. 7 (continued)

comprised two separate microstructural regions, containing areas rich in yttria and
AlN particles and particle-free NiAl areas. In the HIPped samples, the coarse particle-free NiAl regions exhibited a bimodal distribution, with respective grain sizes of
approximately 6 and 1 lm and were arranged in roughly spherical sub-regions. In
the extruded material the particle free regions assumed a more continuous grain size
distribution around a mean grain diameter transverse and grain length parallel to the
extrusion direction [13]. The aspect ratio of the grains was commensurate with the
extrusion reduction ratio. Compression tests and compressive creep tests conducted
with the stress axis longitudinal to the extrusion direction suggested limited inuence

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

25

of processing on mechanical properties [13,17]. Subsequent cryomilling of NiAl without Y2O3 [19] indicated that the presence of a hard second phase was not responsible
for the microstructural inhomogeneity. This was conrmed by cryomilling either
with or without Zr, both of which resulted in a similar microstructure [18].
For cryomilled materials that have included both a primary consolidation step,
such as HIPping, followed by extrusion, the microstructural evolution can be assessed even further. In an earlier study of cryomilled Al 5083, the average grain size
in both HIPped and extruded forms was stated to be approximately 40 nm [66]. The
room temperature tensile behavior of this extruded material diers signicantly from
and is inconsistent with that of subsequent cryomilled extruded Al 5083 and Al
7.5Mg with larger grain sizes [74,75], placing in doubt the interpretation of the grain
size in the earlier work. Comparison of HIPped and extruded microstructures was
made for cryomilled Al10Ti2Cu, but that focused more on the alteration of
coarse-grained regions that arose during HIPping and surrounding the powder particles [73,76]. Extrusion resulted in the elongation of coarse grained regions in the
extrusion direction, similar to the behavior exhibited by the cryomilled NiAl [13].
Grain size distributions for extruded cryomilled Al better show the range of grain
sizes that arise during processing. For material that was HIPped at 300 C
(0.6Tm) and extruded with a 9:1 reduction ratio at 325 C, the average grain size
was 111 nm, with a standard deviation of 77 nm [77]. Cryomilled commercial purity
Al taken from a single cryomilling batch and subjected to HIPping temperatures of
275, 300 and 325 C, before extrusion with a reduction ratio of 8:1 at 260 C had
average grain sizes that correlated with its HIPping temperature: 116, 162 and 179
[78], respectively, although these dierences fall within a single standard deviation.
The grain size distributions for these three extrusions were similar, but higher HIP
temperature was correlated with a larger number of relatively coarse grains (300
1000 nm) [78]. The qualitative dierence in microstructures was that at higher HIP
temperatures, coarse grains were more abundant and were more likely to be found
in clusters. The grain size distributions shown in Fig. 8 indicate similarity in microstructure for the four cryomilled Al extrusions. In comparison, the grain size distributions for cryomilled Al 5083 plotted in Fig. 8 exhibit greater variation.
A complete description of the microstructural evolution from HIPped to extruded
form would include discussion of not only the changes in large features such as the
coarse-grained interparticle regions, but also the changes in microstructure within
the milled and consolidated powder particles. Fig. 9a shows the result of extrusion
on a laboratory-scale cryomilled Al 5083 sample in which the powder particle
boundaries remained intact during HIPping (H-1), while Fig. 9b shows the microstructure of an extrusion from a commercial-scale HIP billet (H-4) in which the particles were deformed more substantially by the HIPping process. Optical
micrographs for the HIP precursors of these extrusions are shown in Fig. 7e and
h, respectively. The extrusion reduction ratio was 6.5:1 for both, and the extrusions
were performed at relatively low temperatures (0.50.60Tm) under isothermal conditions that were maintained in part by using low ram speeds (strain rates on the
order of 103 s1). In the laboratory-scale extrusion shown in Fig. 9a, the negrained regions were elongated, but a regular network of coarse-grained bands

26

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Cumulative Frequency

0.75

Al 2a
Al 2b
Al 2c
Al [77]

0.5

H-1
H-3
H-4a
H-4b

0.25

00
>6

60

0
50

0-

50

0
45

0-

45

0
40

0-

40

0
35

0-

35

0
30

0-

30

0
25

0-

25

0
20

0-

20

0
15

0-

15

00
-1

010

50

<5

Grain Size Interval (nm)


Fig. 8. Cumulative frequency of grain size in cryomilled HIPped and extruded Al and Al 5083.

appear as lamellae that continue to dene the prior particle boundaries. In the commercial material, the coarse-grained regions are also present, but are not in obvious
coincidence with prior particle boundaries (Fig. 9b). The coarse regions in the commercial-scale extrusions are thinner and less prevalent than those in the laboratoryscale extrusions. In Fig. 10, a TEM image of a coarse-grained region in an extrusion
from a laboratory-scale cryomilling experiment shows that it is approximately
500 nm wide, separating two regions of ne-grained material. At the right-hand side
of the image, the band broadens as it approaches the end of an elongated powder
particle. Detailed investigation of several dierent examples of the coarse-grained
bands suggest that when they are very thin, they consist of a series of sub-grains
whose width encompasses the entire width of the band, and whose length in the
extrusion direction is approximately the same as their width. In the thicker regions
of the bands, the individual grains or sub-grains no longer span the entire width.
Rather, the band consists of multiple grains with dimensions usually falling between
500 and 1000 nm. Unlike the ne-grained regions, where the grain boundaries tend
to be irregular, the coarse-grained regions often show evidence of sharply dened
polygonal boundaries, consistent with recrystallization.
Grain size frequency distributions for the extruded Al 5083 are plotted together
with the respective HIP precursors in Fig. 7. In laboratory-scale extrusions, TEM
micrographs indicated elongation of grains consistent with the extrusion direction.
For this reason, both transverse and longitudinal grain dimensions are plotted in
the histograms. While not all grains in the commercial-scale cryomilled extrusions
or in the HIPped materials could be considered equiaxed there was insucient texture to correlate the long axis of the grains with the extrusion direction. The mode of
the grain size distribution for the laboratory scale extrusions was larger than for the
commercial scale extrusions. The mean grain size of the extruded material relative to

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

27

Fig. 9. Optical micrographs of extruded cryomilled Al 5083, sections oriented parallel to extrusion
direction. The image in 10a is for extrusion from HIP precursor shown in Fig. 7e, while that in 10b is
extruded from the HIP precursor in Fig. 7h.

its respective HIPped precursor varied according to the sample, as shown in Table 4.
This ratio was 1.3 and 1.7 for the two laboratory scale materials extruded at a ratio
of 6.5:1, while it was 2.5 for the laboratory scale material extruded at a ratio of 9:1
and for each of the commercial scale extrusions. The cumulative frequency diagrams

28

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Fig. 10. TEM BF image of coarse-grained band region in extruded material.

Table 4
Summary of average grain sizes
Sample

Processing

HIP grain
size (nm)

Extrusion grain size (nm)

H-1
H-2
H-3
H-4a
H-4b

Lab. Scale (6.5:1)


Lab. Scale (6.5:1)
Lab. Scale (9:1, 325 C)
Commercial Scale, Ext. 200 C
Commercial Scale, Ext. 225 C

113 76
85 49
129 75
139 107
139 107

148 63 (43%) Long. 120 58 (48%) Trans.


147 81 (55%) Long. 95 47 (49%) Trans.
308 156 (51%) Long. 215 111 (52%)
220 109 (50%)
235 159 (68%)

in Fig. 8 show that laboratory scale batches of Al and Al 5083 processed in the same
attritor and having similar consolidation and extrusion conditions had similar grain
size distributions regardless of mean grain size. The laboratory scale cryomilling
batch that was processed in a smaller attritor with a lower ball:charge mass ratio
(H-3) deviated from this grouping, but it was also extruded at a higher temperature
and higher extrusion ratio. Extrusions derived from the commercial scale cryomilling
batch of Al 5083 are displaced somewhat from both the Al and Al 5083 from the
laboratory attritor.
The importance of a complete description of the microstructure must be emphasized. The mechanical behavior of nanostructured metals and alloys, and hence their
performance and potential applications, may be best understood in terms of the full

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

29

range of grain sizes, and not simply an average grain size [63,69]. Furthermore,
experimental evidence for materials processed by means other than cryomilling indicates that the presence of occasional coarse grains may play a critical role in the
mechanical properties [67,68,79]. For consolidated materials, at least, reported grain
sizes based on XRD data are clearly insucient to describe the microstructure. In
the case of cryomilled materials, the average grain size determined by XRD corresponded to the smallest grain size in the distribution based on TEM observations
[77].
The grain size distribution of extruded cryomilled materials diers considerably
from the narrow distributions reported for cryomilled powder, which had mean
grain sizes ranging from 20 to 30 nm for Al [34] and AlMg [32]. The nal grain size
in consolidated materials is the result of degassing, consolidation and extrusion, but
the question remains as to what determines the nal grain size. On the basis of thermal analysis and isothermal annealing on cryomilled Al7.5Mg, recrystallization was
found to occur between approximately 280 and 370 C [32], which is below the degassing temperature of the as-milled Al 5083. Therefore, the transformations that occurred on annealing the material would have been completed during the degassing
cycle. On the basis of this and other static annealing experiments of cryomilled
AlMg [40], the average grain size of degassed cryomilled Al 5083 would be expected
to be between 50 and 100 nm, with occasional coarser grains and coarse grain regions present. The HIP cycle for the Al and AlMg materials discussed herein also
falls in this temperature range, but the isothermal extrusion occurred at temperatures
below that of recrystallization.
To understand how each thermomechanical processing step changes the microstructure, the role of dispersoids must also be considered. The dispersoid population
has a dierent character, especially in terms of size distribution, for the dierent systems studied, such as Al [6], NiAl [12] and Zn22Al [49]. In using the grain size data
for cryomilled Al 5083 as a case study, the description of dispersoids given for the
rst cryomilled AlAl2O3 is adopted, namely small platelets measuring a few atomic
layers thick and 10 to 15 nm in diameter [23], as TEM evidence does not suggest the
presence of larger dispersoids. The estimated spacing of these dispersoids was based
on the nitrogen content of the extruded material [6]. In addition, as an aluminum
powder metallurgy product, stable surface oxides will be broken up during the milling process, and incorporated into the cryomilled powders, as occurs with mechanically milled Al [10]. Oxides may also reform on the as-milled powder during
handling, thus coating prior particle boundaries during consolidation [80]. These
oxides are no doubt present as well, and must be considered in evaluating the microstructural evolution.
Previous work on cryomilled fcc Ni [43] suggests that Zener pinning shows good
agreement with the grain size in annealed powders, according to the equation
d

4r
;
3f

where d is the grain size stabilized by particles, r is the radius of pinning particles,
and f is the volume fraction of the particles. The average combined nitrogen and

30

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

oxygen content of the cryomilled Al 5083 extrusions discussed above was 0.55 wt.%
and 0.4 wt.%, respectively. Based on these values, the volume fraction of cryomilling
dispersoids can be approximated to be between 0.01 and 0.02, leading to a particle
stabilized grain size between 700 and 2000 nm. This assumption makes no distinction
between the pinning role of oxide or nitride particles. Another way to evaluate the
grain size on the basis of dispersoids is to consider the distance between particles
based on their size and volume fraction. For plate-shaped particles, the following
relationship holds [81]:
s
D t
DA
;
5
f
where DA and D are the particle spacing and particle diameter, respectively, and t is
the particle thickness. A dispersoid thickness of 12 nm gives an interparticle spacing
of 2245 nm based on Eq. (5). In the earliest reports of the cryomilling technique,
Luton et al. [6] estimated that the spacing of the Al oxynitrides in cryomilled
AlAl2O3 was 80 nm based on TEM evidence, in close agreement for the value calculated from Eq. (5) using their particle volume fraction and the same particle
dimensions. The grain size for this material was reported as 50300 nm, well below
the value calculated based on Zener pinning. High-resolution TEM on this same
material by Susegg et al. [23] showed several dispersoids in random orientation in
clusters within which the center-to-center separation of individual particles is closer
to 20 nm. This result indicates that Luton et al. [6] may have observed clusters of
particles separated by 80 nm, not individual dispersoids. In this case, the wide discrepancy between the actual grain size and the grain size predicted by Zener pinning
can be rationalized in part by noting the close correspondence of the grain size and
the interparticle spacing. Zener pinning is predicated on the assumption of a planar
grain boundary interacting with a particle, which is valid when the grain size is much
larger than the particle spacing [62], so an interparticle spacing of 80 nm and a grain
size range from 50 to 300 nm would suggest that the Zener pinning approach, an
approximation at best, is not applicable over the entire microstructure.
For grain sizes that are closer in dimension to the interparticle spacing, Humphreys and Hatherly have suggested a model in which the stable grain size occurs when
all grain corners are pinned [62]. Such an approach is consistent with Luton et al.s
[6] reported grain size distribution and observed interparticle spacing. Given the clustering of dispersoids observed in this material [23], a range in grain sizes is not surprising. This suggests that the dispersoids are indeed found at grain boundaries.
APFIM has been used to investigate the chemical composition of grain interiors
and grain boundaries in cryomilled Fe10Al (Fe5.1at%Al) consolidated by HIPping [70]. The results showed dramatic dierences in both the alloying element Al
and dispersoid elements O and N. The APFIM data, given here in Table 5, showed
that nearly all the oxygen and nitrogen was segregated to the grain boundaries, and
that Al had been partitioned to the grain boundaries relative to the grain interiors.
The APFIM data lend considerable weight to the notion that the dispersoids are
responsible for grain size stability, but their uneven distribution throughout the

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

31

Table 5
APFIM data for cryomilled Fe5.1 at%Al [70]
Measurement location

At.%Fe

At.%Al

At.%N

At.%O

Grain boundary 1
Grain boundary 2
Grain interior

66 1.3
38 1.3
97 0.33

6.5 0.66
22 0.66
2 0.27

5.3 0.61
16 0.61
0.5 0.14

22 1.1
24 1.1
0.7 0.14

material makes prediction of the average grain size based strictly on the chemical
composition with respect to N and O dubious.
For the Al 5083, both the HIPped and extruded grain sizes lie between the calculated interparticle spacing (from Eq. (5)) and the grain size predicted by Zener pinning. The shortcomings of Zener pinning to predict the grain size have been noted,
but in addition to geometric considerations of grain boundaryparticle interactions,
the Mg solute will also play a role in grain growth [32,82,83], particularly by reducing grain boundary mobility. In addition, the clustering of dispersoids may lead to an
eective interparticle distance that exceeds the value predicted by Eq. (5). High-resolution TEM is needed to clarify this matter, because the small size of the dispersoids
and their presumed segregation to grain boundaries renders their imaging dicult at
best by means of conventional TEM.
In determining the microstructural evolution from as-HIPped to extruded, the situation changes from a combination of high thermal driving force and isostatic stress
to one of lower applied stress and temperature in combination with shear stresses.
After extrusion, the average grain size still remains less than that predicted by the
Zener pinning relationship (Eq. (4)). Insucient understanding about the occurrence
of the cryomilled dispersoids, among other things, makes it dicult based on the
present TEM evidence to produce a compelling model for grain growth. The arguments advanced concerning Luton et al.s [6] material were based on the extruded
material, although they too used HIPping as a primary consolidation technique.
For the Al 5083 in the as-HIPped form, the microstructure could be described by
one of three possible scenarios. First, the dispersoids are coincident with grain
boundaries, a case which requires clustering of dispersoids. Second, the spacing of
dispersoids is more regular (not clustered) and hence less than the grain dimensions,
and the grain boundaries are stabilized by Mg. Third, clustering of disperoids leads
to the clusters having a larger spacing than the as-HIPped grain size, and again
movement of grain boundaries is hindered by Mg.
During extrusion, increasing the average grain size by 130250% from the average
HIP grain size for the various extrusions would play out dierently in each of these
scenarios. In the rst scenario, the grain boundaries must break free of the network
of dispersoids, which would remain in place in the interior of the new, larger grains,
or would themselves migrate during extrusion. In the second and third, the role of
Mg must be examined. The capacity of ne-grained materials for accommodating
an unusually high percentage of solute elements may be explained by the increased
grain boundary volume at lower grain sizes; for a 100 nm material with grain boundaries 1 nm thick, nearly 5% of the atoms would be associated with grain boundaries

32

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

[84]. Cryomilling of Al7.5 wt.% Mg led to a super-saturated solid solution of Mg


[32,40]. Measurements of the lattice constant of cryomilled Al7.5Mg powder after
annealing at dierent temperatures indicated that the amount of Mg in solution decreased for annealing at 150 and 250 C, but was apparently stable at higher annealing temperatures [32]. In this case, the annealed material was powder with a starting
grain size of 25 nm, so that Mg rejected from the lattice could be readily absorbed by
abundant and nearby grain boundaries. While some of the Mg in the HIPped material is within the grain boundaries, the generally high Mg mobility in Al, especially at
temperatures in the range 150250 C (which includes the extrusion temperatures
used) would allow grain boundaries to migrate and lead to grain growth. This situation favors the third scenario, in which the spacing of dispersoid clusters is larger
than the as-HIPped grain size, and the nal extruded grain size reects the nature
of the dispersoid network.
Comparison to other aluminum alloys lends credence to such an interpretation. In
one example, the extrusion of aluminum metalmatrix composites typically leads to
clustering of secondary particles [85]. Perhaps a more analogous material is mechanically alloyed (MA) Al, which contains a highly deformed microstructure as well as Al
oxide and carbide dispersoids. The microstructure of an extruded MA Al described by
Hawk et al. [86] is similar in many respects to that of the cryomilled AlMg. The overall grain size of these alloys was 250500 nm, and dispersoids were found primarily in
grain boundary regions, although they were not homogeneously distributed throughout the material. Some grain boundaries had higher concentrations of dispersoids,
and some grain interiors contained some dispersoids, while other grains were free
of dispersoids. In addition, larger grain sizes were observed for alloys with lower volume fractions of dispersoids. In addition, regions of large elongated grains were
found, and these regions had a relatively low concentration of dispersoid particles.
The core and mantle microstructure of cryomilled NiAl is a more dramatic example, as relatively large regions of apparently particle-free regions arose during HIPping or extrusion [12,13]. A dierent interpretation of grain growth emerges from
study of Al3Mg processed by Equal Angular Channel Pressing (ECAP). As has been
described for cryomilled powders, as-ECAPed Al3Mg had a heterogeneous microstructure, consisting of both elongated grains and equiaxed grains [38]. Annealing
at intermediate temperatures led to a duplex microstructure consisting of coarse,
recrystallized grains within a ne grained (sub-micron range) matrix. The rst appearance of the coarse grains was at 503 K, and the coarse grained fraction continued to
grow in average grain size, while the ne-grained matrix remained stable, up to 550 K,
whereupon the entire sample became coarse grained. This duplex microstructure was
attributed not to heterogeneous distribution of second phase particles, however, but
to the presence of low-angle grain boundaries that were eliminated during annealing
by dislocation recovery and annihilation [87]. Such a mechanism might still be operative in cryomilled materials as an explanation for the occurrence of coarse grains that
contain dispersoids [75], without detracting from the dominant role that dispersoids
play in regulating the microstructure.
In the cryomilled alloy, an interparticle spacing of the dispersoids that is on the
order of the grain size and where the dispersoids stabilize the grain boundaries, clus-

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

33

tering of particles during extrusion would necessarily contribute to the both the increase in grain size as well as the broader distribution of grain sizes. The clustering of
dispersoids, as well as the general inhomogeneity of dispersoid distribution, will lead
to regions of lower volume fraction of dispersoids, for which Eq. (4) predicts a larger
particle-pinned mean grain size. In addition, the conditions for abnormal grain
growth can be generalized as occurring when the pinning eects of particles have arrested the process of normal grain growth [62]. There are thus two possible mechanisms for the coarsening observed during extrusion based on the presence and
distribution of dispersoids: grain boundary pinning by dispersoids in an arrangement
that has been changed by extrusion, and abnormal grain growth. Further evidence
for the rst of these was suggested by APFIM of cryomilled and extruded Al
5083, in which Al nitride and oxide dispersoids were found primarily at grain boundaries, and the radial axis of the dispersoid platelets was coincident with the extrusion
axis [66]. In combination with mobilization of Mg solute during extrusion, the nal,
as-extruded grain size reects a microstructure in which the Mg and the dispersoids
stabilize the grain boundaries, and not surprisingly this microstructure is fairly
complex.
In this discussion, a second component of the microstructure has not been discussed, namely the coarse-grained regions that arise during HIPping and are elongated parallel to the extrusion direction. These regions, which as previously
described represent the lling in of interparticle voids, consist of grain sizes of
approximately 12 lm in the HIPped material and roughly 500 nm (grain or subgrain) in the extrusions (Fig. 10). The coarse-grained region has a grain size roughly
10 to 20 times that of the UFG region, and probably contains signicantly fewer
cryomilling dispersoids, so it can be surmised to have lower strength. Given this situation, Al-MMCs give some context for the nature of the deformation taking place
during extrusion of the HIPped cryomilled materials. For an Al-MMC extruded at
low strain rates, the deformation will be carried preferentially by the particle-free regions [85]. The strain rate employed to extrude the Al 5083 discussed here was sufciently low in order to maintain isothermal conditions within the billet. The softer
coarse-grained regions will be preferentially deformed during extrusion, like the
metal matrix in a particulate MMC. One of the obvious drawbacks of this scenario
is that it makes it more dicult to disrupt prior particle boundaries, as has been observed in other cryomilled Al alloys [73] and in the laboratory-scale samples for cryomilled Al 5083, which would be expected to have a deleterious eect on mechanical
properties [80]. In the commercial-scale billet, however, the HIPping process itself
was eective in disrupting much of the particle structure, so that the coarse-grained
regions form an irregular array, not a network.
In both commercial- and laboratory-scale Al 5083 materials discussed here, the
dimensions of the coarse regions in the HIP and extruded materials were consistent
with the extrusion ratio. These softer regions were elongated to the same extent as
the macroscopic change in the billet. In contrast, grains in the UFG matrix exhibited
aspect ratios of less than 2 for the materials extruded at a ratio of 6.5:1 at a low temperature, while the higher temperature, higher extrusion ratio (9:1) material (H-3)
had a grain aspect ratio of 2.5. The size distribution of the grain dimension measured

34

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

transverse to the extrusion direction is more similar to the HIP material than the longitudinal dimension. This means that the grains are not being elongated to the same
extent as the billet as a whole, and indicates grain growth may proceed transverse to
the extrusion direction. A possible scenario is that grains are aligned during extrusion, and low-angle boundaries parallel to the extrusion direction are consumed,
leading to grains that are slightly longer and wider than the as-HIPped grains.
While the individual extrusions are heterogeneous on a microscopic scale, the
microstructure is consistent over the whole of the extrusion. During the extrusion
process, there are spatial variations in strain rate that depend on the semi-angle of
the die, reduction ratio, and lubrication [88]. Despite the inferred inhomogeneous
distribution of deformation within dierent components of the extrusion, microhardness indentation testing in transects across both the transverse and longitudinal faces
of the extrusions did not indicate any spatial variation in Vickers Hardness, as
shown in Fig. 11a, which graphically presents hardness measurements made from
the center of to the edge of two cylindrical extrusions of approximately 33 mm diam-

Fig. 11. Hardness values from transects measured on transverse and longitudinal sections of circular
extrusions (a); schematic of rectangular extrusion, divided into six dierent faces where hardness was
measured was measured along transects. The average hardness value of each face is given at the bottom of
the diagram (b).

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

35

eter. The transverse hardness values were slightly higher in each extrusion. The same
size HIPped billet that was extruded through a circular die with diameter 33 mm at
low strain rate was also extruded at a higher temperature and strain rate through a
rectangular die with an approximate cross-section of 22 53 mm. The strain-rate
eld in this extrusion was more complicated than for the circular die. Like the circular extrusions, the rectangular extrusion did not exhibit spatial hardness variations in
transects measured in dierent directions in dierent places within the longitudinal
and transverse cross-sections. The locations of these transects are shown in Fig.
11b, along with the average Vickers Hardness values for each region. These data
show that at the scale of the indentation, which exceeded the grain size by approximately two orders of magnitude, the variations in strain rate did not signicantly
aect the mechanical properties.

3. Mechanical properties of cryomilled materials


3.1. Overview of mechanical properties
The rst two cryomilled systems reported in the open literature were AlAl2O3 [6]
and NiAl [14]. For the former, the initial report of mechanical properties included
compression testing of the as-HIPped material and compression and tensile testing
of the material after extrusion and swaging [6]. At room temperature, the yield stress
of the as-HIPped material in compression was nearly 40% higher than the extruded
material in tension; at temperatures around 400 C this situation was essentially reversed. Creep testing at 400 C indicated that the threshold stress for creep was the
same as the yield stress in tension, or approximately 130 MPa. Additional tensile
testing indicated that the Al cryomilled with 3 vol.% alumina had both maximum
strength and ductility at room temperature [89].
Room temperature testing was not reported for NiAl, but initial compression testing conducted between 1200 and 1400 K (0.630.73Tm) indicated that the cryomilled NiAlY2O3 was six times more creep resistant than conventional NiAl and
twice as strong as NiAl reinforced by 10% TiB2 [13,14]. Subsequent study indicated
that the yttria made only a limited strengthening contribution to the material, and
investigators concluded that alloying elements would be needed to exceed the properties of conventionally processed superalloys based on creep rupture behavior [19].
To date, this latter issue has not been successfully addressed [16].
More recent work on cryomilled Inconel 625 following HIPping and extrusion directly compared it to wrought Inconel 625 [90]. The cryomilled alloy had a signicantly smaller grain size, with grain sizes ranging from 200 nm to 10 lm compared
to the conventional alloys grain size of 200 lm. Notably, the cryomilled material
contained a second phase in the form of ne Ni3Nb particles. These were primarily
associated with the ner grains in the microstructure and were absent from the conventional alloy. In this sense, the cryomilled Inconel 625 bears some similarity to the
core and mantle structure of the NiAl. The tensile behavior indicated that while both
cryomilled and conventional IN 625 exhibited a similar trend in yield and ow stress

36

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

as a function of temperature, the cryomilled alloy maintained a consistent improvement of 400600 MPa over the wrought Inconel over the temperature range of 0.19
0.58Tm. Creep testing was performed over the upper end of this temperature range,
and two dierent creep regimes were identied: a low-stress structure-controlled
creep regime dominated by dislocation climb and a higher stress region controlled
by viscous glide.
Cryomilled Al and its alloys have been subjected to more extensive room temperature testing than the Ni-based superalloys, a reection of anticipated operation temperatures. Characterization of the mechanical properties of these alloys, including
cryomilled Al [6,78,89], Al10Ti2Cu [63,9193], AlMg [75,94,95] and Al 5083
[66,74], invariably discusses their deformation behavior with respect to grain size.
While limited high-temperature deformation behavior has been reported for Al,
the existing data suggest similarity to mechanically alloyed Al. From elevated temperature tensile testing of Al7.5Mg0.3Sc, the behavior of the cryomilled material
was deemed inferior to superplastic alloys with respect to uniform deformation [96].
In the case of creep of cryomilled Al4Mg, high creep resistance and stability of the
microstructure during creep testing were both attributed to the presence of cryomilling dispersoids [95,97]. Interaction between grain boundary dislocations and nanometer particles were thought responsible for the creep resistance, while the particle
themselves stabilized the grain boundaries.
3.2. Room temperature deformation: strengthening mechanisms of cryomilled Al
Perhaps the fundamental reason for attempting to reduce the grain size to the
nanocrystalline region is the presumed gain in yield strength (ry) based on the
HallPetch equation:
ry r0 k y d 1=2 ;

where r0 is the lattice friction stress, ky is the HallPetch slope, and d is the mean
grain size. Two cryomilled systems for which grain size variation has been achieved,
allowing HallPetch plots, are commercial purity Al and Al 5083. Average grain
sizes for cryomilled extrusions are in the range 100200 nm for the Al [78,98] and
from approximately 150 to 300 nm for the Al 5083, as previously shown in Table
4. The yield strengths and ultimate tensile strengths of these materials are plotted
in Fig. 12 against the inverse square root of the p
grain size. For the cryomilled Al,
the calculated HallPetch slope ky was 0.09 MPa m [78]. These Al extrusions were
derived from a single batch of cryomilled commercial purity powder, and their subsequent consolidation was varied only by the HIP temperature. Despite the overall
similarity in grain size distributions for these three extrusions, higher HIP temperature correlated with a larger number of relatively coarse grains (3001000 nm) and a
lower strength [78]. As previously discussed, the qualitative dierence in microstructures was that at higher HIP temperatures, coarse grains were more abundant and
were more likely to be found in clusters. The chemical compositions of these dierent
extrusions are given in Table 6, highlighting specic species. The values for Al are
average compositions of ve extrusions derived from a single batch of cryomilled

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

37

800

Yield Stress (MPa)

700

Slope = 0.28 Mpa m^1/2

600
500
400
300

Slope = 0.09 Mpa m^1/2

200
Al

100

Al 5083

0
1500

1700

1900

2100

2300

d-1/2

2500

2700

2900

3100

(m-1/2)

Fig. 12. HallPetch plots for tensile testing of extrusions of cryomilled Al and Al 5083.

Table 6
Chemical composition of cryomilled Al and Al 5083
Extrusion

C (wt.%)

O (wt.%)

H (ppm)

N (wt.%)

Fe (wt.%)

Al 1
Al 2
H-2
H-4

0.25
0.20
0.20
0.17

0.93
1.54
0.45
0.38

425
232
26
18

0.37
0.38
0.71
0.40

0.12
0.05
0.33
0.21

powder. Fe content can be considered a proxy for contamination by the milling


tools. Carbon is most likely derived from reaction of the Al with the carbon in stearic
acid, while H may be from either this source or from contamination by water vapor
or hydration of the stable oxide layer on the powder particles. Oxygen is derived
from a combination of break up of the oxide layer on the pre-milled particles during
milling and reformation of a new layer on the particles after milling, despite eorts to
minimize exposure during handling of the powder. Lastly, N is considered to be the
reaction product of Al and N in the milling environment. There is no obvious basis
in these chemical data for the dierences in yield stress and tensile strength.
Stressstrain curves for two extrusions of cryomilled commercial purity Al with
average grain sizes of 116 nm and 179 nm are shown in Fig. 13. These curves show
only the extent of uniform plastic deformation; after the onset of necking, these
extrusions reached reductions in area of more than 20 and 50%, respectively [78].
These results hint that greater elongations can be achieved if ow localization can
be prevented. Flow localization in ne-grained materials is generally a challenge,
as high strengths must be matched by high work hardening rates according to Consideres criterion [99].

38

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160


450
400

True Stress (MPa)

350
300
250
200
150
100
HIP Temp 275 C
HIP Temp 350 C

50
0
0

0.01

0.02

0.03
True Strain

0.04

0.05

0.06

Fig. 13. Room temperature engineering stressstrain curves for extrusions of cryomilled Al at a strain rate
of 103 s1. The data have been displayed to the extent of uniform plastic strain.

Direct comparison of the mechanical properties of the cryomilled Al to other negrained Al is complicated by dierences that arise due to material preparation. For
example, the yield strengths presented for cryomilled Al greatly exceed mechanically
milled Al with grain sizes of 2140 nm, determined by XRD [100], which p
had yield
strengths between 17 and 26 MPa and a HallPetch slope of 0.012 MPa m. Processing consisted of cold pressing the as-milled powders to 9598% of theoretical
density [100]. While few experimental details were reported, and no chemical composition data, it is unlikely that the dierence in strength between these materials and
the cryomilled materials can be explained by the presence of dispersed Al2O3 alone.
The cryomilled Al contained the equivalent of approximately 1.62.6 vol.% Al2O3,
assuming all measured oxygen is present as alumina. For sintered aluminum product
(SAP), in which Al2O3 ranged from 0.2 to 14.2 vol.%, the increase in yield strength
was found to be approximately 12.5 MPa per volume percent Al2O3 [81,101]. The
yield strength of mechanically alloyed Al was found to correlate well with volume
percent dispersoid content, which varied by the type and volume of process control
agent employed during milling [10], but the slope of the relationship was more than
four times greater than for SAP. In contrast, the cryomilled Al from the same batch
shows a weak correlation with dispersoid volume content, albeit over a relatively
narrow range of volume percent, as shown in Fig. 14. The relative yield strengths
of dierent ne grained Al are presented together in this gure, including cryomilled
Al [78], mechanically alloyed Al [10], SAP [81], and nanocrystalline Al prepared by
plasma evaporation and consolidated by cold compaction followed by sintering
[102]. The carbon content of the mechanically alloyed Al was reported as elemental

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

39

Tensile Strength (MPa)

500

400

300

200

Cryo Al
M.A. Al [10]
Plasma Al [102]
SAP [81]

100

6
8
C+O Dispersoids (vol. %)

10

12

Fig. 14. Relationship between yield stress (0.2% oset) and calculated dispsersoid volume content for
cryomilled Al [78], mechanically alloyed Al [10], sintered aluminum product (SAP) [81], and nanocrystalline Al (grain size 53 nm) prepared by plasma evaporation, compaction and sintering [102].

carbon based on chemical extraction; for the cryomilled Al, the total dispersoid volume is based on the assumption that the carbon is the form Al4C3, with a density of
2.36 g/cm3. In comparison, the strengths obtained by Bonetti et al. [100] for mechanically milled Al were much lower, suggesting signicant aws in the samples.
The cryomilled alloys had a yield stress roughly 15% higher than the mechanically
alloyed Al [10] at a similar fraction of dispersoid particles, except for the cryomilled
Al with the largest dispersoid volume. The dierences between these two sets of
materials were the presence of AlN in the cryomilled alloys and the average grain
size, which is smaller for the cryomilled Al, although both share a generally similar
heterogeneous microstructure. Both strengthening contributions are considered
additive [103], but the grain size distribution and uneven particle spacing may well
complicate the extent to which separate contributions can be determined.
In the absence of chemical composition data, it is dicult to readily compare the
dierence in HallPetch slopes between the cryomilled Al and the nanostructured
ball-milled Al [100]. While the value of ky derived from these data is reliable in
the sense that it is obtained from a single study, the low strength of the ball milled
Al relative to its grain size make comparison to ky of the cryomilled Al inconclusive.
Highly pure Al (99.99%) subjected to ECAP processing gave an average grain size of
1.3 lm in the as-processed
p condition which could be increased by annealing, giving a
ky value of 0.12 MPa m [104], slightly larger than that of the cryomilled Al. Recrysp
tallized and extruded SAP yielded ky values ranging from 0.03 to 0.07 MPa m,
although these were considered not statistically distinguishable [101,103,105]. These
p
values were comparable to a literature value for 99.9% pure Al of 0.059 MPa m
[103], with a chemical composition similar to the cryomilled Al. Furthermore, the

40

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

sub-grain size was correlated with dispersoid content, leading to linear relationships
between ow stress with both microstructure and dispersoid content [103]. As shown
in Fig. 14, the linear trends for yield stress of both MA Al [10] and SAP materials
with respect to dispersoid content converged to a similar intercept of between 100
and 110 MPa for 0% dispersoid content. The value of r0 for cryomilled Al (Fig.
14) was 130 MPa, 25 MPa greater than this gure. The correlation between yield
stress and dispersoid content is not very conclusive, and appears to be less important
than the role of coarse grained regions in the extrusions. There are insucient data
to draw rm conclusions about the similarity of grain size dependency for cryomilled
Al and the dispersoid content for MA Al and SAP, but the similarity is unlikely to be
a complete coincidence. For both the MA and cryomilled Al, the heterogeneous distribution of dispersoid particles may be the determining factor for the heterogeneity
in the microstructure, leading to a coincidence in trends for average grain size and
dispersoid volume.
For the Al 5083, only the longitudinal grain size measurements are included for
those extrusions that exhibited
grain anisotropy. As plotted on Fig. 12, these data
p
gave a ky of 0.28 MPa m. The longitudinal dimension of each grain that was measured in TEM micrographs was converted to an approximate grain volume based on
a rectangular prism by factoring in the average grain aspect ratio. The average volume of the sample was converted back to a linear dimension according to the same
aspect ratio. Because larger grains contribute more proportionally to the overall volume than their frequency in the grain size distributions, the volume averaged grain
size is necessarily larger than the direct measured grain
p size. These values are not
plotted in Fig. 12, but they gave a slope of 0.27 MPa m, consistent with the value
for longitudinal grain size. Values of ky in the ne-grained regime for Al are not
prevalent in the literature, prompting earlier investigators to use published values
for the slope in attempting to explain the strengthening mechanisms operative in
cryomilled AlMg alloys. For p
Al7.5Mg with a grain size of approximately
300 nm, a ky value of 0.068 MPa m was used to estimate the grain-size component
of strengthening at 124 MPa [75]. Forpcryomilled Al 5083 with a yield strength of
334 MPa [66], a value of 0.063 MPa m was cited [106]. Interestingly, this latter
value was based on ky values for several dierent Al alloys, including pure Al, Al
6Mg0.6Mn and Al2Cu2Mg1Ni1Fe (with only two points for the last of these)
[106]. The actual data generated in that experiment,
for mechanically alloyed
p
Al4Mg1.5Li0.4O (IN905XL), was 0.32 MPa m over a grain size range of
0.84.3 lm, while for a grain size range of 4.38.1 lm the value of ky was slightly less.
This latter point, i.e., the extent to which ky for ne grain sizes corresponds to coarser grain sizes, deserves some additional comment. The decrease in ky with increasing
grain size noted for Al4Mg1.5Li0.4O was apparently consistent with behavior observed in Al6Ni, leading Mukai et al. [106] to speculate that a change in slope was to
be expected for grain sizes either above or below approximately 10 lm. In the original
paper for Al6Ni [107], however, the break in slope at a grain size of approximately
4 lm could be eliminated by considering the eect of inhomogeneous yielding caused
by formation of Luders bands. By extrapolating the stressstrain curve from the
work-hardening portion back to the elastic portion, slightly lower yield and ow

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

41

stresses were calculated,


leading to constant ky values over all grain sizes, with a value
p
of 0.14 MPa m based on yield stress. Stressstrain curves of Al4Mg1.5Li0.4O
show evidence of inhomogeneous yielding [106], so it is possible that a constant ky
over all grain sizes may have resulted if the data were considered dierently. While
the authors discussed this fact, they calculated the grain-size renement contribution
to strengthening from the lower value based on other literature.
There are other published values of ky for AlMg alloys that are in better agreement with those for cryomilled Al 5083. Data presented in an investigation of inhop
mogeneous yielding in AlMg alloys [108] gave ky values of 0.12 and 0.25 MPa m
for Al3.5Mg with dierent as-received yield-point characteristics and grain sizes
ranging from 10 to 40 lm. More recently, ky values have been reported
for a number
p
of dierentp AlMg alloys [109]. These ranged from 0.15 MPa m for Al5Mg to
0.27 MPa m for Al 5754 (Al3Mg),
for grain sizes from 10 to 55 lm. ky values ranp
ged from 0.16 to 0.31 MPa m for a number of dierent MA alloys containing Al
and Al2O3 with various additions of yttria, Cu, Nb and Ti [110]. Finally,p ultrane-grained AlMg alloys
p processed by ECAP had ky values of 0.13 MPa m for
Al1Mg and 0.15 MPa m for Al3Mg [104]. The value of the HallPetch
slope
p
for these AlMg alloys tend to fall between 0.15 and 0.30 MPa m, and would
appear to be a function of both material preparation and alloy composition. A
summary of HallPetch slopes is given in Table 7.
Table 7
HallPetch slopes
p

Material

HallPetch slope, ky (MPa

Al and AlAl2O3 composites


Cryomilled Al
Ball-milled Al
ECAP Al
SAP Al

0.09
0.012
0.12
0.030.07

[78]
[100]
[104]
[101,103,105]

AlMg alloys
Cryomilled Al 5083
MA IN905XL (Al4Mg1.5Li0.4O)
Al3.5Mg
Al 5052 (Al2.6Mg)
Al 5754 (Al3Mg)
Al5Mg
Al5Mg0.2Mn
ECAP Al1Mg
ECAP Al3Mg
Al6Ni

0.28
0.32
0.120.25
0.21
0.27
0.15
0.22
0.13
0.15
0.14

This work
[106]
[108]
[109]
[109]
[109]
[109]
[104]
[104]
[107]

Mechanically alloyed Al composites


Al plus C, O (14.1 vol.%)
Al plus C, O, Cu (12.8 vol.%)
Al plus C, O, Y (5.9 vol.%)
Al plus C, O, Nb (16.1 vol.%)
Al plus C, O, Y, Nb (15.9 vol.%)
Al plus C, O, Ti (31.6 vol.%)

0.23
0.31
0.200.27
0.16
0.22
0.25

[110]

m)

References

42

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

The HallPetch slope for Al 5083 plotted in Fig. 12 would intersect the yield
strength of strain-hardened and O temper Al 5083 at grain sizes of approximately
1 and 3 lm, respectively. These grain sizes are somewhat smaller than other
coarse-grained AlMg alloys discussed here. Adding experimental data points for
coarse-grained Al 5083 to the data for cryomilled Al 5083 might thus tend to produce the same trend as observed in Al4Mg1.5Li0.4O [106] and in Al6Ni [107].
As will be discussed, there is evidence that like these alloys, the cryomilled Al
5083 exhibits inhomogeneous yielding. On this basis, the values of ky may slightly
overestimate a true value that would hold at both UFG and coarse grain sizes, or
it may be attributable to the presence of dispersoids. Nevertheless, even if this value
for ky werepcloser to the values cited by Lloyd for Al6Ni [107] or Al5Mg [109], i.e.,
0.15 MPa m, it would signicantly exceed the values chosen by previous investigators to estimate the grain size strengthening component of their respective materials
[66,75,106]. Increasing ky gives a directly proportional increase in the strength derived from the grain size renement.
If the contribution to strengthening has been underestimated for grain size renement, the other sources of strengthening have necessarily been overestimated. These
other components include solute strengthening due to Mg in particular, dispersoid
strengthening based on the Orowan mechanism, and the presence of dislocations
in the as-extruded material. For the rst two, the additional strength is predicated
on the interaction of gliding dislocations with a solute atoms or a network of obstacles. If the dispersoids are more likely to be found at grain boundaries than within
the grains, the contribution due to the Orowan relationship would be less than that
based on calculation of the total volume fraction of dispersoids. The Orowan relationship takes into account both particle dimensions and spacing and can be given as
r/

1
d
ln ;
L
b

where r is the increase in ow stress due to the presence of particles, L is the interparticle spacing, d is the particle diameter and b is the magnitude of Burgers vector.
As discussed above, the particle spacing is likely to be larger than that predicted by
the estimated volume fraction, and the eective volume fraction for the Orowan process, that is those particles within grain interiors, is likely to be lower still [111].
Therefore, while Orowan strengthening cannot be ignored, it is not likely to occur
to the extent predicted by formulas of the type exemplied by Eq. (7). Studies of
MA Al alloys have reached the same conclusion [86,110].
In examining the strengthening role of Mg, it must rst be emphasized that microscopic evidence, both imaging and selected area diraction, indicates that AlMg
intermetallic compounds are either absent, restricted in occurrence to coarse-grained
regions, or are in general irregularly distributed. An appreciation for the strengthening due to Mg can be gained from reviewing data for MA Al alloys. For an extruded
MA AlMg alloy in which the Mg content was varied from 0 to 6.84 wt.%, the oset
yield strength and tensile strength increased by approximately 75% and 65%, respectively [112]. The addition of Mg increased the oxygen content from 1.39 wt.% to
roughly 1.6 wt.%, independent of the amount of Mg added, while the carbon content

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

43

was roughly 0.5 wt.%, and independent of Mg content. Because the investigators
indicated that the microstructures did not vary greatly by alloy, the increased
strength in these alloys could be presumed to result primarily from the increasing
Mg additions. The data in reference [112] are best tted by a quadratic equation
of the form
r Ac2 Bc rAl ;

where c is the Mg content, A and B are empirical constants, and rAl is the yield or
tensile strength of the MA Al without the Mg addition. In contrast to this relationship, a model for ow stress due to solute strengthening derived from principles of
dislocation-solute interactions has a dependence on c1/2 [113]. This model would predict an increase in strength of only 50 MPa for the alloy with 6.84 wt.%Mg, much
less than the actual dierence between Al and Al6.84Mg. Other models for solute
strengthening which demonstrate a linear relationship between solute concentration
and ow stress are valid only at dilute solute concentrations and tend to overestimate
the ow stress [88]. In addition, use of any of these models, like the Orowan model
for dispersoid strengthening, is based on the interaction of gliding dislocations with
solute atoms producing an elastic strain eld in the lattice. The location of the Mg in
these alloys is thus an important consideration in determining the extent to which
Mg strengthens the material.
The exact disposition of Mg cannot be addressed directly for the MA AlMg
mentioned above, except to say that annealing treatments at temperatures up to
316 C at times up to 100 h had essentially no eect on the yield or tensile strength
[112]. In the case of extruded cryomilled Al7.5Mg, positions of the (1 1 1) fcc XRD
peak for as-extruded and annealed samples gave some further indication of the
mobility of Mg [75]. Based on these positions, and the linear relationship between
lattice constant and Mg solute content [114], approximate Mg contents in solid solution for XRD data in Han et al. [75] may be calculated as 4.7 wt.% for the asextruded, 3.4 wt.% for extruded material annealed 2.5 h at 450 C and then water
quenched, and 1.5 wt.% for extruded material annealed 2.5 h at 450 C and furnace
cooled. TEM of the as-extruded material indicated that no AlMg precipitates were
present. The change in solid solution Mg after water quenching suggests more Mg
residing at grain boundaries, while furnace cooling might bring about precipitation
of a second phase. Of particular interest in this case is a comparison of ow stress at
room temperature of the as-extruded with that of the extruded after annealing and
furnace cooling: they are virtually identical [75], despite a signicant dierence in the
presumed Mg content of the grain interiors. This nding suggests that the strengthening derived from Mg is related to its presence at grain boundaries, not grain
interiors.
To better evaluate the role of annealing or heat treatment on mechanical properties, extruded cryomilled Al 5083 from commercial-scale extrusion H-4b was subjected to a number of annealing treatments and tested using microhardness
indentation on its transverse section. These hardness results are shown in Fig. 15,
and the results show little variability in hardness for temperature exposures below
500 C. The temperature 450 C is notable because it is the temperature of maximum

44

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Fig. 15. Microhardness values of extrusion H-4b after various heat treatments.

Mg solubility in solid solution in Al. The short heat treatments at 450 and 500 C
were chosen in part based on the observation by Lloyd et al. [115] that the nature
of Luders strain in Al 5182 could be changed by short anneals at each temperature.
In that case, the amount of Luders strain was reduced only after water quenching
samples subjected to 500 C for 2 min. Samples air cooled from 500 C, as well as
those air cooled and water quenched from 450 C, all had similar amounts of Luders
strain.
3.3. Room temperature mechanical properties: nature of deformation
3.3.1. Inhomogeneous ow
Tensile stressstrain curves of several cryomilled Al 5083 extrusions are shown in
Fig. 16, plotted to show the deformation prior to the onset of necking. The stress
strain curve for material labeled Al4Mg is for material whose creep behavior was
previously described [95]. The Al 5083 curves labeled 1, 2 and 3 are from the same
batch of cryomilled powder. Curve 1 is from material that was extruded with a
9:1 ratio at 300 C, while curves 2 and 3 were taken from the same HIP billet, and
extruded at a 9:1 ratio at temperatures of 325 and 350 C, respectively. The grain size
distribution and microstructure of the material in curve 2 has been given previously
in Fig. 4 as H-3. These curves dramatically show the eect of extrusion temperature

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

45

700
Al-4Mg
Al 5083 (1)

600

Al 5083 (2; H-3)


Al 5083 (3)
Al 5083 (H-4b)

True Stress (MPa)

500

400

300

200

100

0
0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.1

True Strain

Fig. 16. Stressstrain curves of cryomilled Al 5083 in tension at a strain rate of 103 s1. Curve for
Al4Mg is unpublished data from Tellkamp et al., [98]; creep data for this material was published
elsewhere [95]. Curves labeled Al 5083 1, 2 and 3 are from the same laboratory powder batch; grain size for
curve 2 is displayed in Fig. 5 as sample H3. The material for the curve labeled H-4b is from a commercial
scale cryomilling run and HIP billet, followed by extrusion. Its grain size is also displayed in Fig. 5.

on the resulting stressstrain behavior. A similar correlation has been noted for cryomilled Al10Ti2Cu extruded at 205 and 300 C [76]. In comparison, variation in
HIP temperature of cryomilled Al had a much more limited eect on strength
[78]. In comparison is a stressstrain curve for material (H-4b) from a commercial
scale cryomilling run, single HIP billet, and isothermally extruded at 225 C, for
which the grain size distribution is given in Fig. 5. A summary of the room temperature tensile properties of these and other cryomilled AlMg is presented in Table 8.
The strength values are generally higher than those for AlMg alloys processed by
ECAP, for which room temperature yield strengths of 410 MPa for Al4.8Mg
[116], 355 MPa for Al6Mg0.6Mn [117], and roughly 230 and 375 MPa for Al
1Mg and Al3Mg respectively [104] have been obtained. The yield strength of ECAP
5083 was reported as 276 MPa [118]. ECAP AlMg alloys typically have average
grain sizes around 200500 nm, i.e., slightly larger than the cryomilled material,
which may well explain much of the dierence, and it should be noted that the additional strength in cryomilled materials come at the price of lower room temperature
ductility.
The stressstrain curves of Al 5083 typically exhibit serrated ow or yield drop at
low strain, followed by ow softening or constant ow stress, and nally localization
of ow leading to fracture. Flow softening has been noted in ne-grained Al alloys

46

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Table 8
Room temperature mechanical properties of cryomilled extruded Al and Al 5083
Extrusion

Yield stress
(0.2% oset) (MPa)

Tensile strength
(MPa)

% Elongation
(from gauge)

References

Al 1
Al 2a
Al 2b
Al 2c
Al 2
Al 2
Al 5083 1
Al 5083 2
Al 5083 3 (H-3)
H-1
H-2
H-4a
H-4b

440
380
329
346
347
323
523
453
169
660
705
544
555

479
404
346
366
354
345
584
491
199
682
745
611
636

4.3
(17.1)
(31.3)
(27.1)
(24.3)
(41.6)
3.7
5.9
22.8
(1.4)
(1.1)
(8.6)
(6.4)

[98]
[78]
[78]
[78]
[78]
[78]
[74]
[74]
[74]
[74]
[74]
[74,78]
[74,78]

processed by mechanical alloying and extrusion [110,119], plasma spraying followed


by cold pressing [102], and cryomilled AlAl2O3 [89] and Al10Ti2Cu [92], and can
be seen for cryomilled Al in Fig. 13. The yield drop in Al4Mg and the serrated ow
especially evident in sample H-4b in Fig. 16 is notably absent from these other alloys,
but it is present in other ne grained AlMg alloys, such as MA Al4Mg1.5Li
1.2C0.4O [106,120], MA Al4Mg1.5O0.8C [121], and cryomilled Al7.5Mg
[75]. In each of these materials, as in the present cryomilled Al 5083, there is a yield
drop, followed by successive serrations which diminish in relative amplitude with
strain, and vanish after the third or fourth stress drop. After the cessation of serrated
ow, various behaviors have been observed, including the onset of work hardening
in MA Al4Mg1.5Li1.2C0.4O [106], ow softening in MA Al4Mg1.5O0.8C
[121], and nearly perfectly plastic ow in Al7.5Mg. The observation that a yield
drop and serrated ow are observed only in MA and cryomilled alloys containing
Mg is important in discerning its cause. For MA AlMg, the concentration of Mg
has been shown to inuence the nature of the serrated ow, as it was not found in
extruded material containing less than 1% Mg [122]. An important role for Mg
has also been noted to cause serrated ow in conventional materials [123]. If the yield
drop were due solely to unpinning of dislocations from dispersoid particles, as has
been suggested [121], then it would be expected in other MA or cryomilled materials
with a similar microstructure that also contain dispersoids. It is also important
to point out that although inhomogeneous yielding and serrated ow are not uncommon in AlMg, their behavior as described for conventional alloys [108,115,124,125]
is quite dierent from those seen in the dispersoid strengthened cryomilled and MA
AlMg alloys.
There is extensive literature on the subject of yield-point phenomena and dynamic
strain aging, but this discussion has not been extended to materials with ultranegrained or nanostructured grain sizes. The appearance of either a yield point or serrated ow is a function of test temperature and strain rate, and is sensitive to testing

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

47

apparatus. In the original treatment of dynamic strain aging proposed by Cottrell


[126], solute or interstitial atoms diuse to mobile dislocations, thus requiring additional stress to assist dislocation movement. The subsequent unlocking of dislocations from the solute atmospheres leads to a drop in stress necessary to sustain
ow. While modications have been proposed, such as transfer of solute atoms from
forest dislocations to mobile dislocations [127], the essential role of diusion of solute or interstitial atoms can be presumed. As the diusion of the species is assisted by
formation of vacancies, some amount of strain is typically necessary to initiate serrations. The Cottrell treatment, however, is not universal, as it does not hold at temperatures greater than some critical value, which in Al alloys is 0.3Tm [128]. The
critical temperature for Al5Mg was 270 K at a low strain rate (6.4 105 s1)
and 323 K at higher strain rate (102 s1) [128], which brackets the room temperature testing condition. Above the critical temperature, the immediate onset of serrated ow has been interpreted as evidence of a sucient defect concentration
prior to testing, a condition that may be created through quenching, pre-straining
or radiation damage [128,129].
Another generalization about dynamic strain aging is that the magnitude of the
stress drops increases with increasing strain [125], which may also be attributed to
increasing vacancy concentration resulting from deformation. For the MA and cryomilled AlMg materials, the stressstrain curves showed evidence of a yield drop
early in the tensile tests, as shown in Fig. 16 and the literature [120,122], but the
serrations decreased in magnitude after increasing strain. Furthermore, the amount
of strain between stress drops in the cryomilled Al 5083 decreased successively.
The yield drop observed in the cryomilled and MA AlMg alloys did not occur at
the onset of plastic ow, but after a short work hardening period. For the MA
AlMg, this was explained in terms of the highly variable dislocation density in
grains [122]: in grains with existing dislocations, the dislocations are presumed to
be locked by Mg atmospheres, while in dislocation-free grains, the imposed stress
of the tensile test leads to an increasing dislocation density and an increase in the
Mg locking due to strain.
Serrated ow may also occur at low strain during Luders extension [124], but for
the MA and cryomilled materials, the stress drops are of sucient magnitude that
any plateau in stress typical of Luders band propagation is obscured. Mukai et al.
[106] reported Luders band propagation for MA AlMgLi, although it is not clear
if this is based on observation of the bands or inferred from the stressstrain curves.
Also, the Al3Li phase was not discussed, the presence of which may inuence the
mechanical behavior, as has been suggested for Al3Ti in cryomilled AlTiCu [92].
A more thorough depiction of the low-strain behavior was presented for cryomilled
Al7.5Mg [75], which was tested at three dierent strain rates, as well as at the same
strain rate under as-extruded and two dierent annealed conditions. In these cases,
the magnitude of the stress drops (1550 MPa) is considerably higher than that
typically observed in dynamic strain aging [75]. Tensile tests of the as-extruded cryomilled Al7.5Mg at strain rates from 4 104 to 4 102 s1 showed that the magnitude of the yield drops was inversely related to strain rate. The data also indicated
an inverse relationship between the onset of the rst yield drop and strain rate.

48

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Some of the observations above are consistent with the serrations being of the
unlocking type described for an AlMgSi [130,131], which occurred at the beginning of serrated ow and were associated with the negative strain-rate sensitivity
of the strain at initiation. These serrations result from breakaway of dislocations
from pre-existing solute atmospheres (hence, unlocking). As described previously,
pre-existing atmospheres in the as-extruded material are also consistent with the
early onset of serrated ow. On the other hand, the form of the serrations in several
of the stressstrain curves of cryomilled Al 5083 may better t the description of
locking serrations, which appear as rapid increases in ow stress followed by a drop
in stress back to the average level of the curve [130]. The serrations in Al7.5Mg [75]
and in the cryomilled Al 5083, particularly the commercial-scale batch (H-4b in Fig.
16) appear to conform to the description of locking serrations. A further complication is presented by the dependence of yield drop and serrated ow on the orientation of the extruded bar in a MA AlMgLi alloy [120]. Yield drop was only noted in
samples for which the tensile load was applied parallel to the extrusion direction,
whereas tensile tests on forgings or samples oriented normal to the extrusion direction did not exhibit yield drop or serrated ow. If the yield drop and early serrated
ow may be ascribed to a dislocation population that is aged from the onset of the
test [128], then serrations might be expected in MA AlMgLi after sucient strain,
but this was not observed [120]. The discrepancy between forged and extruded material can be rationalized in terms of the dislocation density arising from the respective
forming processes. The dierence in ow behavior arising from sample orientation
has been attributed to the grain aspect ratio, and a critical dislocation density on
the slip plane that would allow some initially pinned dislocations to become mobile
[120]. This interpretation suggests that normal to the extrusion direction, not only
was the dislocation density in the as-extruded condition insucient to cause yield
drop, but it did not reach the critical level during deformation.
The nature of the ow serrations in Al7.5Mg has not been claried by investigation of annealing treatments. The as-extruded material and material subjected to a
2.5 h anneal followed by a furnace cool had very similar ow stress levels and
serration characteristics [75]. For breakaway of a dislocation of length L from a
Cottrell-type atmosphere to occur at temperature T, the applied force F must exceed
the maximum value required to maintain a linear displacement from the atmosphere,
and can be given by [132]
F c0 b2

;
L bkT

where c0 is the lattice concentration of solute forming the atmosphere, b is an interaction energy between solute atoms and dislocation, b is the magnitude of Burgers
vector, and k is Boltzmans constant. The direct relationship between the maximum
stress necessary to cause the breakaway and c0 implies that two materials exhibiting a
similar UTS and serrated ow characteristics contain similar solute atmospheres.
This is contradicted, however, by the signicant change in lattice constant, which
indicates that there was about three times as much Mg in Al solid solution after
extrusion compared to the annealed and furnace cooled condition [75]. The annealed

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

49

and quenched material had a ow stress roughly 25 MPa higher, and exhibited a
lower stress drop on each serration. This result is not consistent with the inferred lattice concentration of Mg.
Finally, the reduction in amplitude of the serrations with increasing strain is not
typical, although it was described by Mulford and Kocks for commercial Inconel 600
at elevated temperatures [127]. If the serrations are of the unlocking type, suggested
by the low strain at which they initiate, as well as the negative strain rate sensitivity
to onset, then their reduction in magnitude would be due to the disappearance of Mg
solute atmospheres. If the serrations are of the locking type, then this eect may tentatively be ascribed to a decrease in the production of vacancies during deformation.
Unlocking serrations have been associated with the propagation of Luders bands
[124,131]. The Luders elongation and serrated ow in AlMg have been correlated
with dierent categories of surface markings [108,124]. The prevalence of these
markings on gauge sections, as well as their manifestation in stressstrain curves,
is a function of numerous factors, including grain size, heat treatment, testing temperature and strain rate [115,125]. For cryomilled Al 5083, surface markings of two
kinds were detected on test specimens made from the commercial-scale cryomilled
extrusions, but not on the laboratory-scale material, which failed abruptly at low
strain. The surface markings were of two distinct kinds. The rst appeared as parallel, shallow grooves lying roughly 6070 to the tensile axis found throughout the
gauge section. These grooves occurred with a spacing of 50150 lm and were readily
distinguished from the grinding marks left by sample preparation, which were perpendicular to the tensile direction, as shown in Fig. 17a. The second type of marking
was found only prominently in the necks of the tensile specimens. These were visible
to the naked eye as thin bands perpendicular to the tensile axis. At higher magnications, they appear as smooth bands approximately 150 lm wide as shown in Fig.
17b, from which it is obvious that these bands formed after the shallow grooves.
Aluminum alloys which undergo inhomogeneous yielding typically exhibit two
types of surface markings [125]. These are related to the propagation of a primary
Luders band across the sample, followed by the onset of serrated ow throughout
the specimen. In AlMg alloys, the rst of these, or Type A markings, are often random in appearance [125], while in mild steel they appear as parallel bands aligned
obliquely to the tensile direction. The bands associated with serrated ow, or Type
B, appear in AlMg as parallel bands [125]. In cryomilled Al 5083, both sets of markings are parallel to each other. From SEM images, the shallow grooves are observed
to be 10 lm wide. Given their spacing, the length of the gauge section (19.05 mm),
and the amount of uniform plastic strain in the sample (slightly more than 4%, or
0.8 mm), the total uniform strain in the gauge may be accounted for solely by the
extension in these parallel grooves. In this sense, they were similar to traditional conceptions of Luders bands. After the onset of necking, the plastic ow within the stilldeforming neck region was more uniform, and the surface of the neck is smoothed
over by homogeneous ow. In this interpretation, the absence of Luders bands reported in other cryomilled AlMg [75,94] may be attributed to the greater uniform
elongation, i.e., the bands were not visible because they have been removed by subsequent deformation. The serrations observed in the various ne grained MA and

50

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

Fig. 17. (a) SEM image of parallel surface markings in tensile specimen from extrusion H-4a; (b) Smooth
parallel bands within the neck of tensile specimen from same extrusion.

cryomilled AlMg alloys may thus be related to the propagation of Luders bands.
The gradually decreasing stress with each successive serration is the result of an eectively decreasing gauge section radius. Less stress is required to continuing deforming the samples because the radius has been reduced by the grooves that appear.
After the nal serration, the material continues to ow soften, a characteristic common to cryomilled and MA Al alloys [86,92,119121]. The fracture surfaces of cryomilled Al 5083 exhibited evidence of ne ductile dimples, even in materials with very
low ductility. This has been attributed to a deformation mechanism in which

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

51

dislocations rapidly accumulate at grain boundaries and triple junctions, leading to


void nucleation and an eective reduction of the gauge cross-section. In the case of a
cryomilled Al, microvoids were surmised to nucleate at dispersoid clusters and coalesce, leading to strain localization [89]. The role of dynamic strain aging in ow
localization at temperatures where dislocations are aged from the start of the tensile
test has also been described [129]. In this case, the planes of maximum shear stress
that give rise to initial dislocation motion become softer than those where strain
aging is occurring, becoming bands of preferential ow. The imposed load is accommodated through ow in these bands and stress on other glide planes is never high
enough to activate ow.
In addition to strain localization, another mechanism for reducing the ow stress
would be to reduce the dislocation density. For cryomilled materials, this could happen in two ways. The high initial dislocation density is supported by the nature of the
tensile stressstrain curves as well as TEM images which show high numbers of dislocations in relatively coarse grains of the as-processed materials [75,133]. During
deformation, these pre-existing dislocation densities may not be sustainable, and dislocations glide rapidly to grain boundaries. A second way to reduce ow stress is by
a change in grain boundary structure during deformation, specically by a reduction
in the stress necessary to propagate glide from one grain to the next, leading to a decrease in HallPetch slope with increasing strain [110]. In conjecturing this hypothesis for an MA alloy, it should be noted that the notion of decreasing dislocation
density was rejected [110], despite the theorys being developed for a stressstrain
curve that exhibited more ow softening than cryomilled Al 5083. However, a relationship between decreasing ky and decreasing dislocation density with strain was
noted in conventional AlMg based alloys [109]. A reduction in dislocation density
would also be expected to reduce the maximum stress, as this stress has been related
to interactions between dislocations and Mg atmospheres, as already discussed, as
well as ne dispersoids that pin the dislocations, as has been suggested for MA
AlMg alloys [121].
In summary, most of the interpretation and discussion of inhomogeneous ow in
the literature are based on materials with much coarser grain sizes, and indeed more
homogeneous microstructures than nanostructured or ultrane-grained materials
made via cryomilling. In conventional ingot-processed materials, the appearance
of Luders strain and ow serrations are sensitively related to various material properties and testing conditions, so a complete understanding of the early deformation
of the ne grained cryomilled materials requires additional study based on a wider
matrix of pre-treatments, testing temperatures and strain rates. For the ow softening evidenced by the H-4a and H-4b extrusions, there are several plausible explanations. The observation of parallel grooves in the gauge surface and the dimpling of
the fracture surface both indicate that an eective reduction in the gauge cross-section reduces the load that the tensile specimen can bear. The work hardening before
onset of serrated ow indicates an accumulation of dislocations within grain interiors, before the breakaway of dislocation segments pinned by solute atmospheres.
One explanation for reduction in amplitude of serrations with increasing strain is
that there are simply fewer dislocations to interact with Mg solute atmospheres,

52

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

indicating a decrease in the dislocation density. This would also be consistent with
the transfer of Mg to grain boundaries. As dislocations with solute atmospheres
are absorbed at grain boundaries, the atmospheres are absorbed as well. Such a scenario would also explain, in part, the absence of any critical strain for serrated ow
in MA AlMgLi tested normal to the axis of extrusion [120]. In this case, a critical
defect density is not achieved because the shorter glide distance between grain
boundaries does not permit it. At a certain strain, the increase in dislocation density
at grain boundaries leads to the formation of microvoids, and the fracture process
begins. In cryomilled Al 5083, this was accompanied by ow localization, while in
Al7.5Mg ow softening and localization were not observed [75].
3.4. Room temperature mechanical properties: compression testing of Al 5083
Compression testing allows for more extensive analysis of the materials deformation properties. For example, cryomilled Al 5083 extrusions H-1 and H-2, both of
which failed abruptly after only about 1% elongation in tension (Table 8), were compressed at room temperature and a strain rate of 103 s1 to strains of greater than
20% before barreling and 10% before fracture, respectively. In each of these extrusions, the maximum ow stress was followed by a gradual ow softening that resembled the compression behavior of cryomilled Al10Ti2Cu at room temperature [91].
For extrusions H-4a and H-4b, uniform strains in compression were greater than
25%, as shown in Fig. 18. Due to the limited ductility in tension, the potential tensile
strengths of laboratory-scale extrusions H-1 and H-2 were not achieved, so an under-

900
H-2
800

H-1

True Stress (MPa)

700

H-4a

600

H-4a

500
400
300
200
100
0
0

0.05

0.1
True Strain

0.15

0.2

Fig. 18. Compressive stressstrain curves for various Al 5083 extrusions tested at a strain rate of 103 s1
at room temperature.

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

53

standing of the deformation was limited to the brief period of work hardening after
the elastic limit. The yield stress in compression was approximately 5% greater in
compression than in tension, which is similar to a cryomilled Al10Ti2Cu alloy that
exhibited an 810% dierence in ultimate strength between tension and compression
[93,134]. Both this AlTiCu alloy and the Al 5083 extrusions share an important
microstructural characteristic, large elongated coarse-grained regions arising during
the HIPping process and remnant prior particle boundaries [73,92]. TEM images of
the ne graincoarse grain interfaces indicate that they are free of microporosity,
with growth of the coarse-grained region occurring at the expense of the ne-grained
region once the inter-particle void spaces have been lled. These interfaces, which are
really remaining prior particle boundaries, are thought to be a source of residual
stress and thus contribute to failure through stress concentration and delamination
[93]. Tensioncompression asymmetry has been previously noted for other bulk
UFG/NC materials [135], but the multi-phase structure is likely responsible for
the asymmetry observed in cryomilled materials, as the two distinct regions have differing levels of strength. The strength of these extrusions is quite high, but the asymmetry in this case can be attributed more to the microstructural variation than to an
intrinsic feature of the grain size. For cryomilled Al10Ti2Cu, which also exhibited
this microstructure, homogenization of the material by friction stir processing reduced the strength in tension by approximately 12% compared to the as-extruded,
although it still remained relatively high (650 MPa), and the ductility was improved
by more than ve times [76].
In comparison to the two laboratory-scale extrusions of Al 5083, the yield stress
of commercial-scale extrusions H-4a and H-4b were less than 2% greater in compression than in tension. Cryomilled Al7.5Mg exhibited a similar behavior, with nearly
overlapping tensile and compressive stressstrain curves up to the point of the stress
drop in tensile testing. While there were artifacts of the processing evident in the
commercial-scale Al 5083 extrusions, again in the form of coarse grained bands, they
were no longer well-dened prior particle boundaries. Given the similarity of tensile
and compressive behavior, a general model for the deformation mechanism consistent with both tests is more likely to be applicable to materials in the same grain size
regime than one developed based on a two-phase microstructure.
Compression tests at room temperature of both H-4a and H-4b exhibit a brief region of work hardening after yield, followed by uniform deformation with work
hardening as shown in Fig. 18, although the ow is nearly perfectly plastic. In each
curve, there is a point after several percent plastic strain where the work hardening
rate dr/de approaches zero, and thereafter assumes a nearly constant positive value.
This point occurs at a stress value that is close to the ultimate stress in tension of the
material. In tension, the UTS appears to represent the onset of Luders band propagation associated with the breakaway of dislocations from Mg solute atmospheres or
initial unpinning of dislocations from dispersoid particles within grain interiors. Because the zero value of work hardening rate in compression occurs at a similar stress
level, a similar unpinning phenomenon and mobilization of dislocations is surmised.
After this point of deformation, the nearly constant stress level suggests a close
balance between dislocation generation and annihilation. Considering the complex

54

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

nature and scale of the microstructure, this observation has several important implications. First, the dislocation sources may be FrankRead sources within the grain
interiors, especially the larger grains. For larger grains containing dislocations pinned by nanometer-scale dispersoids, these sources may operate continuously during
plastic deformation. Even the coarsest grains are less than 1 lm, so the gliding distance from grain interior to grain boundary is still limited, and would allow continuous production of dislocations that do not interact with other gliding dislocations
in grain interiors. A second source is the grain boundaries themselves, which has
been suggested for MA alloys [119] and micrometer-scale grained Al [115,136]. In
this case, the balance between dislocation generation and annihilation is maintained
again by the short glide distances of ne-grained materials. Grain boundary dislocations are emitted under stress, and then glide unobstructed until they are absorbed at
the opposite side of the grain, leading to net plastic strain with no gain in dislocation
density, and hence no change in ow stress [63,110,134,137].
In tension testing of the same cryomilled Al 5083, the strain occurring after the
UTS was reached was marked by the appearance of surface markings on the gauge,
but the surface of the compression samples did not show evidence of inhomogeneous
ow, such as the shear bands that have been found in compression of NC and UFG
Fe [79,135]. It should be noted that in both Fe studies, polished prismatic samples
were employed to allow observation of shear bands [79]. For cryomilled Al 5083,
cross-sections of compression samples deformed to approximately 25% strain and
examined after polishing and etching did not show macroscopic evidence of shear
banding either. Given the presence of the elongated coarse-grained regions, propagation of shear bands through the sample would result in a change in the appearance of
these regions, such as displacement of one end of a band relative to the other. This,
however, was not observed in analysis of materials after compression testing. If the
deformation in compression initiates within the ne-grained matrix, as suggested by
the high strength levels, the coarse-grained regions may serve as a barrier to shear
band propagation. At the scale of optical or scanning electron microscopy, changes
within the ne-grained matrix are not accessible, and TEM study of deformed specimens will be necessary to determine the extent to which the ne-grained matrix has
been deformed. Because the coarse grain regions have a larger grain size, and are presumed to have fewer dispersoids, they may deform favorably relative to the negrained regions.
At the same time, TEM of the coarse-grained regions in both the low-ductility
laboratory-scale specimens and the more ductile commercial-scale extrusions showed
that they are composed mainly of aligned sub-grains with sub-micron dimensions.
Room temperature deformation, even if concentrated in the coarser regions, is thus
occurring by dislocation movement in sub-grains of similar length scales as those inferred from the grain size distributions. In addition, the HallPetch relationship
exhibited by the cryomilled Al 5083 can arise due to dislocations emanating from
FrankRead sources and forming pile ups within grain interiors (the original concept
of the model), or at subgrain or grain boundaries [138]. In this model, the high levels
of ow stress sustained during room temperature compression testing are consistent
with strain within either component of the microstructure.

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

55

Considerable deformation has been observed in NC and UFG materials with a


wide distribution of grain sizes, in which the ne grained matrix appeared undeformed after strain of greater than 100% [68]. This was attributed to a combination
of grain-boundary sliding in ne grains and dislocation-based plasticity in the larger
grains. In the tensile deformation of an MA AlMg alloy, which showed more ow
softening than the cryomilled material, the deformed microstructure appeared similar to the untested material [121]. In an AlMg material cryomilled with the addition
of 20.6 vol.% AlN particles, deformation by compression at 550 C to a strain of
100% did not result in signicant microstructural change [53]. This material was
characterized by coarser Al regions without AlN particles, so it too had a duplex
microstructure. TEM of deformed specimens will be necessary to better determine
the nature and location of the plastic deformation. In particular, the dislocation distribution with respect to grain size and microstructure is necessary to better implicate
the various dislocation sources, and hence the mechanism of plasticity.

4. Conclusions
The properties of cryomilled materials as described indicate both their promise as
well as some of the areas where additional work is necessary. In the powder form, the
cryomilled materials exhibit grain sizes that are comparable to the minimum grain
sizes achieved in mechanical milling. The thermal stability of cryomilled Fe is similar
to that of mechanically milled Fe, and can be increased dramatically by the addition
of alloying elements. In the case of cryomilled Al, the relative thermal stability exceeds that of Fe, probably due to the presence of oxides in addition to the nitrides
that form during milling. Once consolidated, the materials typically possess considerable strength compared to both conventional materials as well as to ne grained
materials processed by other methods.
While the as-milled powders are truly nanocrystalline, once consolidated the
materials are better described as ultrane-grained. The microstructures of the consolidated materials are fairly complex, and thus understanding of the mechanical
behavior remains tentative. The deformation mechanisms of materials with microstructures at both nanocrystalline and ultrane-grained scales are currently the subject of considerable attention, so as better models emerge for these grain sizes, the
better the behavior of the cryomilled materials will be understood.
Despite the complexity of the microstructure, existing data on Al and AlMg indicate that the HallPetch relationship describes the mechanical properties to a reasonable degree. This in turn suggests that dislocation interactions at grain boundaries
are responsible for the plastic deformation, and not alternative mechanisms such
as grain boundary sliding. Nevertheless, the exact nature of deformation has not
been identied, and is complicated by the wide distribution in grain sizes as well
as duplex microstructures. The large grained regions arise in two ways. The rst is
through grain growth during thermal treatment or thermomechanical processing,
in which some nanostructured grains grow to 500 nm or more, while the majority
of the grains only grow to 100 to 200 nm. The second is the result of consolidation

56

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

of the powders and lling in of the interparticle voids, resulting coarser grained regions that do not have a nanostructured history. The resulting microstructures range
from the core and mantle structure (NiAl) to a network of coarse grains dening
prior particle boundaries (Al10Ti2Cu and several of the Al 5083 extrusions discussed here). The prior particle boundaries can be disrupted to a signicant extent
during consolidation by HIPping, and while elongated coarse grained regions are
still found in these extrusions, the increased ductility can be attributed to the greater
degree of homogenization. Friction stir processing of cryomilled material that had
extensive coarse grained regions resulting from prior particle boundaries has demonstrated that high strength can be retained while simultaneously enhancing the
ductility.
The relationship between processing and properties has not been explored to an
extent necessary to guide optimization or tailoring of materials through consolidation and secondary processing steps. Extrusion temperature exerts a strong inuence
on the properties of Al alloys, and maintaining a low temperature during extrusion is
probably benecial to limiting grain growth. Higher extrusion ratios would probably
better homogenize several of the materials discussed here, and thus improve their
ductility, but greater reduction during the extrusion step would also require higher
temperatures. While existing research on powder metallurgy Al alloys may prove
useful guides to processing steps like degassing, the lower temperatures used in consolidation and extrusion for cryomilled Al will require wider experimental matrices.
The available literature does not allow a comprehensive comparison of cryomilled
materials and their mechanically alloyed analogues. For example, much of the discussion of the cryomilled Al has emphasized the grain size. Literature on mechanically alloyed Al indicates similarity in nal microstructure, but does not oer
enough detail to assess to what extent the addition of cryogenic liquid and resulting
formation of AlN justies the added expense of the process. From purely practical
considerations, cryomilling in a planetary ball mill may oer greater potential for
scale-up to manufacturing than shaker mills. Studies of milling time to achieve minimum grain size conducted on 500 or 1000 g batches of cryomilled powders may thus
be more instructive in conceiving of industrial processes than those conducted on
gram quantities of powders in high-speed shaker mills. The cryomilling technique
was originally developed with the idea of improving the dispersion-strengthened
properties as well as processing time relative to mechanical alloying. For the earlier
eorts on Fe-based and NiAl systems, the cryomilled materials showed improvements in mechanical properties in some cases, but not universally. The shorter processing time in terms of minimum grain size has been borne out in more recent
comparisons of dierent milling equipment and temperatures.
Preliminary investigations of other secondary processing methods, such as rolling
and forging of cryomilled Al alloys, are currently underway. These processes will
permit a greater variety of forms, such as plate and sheet, thus expanding the potential applications beyond that which can be gained from extrusion. These must be
accompanied by examination of high-temperature mechanical properties. For the
Al alloys, this is important from the perspective of determining best forming practices and limits, and not necessarily with an eye towards high-temperature service

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

57

conditions, as was the case for cryomilled Inconel and NiAl. In addition, dierent
ways to tailor the microstructure are being investigated. The observation that coarse
grains within the ne-grained matrix has improved ductility has prompted an eort
to fracture toughness by blending unmilled Al 5083 powder with cryomilled Al 5083
prior to consolidation and extrusion [139,140]. Such eorts are not unique to cryomilling, of course, as limited ductility in nanostructured materials has attracted
increasing attention as a research topic [69,141]. The increased strength gained from
grain size renement will ultimately be weighed against how well that strength can be
maintained during forming, whether the ductility can be improved to allow forming,
and the extra expense of additional processing compared to conventional materials.

Acknowledgements
Original research presented in this paper on cryomilled Al 5083 has been made
possible by nancial support from the Oce of Naval Research, under Contract
Numbers N00014-01-C-0149 and N00014-03-C-0163. We also thank Professor
R.S. Mishra, University of Missouri, Rolla, for sharing with us High-Resolution
TEM images shown in Fig. 4 prior to their publication. This manuscript beneted
signicantly from the critical reading and suggestions of A.P. Newbery.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]

Zhang X, Wang H, Narayan J, Koch CC. Acta Mater 2001;49:1319.


He J, Schoenung JM. Mater Sci Eng A 2002;336:274.
Ajdelsztajn L, Chung JLK, Bastian FL, Lavernia EJ. Metall Mater Trans A 2002;33:647.
Ajdelsztajn L, Picas JA, Kim GE, Bastian FL, Schoenung JM, Provenzano V. Mater Sci Eng A
2002;338:33.
Petkovic-Luton R, Vallone J. Exxon Research and Engineering Company, Florham Park, NJ,
United States of America US Patent Number 4619699; 1986.
Luton MJ, Jayanth CS, Disko MM, Matras S, Vallone J. Mater Res Soc Symp Proc 1989;132:79.
Benjamin JS, Volin TE. Met Trans 1974;5:1929.
Benjamin JS. Met Trans 1970;1:2943.
Gilman PS, Benjamin JS. Ann Rev Mater Sci 1983;13:279.
Benjamin JS, Bomford MJ. Metall Trans A 1977;8:1301.
Gleiter H. Prog Mater Sci 1989;33:223.
Aikin BJM, Dickerson RM, Jayne DT, Farmer S, Whittenberger JD. Scripta Metall Mater
1994;30:119.
Whittenberger JD, Arzt E, Luton MJ. J Mater Res 1990;5:2819.
Whittenberger JD, Arzt E, Luton MJ. J Mater Res 1990;5:271.
Garg A, Whittenberger JD, Aikin BJM. Mater Res Soc Symp Proc 1994;350:231.
Whittenberger JD, Garg A, Hebsur MG. J Mater Res 1999;14:2418.
Whittenberger JD, Arzt E, Luton MJ. Scripta Metall Mater 1992;26:1925.
Whittenberger JD, Luton MJ. J Mater Res 1992;7:2724.
Whittenberger JD, Luton MJ. J Mater Res 1995;10:1171.
Huang B, Vallone J, Luton MJ. Nanostruct Mater 1995;5:631.
Whittenberger JD, Noebe RD, Garg A. Metall Mater Trans A 1996;27:3170.
Whittenberger JD, Grahle P, Behr R, Arzt E, Hebsur MG. Mater Sci Eng A 2000;291:73.

58
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]

[66]

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160


Susegg O, Hellum E, Olsen A, Luton MJ. Philos Mag A 1993;68:367.
Gilman PS, Nix WD. Metall Trans A 1981;12:813.
Suryanarayana C. Prog Mater Sci 2001;46:1.
Fecht H-J. Nanostruct Mater 1995;6:33.
Xun Y, Lavernia EJ, Mohamed FA. Metall Mater Trans A 2004;35:573.
Koch CC. Nanostruct Mater 1997;9:13.
Eckert J, Holzer JC, Krill III CE, Johnson WL. J Mater Res 1992;7:1751.
Mohamed FA. Acta Mater 2003;51:4107.
Mohamed FA, Xun Y. Mater Sci Eng A 2003;354:133.
Zhou F, Liao XZ, Zhu YT, Dallek S, Lavernia EJ. Acta Mater 2003;51:2777.
Liao XZ, Huang JY, Zhu YT, Zhou F, Lavernia EJ. Philos Mag 2003;83:3065.
Zhou F, Witkin D, Nutt SR, Lavernia EJ. Mater Sci Eng A 2004;375377:917.
Zhou F, Nutt SR, Bampton CC, Lavernia EJ. Metall Mater Trans A 2003;34:1985.
Hansen N. Metall Mater Trans A 2001;32:2917.
Hughes DA, Hansen N. Acta Mater 1997;45:3871.
Wang J, Iwahashi Y, Horita Z, Furukawa M, Nemoto M, Valiev RZ, et al. Acta Mater
1996;44:2973.
He J, Lavernia EJ. J Mater Res 2001;16:2724.
Zhou F, Rodriguez R, Lavernia EJ. Mater Sci Forum 2002;386388:409.
Hellstern E, Fecht H-J, Fu Z, Johnson WL. J Appl Phys 1989;65:305.
Oleszak D, Shingu PH. J Appl Phys 1996;79:2975.
Lee J, Zhou F, Chung KH, Kim NJ, Lavernia EJ. Metall Mater Trans A 2001;32:3109.
Perez, RJ. PhD thesis, University of California, Irvine; 1997. 128p.
Perez RJ, Jiang HG, Dogan CP, Lavernia EJ. Metall Mater Trans A 1998;29:2469.
Malow TR, Koch CC. Acta Mater 1997;45:2177.
Zhou F, Lee J, Dallek S, Lavernia EJ. J Mater Res 2001;16:3451.
Hayes RW, Berbon PB, Mishra RS. Metall Mater Trans A 2004;35:3855.
Xun Y, Mohamed FA. Philos Mag A 2003;83:2247.
Chung KH, Lavernia EJ. Metall Mater Trans A 2002;33:3795.
Goujon C, Goeuriot P, Chedru M, Vicens J, Chermant JL, Bernard F, et al. Powder Tech
1999;105:328.
Goujon C, Goeuriot P, Delcroix P, Le Caer G. J Alloys Compd 2001;315:276.
Goujon C, Goeuriot P. Mater Sci Eng A 2001;315:180.
Goujon C, Goeuriot P. Mater Sci Eng A 2003;356:399.
Chung KH, He J, Shin DH, Schoenung JM. Mater Sci Eng A 2003;356:23.
Feldheim DL, Foss Jr CA, editors. Metal nanoparticles: synthesis, characterization and applications. New York: Marcel Dekker; 2002.
Rotello V, editor. Nanoparticles: building blocks for nanotechnology. New York: Kluwer Academic/
Plenum Press; 2004.
Chung KH, Lee J, Rodriguez R, Lavernia EJ. Metall Mater Trans A 2002;33:125.
Lau ML, Jiang HG, Perez RJ, Juarez-Islas J, Lavernia EJ. Nanostruct Mater 1996;7:847.
Moelle CH, Fecht H-J. Nanostruct Mater 1995;6:421.
Holzer JC, Birringer R, Eckert J, Krill III CE, Johnson WL. Mater Res Soc Symp Proc
1992;272:283.
Humphreys FJ, Hatherly M. Recrystallization and related annealing phenomena. Oxford: Elsevier
Scientic; 1996.
Milligan WW. In: Gerberich W, Yang W, editors. Interfacial and nanoscale fracture, vol.
8. Amsterdam: Elsevier Pergamon; 2003. p. 529.
Youngdahl CJ, Weertman JR, Hugo RC, Kung HH. Scripta Mater 2001;44:1475.
Mitra R, Ungar T, Morita R, Sanders PG, Weertman JR. In: Chung T-W, Dunand DC, Liaw PK,
Olson GB, editors. Advanced materials for the 21st century: the Julia R. Weertman Symposium. The Minerals, Metals and Materials Society; 1999. p. 553.
Tellkamp VL, Melmed A, Lavernia EJ. Metall Mater Trans A 2001;32:2335.

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160

59

[67] Wang Y, Chen M, Zhou F, Ma E. Nature 2002;419:912.


[68] Zhang X, Wang H, Scattergood RO, Narayan J, Koch CC, Sergueeva AV, et al. Appl Phys Lett
2002;81:823.
[69] Koch CC. Scripta Mater 2003;49:657.
[70] Melmed AJ, Tambakis NC, Lau M, Lavernia EJ. Progress Surf Sci 1998;59:313.
[71] Atkinson HV, Davies S. Metall Mater Trans A 2000;31:2981.
[72] Ashby MF. In: Jenkins I, Wood JV, editors. Powder metallurgy: an overview. London: The
Institute of Metals; 1991. p. 144.
[73] Lee Z, Rodriguez R, Hayes RW, Lavernia EJ, Nutt SR. Metall Mater Trans A 2003;34:1473.
[74] Witkin D, Lavernia EJ. In: Shaw LL, Suryanarayana C, Mishra RS, editors. Processing and
properties of structural nanomaterials. Chicago: TMS; 2003. p. 117.
[75] Han BQ, Lee Z, Nutt SR, Lavernia EJ, Mohamed FA. Metall Mater Trans A 2003;34:603.
[76] Berbon PB, Bingel WH, Mishra RS, Bampton CC, Mahoney MW. Scripta Mater 2001;44:61.
[77] Zhang Z, Zhou F, Lavernia EJ. Metall Mater Trans A 2003;34:1349.
[78] Hayes RW, Witkin D, Zhou F, Lavernia EJ. Acta Mater 2004;52:4259.
[79] Jia D, Ramesh KT, Ma E. Acta Mater 2003;51:3495.
[80] Kim Y-W, Grith WM, Froes FH. JOM 1985;37:27.
[81] Hansen N. Acta Metall 1970;18:137.
[82] Michels A, Krill III CE, Ehrhardt H, Birringer R, Wu DT. Acta Mater 1999;47:2143.
[83] Zhou F, Lee J, Lavernia EJ. Scripta Mater 2001;44:2013.
[84] Siegel RW. J Phys Chem Solids 1994;55:1097.
[85] Ghosh AK. In: Suresh S, Mortensen A, Needleman A, editors. Fundamentals of metalmatrix
composites. Boston: Butterworth-Heineman; 1993. p. 23.
[86] Hawk JA, Mirchandani PK, Benn RC, Wilsdorf HGF. In: Kim Y-W, Grith WM, editors.
Dispersion strengthened aluminum alloys. Warrendale, PA: TMS; 1988. p. 517.
[87] Morris DG, Munoz-Morris MA. Acta Mater 2002;50:4047.
[88] Dieter GE. Mechanical metallurgy. 2nd ed. New York: McGraw-Hill; 1976.
[89] Kim S-S, Haynes MJ, Ganglo RP. Mater Sci Eng A 1995;203:256.
[90] Rodriguez R, Hayes RW, Berbon PB, Lavernia EJ. Acta Mater 2003;51:911.
[91] Semiatin SL, Jatta KV, Uchic MD, Berbon PB, Matejczyk DE, Bampton CC. Scripta Mater
2001;44:395.
[92] Hayes RW, Rodriguez R, Lavernia EJ. Acta Mater 2001;49:4055.
[93] Han BQ, Lavernia EJ, Mohamed FA. Mater Sci Eng A 2003;358:318.
[94] Han BQ, Lavernia EJ, Mohamed FA. Philos Mag Lett 2003;83:89.
[95] Hayes RW, Tellkamp VL, Lavernia EJ. J Mater Res 2000;15:2215.
[96] Han BQ, Verma A, Mohamed FA, Lavernia EJ. In: Jin Z, Beaudoin A, Bieler TA, Radhakrishnan
R, editors. Hot deformation of aluminum alloys III. San Diego: TMS; 2003. p. 343.
[97] Hayes RW, Tellkamp VL, Lavernia EJ. Scripta Mater 1999;41:743.
[98] Tellkamp VL, Zhou F, Hayes R, Han BQ, Lavernia EJ. Mechanical behavior of nanostructured Al
synthesized by cryomilling and consolidation. In: NANO 2002: 6th International Conference on
Nanostructured Materials, Orlando; 2002.
[99] Wang YM, Ma E. Mater Sci Eng A 2004;375377:46.
[100] Bonetti E, Pasquini L, Sampaolesi E. Nanostruct Mater 1997;9:611.
[101] Hansen N. Met Trans 1970;1:545.
[102] Sun XK, Cong HT, Sun M, Yang MC. Metall Mater Trans A 2000;31:1017.
[103] Hansen N. Trans Metall Soc AIME 1969;245:1305.
[104] Hasegawa H, Komura S, Utsunomiya A, Horita Z, Furukawa M, Nemoto M, et al. Mater Sci Eng
A 1999;265:188.
[105] Hansen N. Acta Metall 1969;17:637.
[106] Mukai T, Ishikawa K, Higashi K. Mater Sci Eng A 1995;204:12.
[107] Lloyd DJ. Met Sci 1980;14:193.
[108] Phillips VA, Swain AJ, Eborall R. J Inst Metals 1953;81:625.
[109] Lloyd DJ, Court SA. Mater Sci Tech 2003;19:1349.

60
[110]
[111]
[112]
[113]
[114]
[115]
[116]
[117]
[118]
[119]
[120]
[121]
[122]
[123]
[124]
[125]
[126]
[127]
[128]
[129]
[130]
[131]
[132]
[133]
[134]
[135]
[136]
[137]
[138]
[139]
[140]
[141]
[142]

D.B. Witkin, E.J. Lavernia / Progress in Materials Science 51 (2006) 160


Wilsdorf HGF, Kuhlman-Wilsdorf D. Mater Sci Eng A 1993;164:1.
Rosler J, Joos R, Arzt E. Metall Mater Trans A 1992;23:1521.
Benjamin JS, Schelleng RD. Metall Trans A 1981;12:1827.
Hull D, Bacon DJ. Introduction to dislocations. 4th ed. Butterworth-Heinemann; 2001.
Pearson WB. A handbook of lattice spacings and structures of metals and alloys, 2. London: Pergamon Press; 1967.
Lloyd DJ, Court SA, Gatenby KM. Mater Sci Tech 1997;13:660.
Kawazoe M, Shibata T, Mukai T, Higashi K. Scripta Mater 1997;36:699.
Markushev MV, Murashkin MY, Prangnell PB, Gholinia A, Maiorova OA. NanoStruct Mater
1999;12:839.
Chang S-Y, Lee JG, Park K-T, Shin DH. Mater Trans 2001;42:1074.
Wilsdorf HGF. In: Kim Y-W, Grith WM, editors. Dispersion strengthened aluminum alloys.
Warrendale, PA: TMS; 1988. p. 3.
Last HR, Garret RK. Metall Mater Trans A 1996;27:737.
Kim Y-W, Bidwell LR. Scripta Metall 1982;16:799.
Kang SK, Erich DL, Merrick HF. The mechanical behavior of mechanically alloyed Al alloys.
TMS-AIME; 1982. p. 317.
Robinson JM. Mater Sci Eng A 1995;203:238.
Thomas AT. Acta Metall 1966;14:1363.
Hall EO. Yield point phenomena in metals and alloys. London: MacMillan and Company; 1970.
Cottrell AH. Philos Mag 1953;44:829.
Mulford RA, Kocks UF. Acta Mater 1979;27:1125.
Charnock W. Philos Mag A 1969;20:427.
King JE, You CP, Knott JF. Acta Metall 1981;29:1553.
McCormick PG. Scripta Metall 1970;4:221.
McCormick PG. Acta Metall 1971;19:463.
Hirth JP, Lothe J. Theory of dislocations. 2nd ed. New York: John Wiley and Sons; 1982.
Huang JY, Liao XZ, Zhu YT, Zhou F, Lavernia EJ. Philos Mag 2003;83:1407.
Cheng S, Spencer JA, Milligan WW. Acta Mater 2003;51:4505.
Carsley JE, Fisher A, Milligan WW, Aifantis EC. Metall Mater Trans A 1998;29:2261.
Lloyd DJ, Morris LR. Acta Metall 1977;25:857.
Morris DG, Morris MA. Acta Metall Mater 1991;39:1763.
Li JCM. Trans Metall Soc AIME 1963;227:239.
Witkin D, Lee Z, Rodriguez R, Nutt SR, Lavernia EJ. Scripta Mater 2003;49:297.
Pao PS, Jones HN, Feng CP. Mater Res Soc Symp Proc 2003;791:17.
Koch CC, Morris DG, Lu K, Inoue A. Mater Res Soc Bull 1999;24:54.
Brandes EA, editor, Smithells metals reference book. London: Butterworths; 1983.

Das könnte Ihnen auch gefallen