You are on page 1of 10

10-03-80A-4.

qxd 27/4/04 17:24 Page 1735

Achieving Enhanced Ductility in a Dilute Magnesium


Alloy through Severe Plastic Deformation
KIYOSHI MATSUBARA, YUICHI MIYAHARA, ZENJI HORITA,
and TERENCE G. LANGDON

Experiments were conducted to evaluate the utility of a new processing procedure developed for
Mg-based alloys in which samples are subjected to a two-step processing route of extrusion followed
by equal-channel angular pressing (designated as EX-ECAP). The experiments were conducted
using a Mg-0.6 wt pct Zr alloy and, for comparison purposes, samples of pure Mg. It is shown that
the potential for successfully using ECAP increases in both materials when adopting the EX-ECAP
procedure. For the Mg-Zr alloy, the use of EX-ECAP produces a grain size of ⬃1.4 ␮m when the
pressing is undertaken at 573 K. By contrast, using EX-ECAP with pure Mg at 573 K produces a
grain size of ⬃26 ␮m. Tensile testing of the Mg-Zr alloy at 523 and 573 K after processing by
EX-ECAP revealed the occurrence of significantly enhanced ductilities with maximum elongations
of ⬃300 to 400 pct.

I. INTRODUCTION bulk material,[17] and high strains may be imposed through


the use of a multipass die[18] or by using a rotary die[19] or
THE low density and good machinability of magnesium a side-extrusion facility.[20] In addition, it has been demon-
alloys makes them attractive for a wide range of structural
strated that the principles of ECAP processing may be incor-
applications in transportation and, especially, in the automotive
porated into conventional rolling using the conshearing[21]
field.[1,2] These alloys also exhibit excellent damping capac-
or continuous confined strip-shearing (C2S2)[22–25] processes.
ities, so that they are especially attractive for use in appli-
In contrast to the work with fcc metals, there has been
cations where it is necessary to dampen external vibrations,
less success when using ECAP processing with magnesium
as in helicopters and satellites. These beneficial features have
and Mg-based alloys. Although there are several reports of
led to the prediction by Friedrich and Schumann[3] of a “new
the application of ECAP to a range of Mg alloys,[26–30] these
age of magnesium.”
investigations generally relate to the use of fairly complex
Despite the inherent potential advantages of magnesium-
alloy systems where there are widespread dispersions of inter-
based alloys, their hexagonal crystal structure provides only
metallic precipitates. By contrast, the processing by ECAP of
a limited number of active slip systems, giving low levels
as-cast pure Mg and dilute Mg solid-solution alloys has been
of ductility and, consequently, difficulties in forming com-
generally unsuccessful, and ultrafine-grained microstructures
plex components. It is well established in recent work that
have not been achieved. For example, pure Mg was subjected
the ductilities of metallic alloys may be significantly
to ECAP through two passes at 673 K to give a measured
enhanced, often to the superplastic range, by processing
as-pressed grain size of ⬎100 ␮m, and there was a meas-
the alloys through the introduction of severe plastic deform-
ured grain size of ⬃17 ␮m in an Mg-0.9 pct Al solid-solution
ation (SPD).[4] Conventional methods of SPD processing
alloy after ECAP through two passes at 473 K,[31] where this
include equal-channel angular pressing (ECAP),[5,6,7] where
and all subsequent alloy compositions are given in weight
a sample is pressed repetitively through a die confined within
percent. It is apparent, therefore, that both of these grain sizes
a channel bent through an angle at, or close to, 90 deg, and
are outside of the range of conventional ultrafine-grained
high-pressure torsion (HPT),[8,9,10] where a disk is subjected
materials, so that, although the experiments on pure Mg and
to a high pressure and concurrent torsional straining. The
the Mg-0.9 pct Al alloy demonstrated an improvement in the
procedures of ECAP and HPT are both effective in pro-
mechanical properties of these materials at room temperature,
ducing substantial grain refinement in fcc metals such as
the as-pressed grain sizes were too large to give enhanced or
pure aluminum,[11,12] pure nickel,[9,10] and a range of alu-
superplastic ductilities when testing at elevated temperatures.
minum alloys,[13–16] but, in practice, ECAP appears to be the
The present investigation was motivated by recent evidence
more useful technique because it utilizes fairly large sam-
that very significant grain refinement may be introduced in a
ples, it can be scaled up relatively easily to produce large
Mg-0.6 pct Zr alloy through the use of a two-step processing
procedure in which the alloy is first extruded and then sub-
jected to ECAP processing.[32] This two-step process of extru-
KIYOSHI MATSUBARA and YUICHI MIYAHARA, Graduate Stu- sion and subsequent ECAP has been given the acronym
dents, and ZENJI HORITA, Professor, are with the Department of Mater- EX-ECAP, and recent experiments have shown that, for the
ials Science and Engineering, Faculty of Engineering, Kyushu University, same conditions of temperature and strain rate, this process
Fukuoka 812-8581, Japan. TERENCE G. LANGDON, Professor, is with is capable of increasing the ductility of a Mg-9 pct Al alloy
the Departments of Aerospace & Mechanical Engineering and Materials
Science, University of Southern California, Los Angeles, CA 90089-1453.
from ⬃30 pct in the as-cast condition to ⬃110 pct after casting
Contact e-mail: langdon@usc.edu plus extrusion and up to ⬃840 pct after casting and process-
Manuscript submitted February 18, 2003. ing by EX-ECAP.[33]

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 35A, JUNE 2004—1735


10-03-80A-4.qxd 27/4/04 17:24 Page 1736

The basic principles underlying the EX-ECAP process are prepared by casting into an ingot of 60 mm in diameter. The
threefold. First, early experiments showed that it was possible grain size of the cast Mg was estimated to be ⬃1.4 mm, and
to use conventional extrusion in order to achieve significant analysis revealed the presence of 0.02 wt pct Al as the only
grain refinement in commercial-purity Mg.[34] Second, it is measurable impurity.
well known that the application of extrusion to Mg-based The melting temperature of the Mg-Zr alloy was deter-
alloys is often effective in promoting superplastic character- mined using differential scanning calorimetry (DSC). A small
istics in these materials,[35] primarily because the extrusion sample of the alloy, with a weight of ⬃50 mg, was heated
step produces some limited grain refinement, in addition to a in an argon atmosphere together with a Mo standard within
strong texture where the basal planes become reasonably a MacScience DSC3300S apparatus, using a heating rate of
aligned.[36,37] Third, by applying ECAP to samples in the 10 K min⫺1. The incipient melting temperature of the alloy
extruded condition, it is reasonable to anticipate that there may was recorded as 920 K.
be an additional grain refinement and, thus, an ability to attain To evaluate the effect of incorporating extrusion into the
even higher superplastic elongations. Preliminary experiments processing procedure, parts of the Mg-Zr alloy and the pure
on a Mg-0.6 pct Zr alloy showed that this two-step EX-ECAP Mg were extruded at a speed of 5 mm s⫺1 and a tempera-
procedure was effective in reducing the grain size to ⬃1 ␮m ture of 623 K into rods having a diameter of 10 mm, where
through only a single pass of ECAP at 573 K, and the as- this extrusion condition corresponds to a reduction ratio of
pressed alloy then exhibited superplastic characteristics when 36:1. These materials are, henceforth, designated as the
testing in tension at 573 K.[32] “cast ⫹ extruded” condition.
The earlier exploratory work on the Mg-0.6 pct Zr alloy For ECAP processing, samples were prepared with dia-
was designed only to evaluate the potential for using an meters of 10 mm and lengths of 60 mm, and the pressing
EX-ECAP processing route to achieve an ultrafine microstruc- was conducted using a solid die with an angle between
ture and good tensile ductilities in a dilute magnesium-based the channels of ⌽ ⫽ 90 deg and an outer arc of curvature
alloy. It was recognized, however, that a detailed investigation of ⌿ ⬇ 20 deg at the point where the two parts of the
was required in order to more fully characterize the poten- channel intersect. It can be shown from first principles that
tial of this processing procedure. Accordingly, the present these internal angles give an imposed strain of ⬃1 on each
investigation was initiated with four specific objectives. The passage of the sample through the die,[40] and experiments
first was to evaluate the feasibility of using the EX-ECAP have demonstrated that an arc of curvature of ⌿ ⬇ 20 deg
procedure on the Mg-0.6 pct Zr alloy with the ECAP con- has no significant influence on the subsequent homogene-
ducted at different temperatures and through different numbers ity of the microstructure.[41] When samples were pressed
of passes, so that varying amounts of strain are introduced repetitively through the die, each sample was rotated by
in each pressing operation. The second was to examine the 90 deg in the same sense between each pass in the pro-
thermal stability of any ultrafine-grained microstructure intro- cessing procedure designated as route BC.[42,43] All pressings
duced into the alloy. The third objective was to evaluate the were conducted at elevated temperatures, with the sam-
potential for achieving enhanced ductility or superplastic ples held at temperature within the die for ⬃10 minutes
elongations in the Mg-0.6 pct Zr alloy after processing by prior to the first pass and then, when repetitive pressings
EX-ECAP. The fourth was to make a direct comparison were undertaken, they were held at temperature within the
between the microstructures achieved in the Mg-0.6 pct Zr die for ⬃1 minute before pressing.
alloy using EX-ECAP with those obtained when using the Attempts were made to press the Mg-Zr alloy for up to four
same processing route with pure Mg. passes using temperatures in the range from 473 to 573 K.
Preliminary experiments showed that higher temperatures
were needed for the pressing of pure Mg, and, for this mater-
II. EXPERIMENTAL MATERIALS ial, the pressing temperatures ranged from 573 to 673 K.
AND PROCEDURES Some samples were subjected to ECAP after casting but with-
out an intermediate extrusion; these samples are designated
The material used in this investigation was from a dif- as “cast ⫹ equal-channel angular pressed.” Other samples
ferent batch, but of the same nominal composition, as the were prepared by casting, extruded to the diameter required
Mg-0.6 wt pct Zr alloy documented in an earlier report.[32] for ECAP, and then subjected to ECAP at selected temper-
As previously, this alloy was selected specifically because atures through up to a maximum of four passes: the latter
it has been demonstrated that this represents the optimum samples represent the EX-ECAP process, and these samples
composition in order to achieve a very high damping capa- are designated as “cast ⫹ extruded ⫹ equal-channel angular
city.[38] The alloy was produced through a conventional melt- pressed.”
ing and casting process, and it was supplied in the form of The thermal stability of the microstructure was exam-
a cylindrical ingot having a diameter of 60 mm and a length ined for samples of the Mg-Zr alloy subjected to EX-ECAP
of ⬃100 mm. A chemical analysis of the cast alloy gave processing by slicing samples with thicknesses of ⬃0.4 mm
a zirconium content of ⬃0.7 wt pct with an addition of perpendicular to the longitudinal axes after EX-ECAP and
⬃0.01 wt pct Al but with no other measurable impurities. then sealing these disks in glass tubes in an argon environ-
Since the maximum solubility of Zr in cast Mg is ⬃0.3 wt pct ment and annealing for 1 hour at selected temperatures up
from room temperature to ⬃573 K, it is apparent that the to a maximum of 673 K.
alloy contained a small quantity of an ␣-Zr phase.[39] The microstructures of the Mg-Zr alloy and the pure Mg
Inspection showed the measured grain size in the cast alloy were subject to extensive examination using combinations
to be ⬃70 ␮m. For comparison purposes, some additional of optical microscopy (OM), transmission electron microscopy
experiments were conducted using pure Mg, which was also (TEM), and electron-probe microanalysis (EPMA). For

1736—VOLUME 35A, JUNE 2004 METALLURGICAL AND MATERIALS TRANSACTIONS A


10-03-80A-4.qxd 27/4/04 17:24 Page 1737

inspection by OM, samples were cut parallel to the solidifi- be rich in Zr. The presence of these Zr-rich particles is effective
cation direction after casting or parallel to the extrusion in retaining a grain size of only ⬃70 ␮m in the cast condi-
direction in the cast ⫹ extruded or cast ⫹ extruded ⫹ equal- tion, and the grain size is further reduced to ⬃11 ␮m in the
channel angular pressed conditions, and they were ground cast ⫹ extruded condition, where there is again an array of
mechanically to a mirrorlike surface using abrasive papers reasonably equiaxed grains suggesting the occurrence of recrys-
and alumina powders. These surfaces were then etched using tallization. As is evident from Figure 1, dark bands are also
a solution of 1 pct HNO3, 24 pct C2H5OH, and 75 pct H2O, visible in the alloy in the cast ⫹ extruded condition, with the
with an immersion time of ⬃15 to 20 seconds. For TEM, bands lying essentially parallel to the extrusion direction: it
the extruded or pressed rods were sliced perpendicular to their was shown using EPMA that these bands are also rich in Zr,
longitudinal axes to thicknesses of ⬃0.4 mm, cut and ground and it is reasonable to conclude they are instrumental in retain-
to disks with diameters of 3 mm and thicknesses of ⬃0.15 mm, ing the grain size at this very low level.
and then thinned to perforation at room temperature in a
twin-jet polishing facility using a solution of 15 pct HClO4,
15 pct C3H8O3, and 70 pct C2H5OH. Following perforation, B. Nature of the Microstructures after ECAP
they were ion milled in an argon environment. The TEM A detailed overview of the ECAP experiments is given
observations were conducted using an Hitachi H-8100 electron in Table I for both pure Mg and the Mg-Zr alloy: this tab-
microscope operating at 200 kV, and selected-area electron ulation distinguishes between samples pressed successfully
diffraction (SAED) patterns were recorded from regions hav- without any cracking (denoted by X), samples pressed but
ing diameters of 6.2 ␮m. The microanalysis was undertaken showing visible macroscopic surface cracking (denoted by
using a Shimadzu EPMA-1600 facility operating at 15 kV A), and samples which broke during the pressing (denoted
with a probe current of ⬃10 nA, and X-ray maps were col- by B), where the other possible pressing conditions were
lected over square areas of 80 ⫻ 80 ␮m2. not attempted. This tabulation confirms the advantage of
The tensile properties of the Mg-Zr alloy were evaluated introducing the extrusion step between casting and ECAP.
by machining tensile specimens with gage lengths of 5 mm Thus, the Mg-Zr alloy was successfully pressed through
and cross-sectional dimensions of 2 ⫻ 3 mm2. These speci- four passes at 573 K in the EX-ECAP condition, but the
mens were prepared with their tensile axes cut parallel to the alloy broke on the second pass when pressed at this tem-
longitudinal axes after extrusion or EX-ECAP and parallel perature in the cast condition without any extrusion. Simi-
to the solidification direction in the cast condition. Tensile larly, the pure Mg was successfully pressed through two
tests were conducted at 523 and 573 K using initial strain passes without cracking at 573 and 623 K using the EX-
rates in the range from 3.3 ⫻ 10⫺4 to 3.3 ⫻ 10⫺1 s⫺1 and ECAP procedure, whereas, when pressing the cast mater-
with a testing machine operating at a constant rate of crosshead ial at these two temperatures, the samples broke in the
displacement. All tensile tests were taken to failure to pro- first and second pass, respectively. The additional strains
vide information on the total elongations. The values of the that may be imposed through ECAP in the cast ⫹ extruded
strain-rate sensitivity (m) were determined for these two testing samples is attributed to the significant grain refinement
temperatures, where m is defined as occurring during the extrusion step. It is apparent also that
the introduction of 0.6 pct Zr leads to a significant increase
⭸ ln s in the pressing capability by comparison with pure Mg, and
m⫽ # [1]
⭸ ln ␧ this is also associated with the greater grain refinement that
# may be achieved in the dilute alloy.
where ␴ is the flow stress and ␧ is the imposed strain rate.[44]
Microstructural information for the Mg-Zr alloy is shown
The stress-strain curves were examined to determine the
in Figure 2, where X-ray maps are presented for the cast,
values of the flow stresses at a strain of 0.1 for each test-
extruded, and EX-ECAP condition, where the latter corres-
ing condition, and these data were then plotted logarithmi-
ponds to a sample pressed under the optimum condition of
cally as flow stress against strain rate to provide estimates
four passes at a temperature of 573 K. These maps were
of the values for m.
constructed using, for the same field of view for any condi-
tion, the Mg K␣ (upper row) and Zr L␣ (lower row) lines in
the EPMA, and the gradations in color from blue through green
III. EXPERIMENTAL RESULTS to yellow and red depict the increasing local enrichments of
the selected elements. Thus, it is apparent that there are Zr-
A. Microstructures after Casting and Extrusion rich regions where Mg is depleted in all three conditions, but
Figure 1 shows optical micrographs of the pure Mg (on the these Zr-rich regions are well defined and reasonably rounded
left-hand side) and the Mg-0.6 pct Zr alloy (on the right-hand in cross section in the cast condition and they are well-defined
side) in the as-cast and cast ⫹ extruded conditions, respec- and elongated in the extrusion direction after the extrusion
tively: the extrusion direction is horizontal for the samples step, but they become diffuse and have ill-defined boundaries
subjected to extrusion. After casting, many twins are visible after the subsequent ECAP.
in the pure Mg, but extrusion leads to an array of equiaxed The microstructure of pure Mg after EX-ECAP is shown
grains indicating that recrystallization probably occurs during in Figure 3, where the sample was pressed for two passes
extrusion at 623 K. Measurements showed that the aver at 573 K. There is an equiaxed array of grains in this sample,
age grain size was reduced from ⬃1.4 mm after casting to with an average size of ⬃26 ␮m. Thus, EX-ECAP refines
⬃55 ␮m in the cast ⫹ extruded condition. Inspection shows the grain size in pure Mg, but it is not capable, at least at
also that the cast Mg-Zr alloy contains second-phase particles, 573 K, of reducing the grain size to the submicrometer level.
and examination by EPMA demonstrated these particles to Twinning is also evident in this microstructure, but the twins

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 35A, JUNE 2004—1737


10-03-80A-4.qxd 27/4/04 17:24 Page 1738

Fig. 1—Optical micrographs of pure Mg (on left) and the Mg-0.6 pct Zr alloy (on right) in the as-cast and cast ⫹ extruded conditions: the extrusion direction
is indicated.

Table I. Summary of the Effect of Applying ECAP to Pure Mg and the Mg-0.6 pct Zr Alloy in the Cast
and the Cast ⫹ Extruded Conditions

Cast Cast ⫹ Extruded


ECAP at T (K) ECAP at T (K)
Number
Material of Passes 573 623 673 473 523 573 623
Pure Mg 1 B X X X X
2 B X X X
3 B B B
Mg-0.6 pct Zr 1 X B A X
2 B B X
3 X
4 X
X ⫽ successful pressing without any visible cracking; A ⫽ evidence for macroscopic surface cracking after ECAP; and B ⫽ specimen broken during ECAP.

are less clearly defined than in the as-cast structure shown Figure 4 shows examples of the microstructures in the alloy
in Figure 1. By contrast, the microstructures of the Mg-Zr after EX-ECAP with the pressing conducted at 573 K through
alloy after EX-ECAP were more representative of those one, two, and four passes, respectively, together with the
achieved in the ECAP of fcc metals. appropriate SAED patterns. After one pass, the boundaries

1738—VOLUME 35A, JUNE 2004 METALLURGICAL AND MATERIALS TRANSACTIONS A


10-03-80A-4.qxd 27/4/04 17:24 Page 1739

Fig. 2—EPMA for the Mg-0.6 pct Zr alloy showing the three conditions of as-cast, cast ⫹ extruded and cast ⫹ extruded–equal-channel angular pressed:
the upper and lower rows show the characteristics of Mg K␣ and Zr L␣, respectively.

and these types of boundaries are generally considered to be


in high-energy, nonequilibrium configurations.[14,45–47] The
microstructure is inhomogeneous after a single pass, with
some areas showing arrays of high-angle boundaries, as
demonstrated by the SAED pattern taken at point (a), and
other areas showing arrays of subgrains with low-angle bound-
aries, as demonstrated by the pattern taken at point (b). The
diffuse and ill-defined nature of the microstructure after one
pass makes it impossible to provide a quantitative estimate of
the relevant proportions of the regions containing predomi-
nantly high- and low-angle boundaries, but, nevertheless, the
overall impression was that both regions were present through-
out the sample. After two passes, there are well-defined
equiaxed grains in many regions, but also some areas of poorly
defined boundaries with high dislocation densities. The mea-
sured grain size in this condition was ⬃1.3 ␮m, and it is
evident from the SAED pattern in Figure 4 that these grain
boundaries have predominantly high angles of misorientation.
Fig. 3—Microstructure of pure Mg after extrusion and ECAP through two
passes at 573 K.
A well-defined microstructure of equiaxed grains was visi-
ble throughout most of the material after four passes, and there
were relatively few dislocations within the grains. The grain
tend to be poorly defined, and there is a very high disloca- size in this condition was ⬃1.4 ␮m, and the SAED pattern
tion density: microstructures with similar characteristics were in Figure 4 suggests the presence of predominantly high-angle
reported earlier for various fcc metals after SPD processing, boundaries. In practice, a grain size of ⬃1.4 ␮m is slightly

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 35A, JUNE 2004—1739


10-03-80A-4.qxd 27/4/04 17:24 Page 1740

Fig. 4—Microstructures in the Mg-0.6 pct Zr alloy after one, two, and four passes at 573 K, plus appropriate SAED patterns: the two SAED patterns for
one pass relate to the areas marked (a) and (b).

outside of the range of conventional ultrafine-grained materials,


where the grain sizes are usually defined as ⬍1 ␮m.

C. Thermal Stability of the Ultrafine-Grained


Microstructure
The preceding observations show that the EX-ECAP process
is effective in reducing the grain size to close to ⬃1 ␮m in
the Mg-0.6 pct Zr alloy, but any attempts to achieve high
tensile ductilities at elevated temperatures requires also
that these very fine grains are reasonably stable at high tem-
peratures. Figure 5 shows results obtained from annealing
samples for 1 hour at selected temperatures up to a maxi-
mum of 673 K: these results were obtained with samples
prepared by EX-ECAP including four passes at 573 K. It is
apparent from these measurements that the fine-grained struc-
ture is reasonably stable up to 573 K, but some limited grain
growth occurs at higher temperatures. At 623 K, it was
observed that fine grains having sizes of ⬍2 ␮m covered
an estimated ⬃30 pct of the sample, and these smaller grains
tended to be retained along longitudinal stringers lying par- Fig. 5—Grain size vs temperature for samples of the Mg-Zr alloy processed
by extrusion and ECAP for four passes at 573 K and then annealed for 1 h
allel to the extrusion and pressing directions, as illustrated at selected temperatures.
in Figure 6 where an arrow delineates some of the smaller
grains present after annealing at 623 K. A comparison with Zr-rich, thereby confirming that the presence of the Zr-rich
the EPMA Zr mapping in Figure 2 suggests that these smaller particles prevents significant grain growth. These observa-
grains are retained preferentially in those regions that are tions suggest, therefore, that it would be beneficial to achieve

1740—VOLUME 35A, JUNE 2004 METALLURGICAL AND MATERIALS TRANSACTIONS A


10-03-80A-4.qxd 27/4/04 17:24 Page 1741

Fig. 6—Microstructure of the Mg-0.6 pct Zr alloy after preparing by EX-


ECAP including four passes at 573 K and then annealing for 1 h at 623 K:
the arrow points to smaller grains contained in longitudinal stringers lying Fig. 8—Elongation to failure vs strain rate for the Mg-0.6 pct Zr alloy at
horizontal and parallel to the extrusion and pressing directions. 573 K for cast ⫹ extruded samples in an unpressed condition and for
samples pressed at 573 K through one, two, and four passes.

were pressed through only one pass at 573 K. All samples


were pulled to failure at 573 K, and the results are shown
in Figure 7. These results are quite similar to those reported
in the earlier investigation using a Mg-0.6 pct Zr alloy, but
where the alloy was obtained from a different batch.[32]
The overall similarity between these two batches of alloys
confirms the general reproducibility of the data.
It is evident from Figure 7 that the cast alloy exhibits only
limited ductility, but the elongation to failure is increased to
a small extent by applying ECAP, is increased to a greater
extent if the cast alloy is extruded, and is increased even fur-
ther when the material is processed using both extrusion and
ECAP in the EX-ECAP procedure. Thus, the imposition of
only a single pass of ECAP after extrusion has a very sig-
nificant effect on the elongation to failure, despite the fact
that the microstructure after one pass is ill-defined and con-
tains a large number of dislocations within the grains. These
results confirm, therefore, the importance of using a two-step
processing route when working with Mg-based alloys.
Fig. 7—Elongation to failure vs strain rate for the Mg-0.6 pct Zr alloy at The significance of the number of passes on the measured
573 K in four different processing conditions: in the as-received cast con- elongations to failure at 573 K is shown in Figure 8, where
dition, in the cast ⫹ extruded condition, in the cast ⫹ extruded ⫹ equal- results are presented for extruded samples in an unpressed
channel angular pressed condition and in the cast ⫹ equal-channel angular
pressed condition without any intermediate extrusion.
condition plus for extruded samples taken through one, two,
and four passes of ECAP at 573 K. In general, it appears that
the total elongation tends to increase slightly with increas-
a homogeneous dispersion of Zr-rich particles throughout ing numbers of passes, and this trend is especially evident
the material, since this would assist in attaining a uniform at the faster strain rates. This increase is attributed, as shown
distribution of ultrafine grains in the processed alloy. in Figure 4, to the more homogeneous array of equiaxed
grains present in the alloy after the larger number of passes.
Taking the measured flow stress for each sample at a strain
D. Mechanical Properties at Elevated Temperatures
(␧) of 0.1, Figure 9 shows a logarithmic plot of flow stress
Samples were tested in tension under four different vs strain rate for the same samples recorded in Figure 8. Thus,
microstructural conditions: in the as-received cast condition; the flow stress is significantly reduced through the applica-
in the cast ⫹ extruded condition; in the cast ⫹ equal-channel tion of ECAP and, since these results were obtained at a high
angular pressed condition, where the ECAP was conducted temperature in a region where diffusion-controlled processes
for one pass at 573 K without any intermediate extrusion; are dominant, this trend is consistent with the smaller
and in the cast ⫹ extruded ⫹ equal-channel angular pressed grain sizes in these samples. The strain-rate sensitivity (m)
condition where, in order to provide consistent data with the was recorded as ⬃0.4 at the lower strain rates where the
cast ⫹ equal-channel angular pressed specimen, the samples elongations are a maximum for those samples subjected to

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 35A, JUNE 2004—1741


10-03-80A-4.qxd 27/4/04 17:24 Page 1742

Fig. 9—Flow stress at a strain of 0.1 vs strain rate for the samples shown Fig. 11—Flow stress at a strain of 0.1 vs strain rate for the samples shown
in Fig. 8: the strain-rate sensitivity under conditions of maximum ductil- in Fig. 10: the strain-rate sensitivity under conditions of maximum ductil-
ity is ⬃0.4. ity is ⬃0.3.

these data, Figure 11 shows a logarithmic plot of flow stress


against strain rate, where the flow stress was recorded at a
strain of ␧ ⫽ 0.1. Under these conditions, the flow stress
again decreases with increasing numbers of passes in ECAP,
and there is a maximum value of m ⬇ 0.3 for the sample
subjected to four passes of ECAP when tested at the two
lowest strain rates.

IV. DISCUSSION
These experiments provide important information on the
processing of dilute Mg-based alloys through ECAP, and
they lead to four significant conclusions.
First, the EX-ECAP procedure is remarkably effective in
increasing the potential for applying the ECAP process to
both pure Mg and a dilute Mg-based alloy. It is difficult to
use ECAP processing with pure Mg, but, nevertheless, as
shown in Table I, the introduction of an intermediate extru-
Fig. 10—Elongation to failure vs strain rate for the Mg-0.6 pct Zr alloy at
sion increases the viability of the ECAP process so that sam-
523 K for unpressed samples and for samples pressed at 573 K through ples can be pressed through two passes without exhibiting
one, two, and four passes. any deleterious cracking. For the dilute Mg-0.6 pct Zr alloy,
the use of EX-ECAP increases the viability of the pressing
EX-ECAP through four passes. This value of m is interme- procedure at 573 K from a single pass to four passes.
diate between the values of m of ⬃0.5 and ⬃0.3 reported for Second, the microstructures are substantially refined when
fcc aluminum-based alloys tested after ECAP for conditions an extrusion step is introduced prior to ECAP. Thus, it is
where the rate-controlling processes were separately attrib- possible to achieve a grain size of ⬃26 ␮m in pure Mg
uted to superplasticity and dislocation glide, respectively.[48,49] after EX-ECAP, as given in Figure 3, whereas, in the
Similar results are shown in Figure 10 for the same set absence of an extrusion step, an earlier report showed the
of samples, but with the testing conducted at a lower tem- application of ECAP to the cast alloy gave an as-pressed
perature of 523 K. These results match those obtained at grain size of ⬎100 ␮m.[31] In the Mg-0.6 pct Zr alloy,
573 K in Figure 8, except that the total elongations tend to processing by EX-ECAP produced a grain size of ⬃1.4 ␮m,
be slightly lower at all strain rates. The highest elongation which is within the range of ultrafine grain sizes generally
at this temperature is ⬃320 pct when using an initial strain associated with the processing of fcc metals using severe
rate of 3.3 ⫻ 10⫺4 s⫺1 after EX-ECAP through four passes. plastic deformation: for example, experiments have recorded
When it is noted that the melting temperature of the alloy a grain size of ⬃1.3 ␮m in pure aluminum after conven-
was measured as 920 K, it is apparent that these tests relate tional ECAP.[12] It is apparent from the TEM observations
to a homologeous testing temperature of only 0.57 Tm, where and associated EPMA that the grains are much smaller in
Tm is the absolute melting temperature of the alloy. Using the Mg-0.6 pct Zr alloy by comparison with pure Mg

1742—VOLUME 35A, JUNE 2004 METALLURGICAL AND MATERIALS TRANSACTIONS A


10-03-80A-4.qxd 27/4/04 17:24 Page 1743

because of the presence of Zr-rich bands and Zr particles of ⬃420 pct was recorded for the same alloy at the same
that inhibit grain growth. testing temperature and strain rate in an earlier investiga-
Third, the ultrafine grain size attained in the Mg-0.6 pct Zr tion, when the material was also subjected to one pass at
alloy is reasonably stable at high temperatures, at least up to 573 K.[32] It is reasonable to speculate that these higher duc-
⬃550 K. As illustrated in Figure 5, a grain size of ⬍10 ␮m tilities may be a consequence of an inhibition in grain growth
is retained in the alloy up to temperatures of at least ⬃600 K. during tensile testing in samples having inhomogeneous
This thermal stability provides an opportunity for achieving microstructures with a preponderance of high-energy, non-
enhanced ductilities in tensile tests at elevated temperatures. equilibrium boundaries, but more experiments are needed
Fourth, the maximum elongations achieved in the Mg- to clarify the significance of these exceptionally high elon-
0.6 pct Zr alloy lie typically in the range of ⬃300 to 400 pct, gations at the slowest experimental strain rate.
as shown in Figures 8 and 10, with higher elongations recorded
at the higher testing temperature of 573 K. These elongations
are generally not high by comparison with the very large V. SUMMARY AND CONCLUSIONS
superplastic elongations recorded in fcc metals after conven-
tional ECAP without an extrusion step: for example, elong- 1. Experiments were conducted to examine the feasibility
ations of ⬎2000 pct have been reported during tensile of achieving an ultrafine-grained structure in a Mg-0.6
testing of an Al-3 pct Mg-0.2 pct Sc alloy after ECAP.[48] pct Zr alloy using ECAP. For comparison purposes, addi-
Nevertheless, the present elongations compare very favorably tional experiments were also conducted using samples of
with those reported for other Mg-based alloys prepared using pure Mg.
alternative processing procedures. For example, there are 2. The potential for successfully using ECAP is markedly
reports of maximum elongations of 320 pct at a testing tem- increased in both the Mg-Zr alloy and pure Mg by using
perature of 773 K in a hot-rolled AZ31 (Mg-3 pct Al-1 pct the EX-ECAP process, in which the material is subjected
Zn) alloy where dynamic recrystallization occurred during to extrusion prior to processing by ECAP. When using
testing,[50] of 360 pct in a cross-rolled AZ31 alloy subjected EX-ECAP, pure Mg may be pressed through two passes
to dynamic recrystallization at a low temperature and then without breaking at temperatures of 573 and 623 K, and
tensile testing at 723 K,[51] and of 580 pct at 673 K in a hot- the Mg-Zr alloy may be pressed through four passes with-
rolled AZ61 (Mg-6 pct Al-1 pct Zn) alloy.[52] All of these out breaking at 573 K.
elongations are comparable to those obtained in the present 3. Ultrafine grain sizes cannot be achieved in pure Mg
experiments, but the testing temperatures are consistently through the use of EX-ECAP: the minimum grain size
100 °C to 200 °C higher. In the present work, the tensile elon- in these experiments was ⬃26 ␮m. By contrast, the pres-
gations of ⬎300 pct achieved at 573 K, equivalent to 0.57 Tm, ence of Zr particles inhibits grain growth in the Mg-Zr
are sufficient for use in superplastic forming operations where alloy, and it was possible to use EX-ECAP to achieve
the requisite strains are equivalent, typically, to tensile strains grain sizes of ⬃1.4 ␮m.
of the order of ⬃300 to 400 pct.[53,54] 4. Tensile testing at 523 and 573 K revealed enhanced duc-
Finally, additional clarification is required with refer- tilities in the Mg-Zr alloy, with maximum elongations up
ence to two of the experimental observations. to ⬃300 to 400 pct. The measured values for the strain-
First, the measured strain-rate sensitivities in the Mg-0.6 rate sensitivity under these conditions were ⬃0.3 to 0.4.
pct Zr alloy are ⬃0.3 to 0.4, and this appears to be more This suggests that a dislocation-glide process may be rate-
consistent with dislocation glide as the rate-controlling controlling, but probably with some contribution from
process rather than superplasticity. In dislocation glide, it grain-boundary sliding because of the very fine grain size.
is well established that the value of n ⫽ 1/m ⫽ 3.0, where
n is the stress exponent. By contrast, true superplasticity is
generally associated with a value of n ⬇ 2, equivalent to ACKNOWLEDGMENTS
m ⬇ 0.5.[55] Similar observations and conclusions were We are grateful to Dr. Koichi Makii (Kobe Steel, Ltd.,
reported recently for an aluminum-based 2024 (Al-4.4 pct Kobe, Japan) for providing the material used in this inves-
Cu-1.5 pct Mg-0.6 pct Mn) alloy, where m ⬇ 0.3 and the tigation and we thank Takayoshi Fujinami for experimental
maximum elongations were also ⬃300 to 500 pct after ECAP assistance. This work was supported in part by the Light
processing.[49] It has been shown that dislocation glide can Metals Education Foundation of Japan, in part by the
be the rate-controlling process during the high-temperature Mitsubishi Foundation, and in part by the United States
creep of dilute Mg-based alloys when the grain sizes are Army Research Office under Grant No. DAAD19-00-1-0488.
large: for example, experiments on a Mg-0.8 pct Al alloy
with a grain size of ⬃240 ␮m gave n ⬇ 3.0 at tempera-
tures in the range from 523 to 623 K.[56] However, more REFERENCES
work is now needed to clarify the precise nature of the
flow mechanism in ultrafine-grained materials exhibiting 1. F.H. Froes, D. Eliezer, and E. Aghion: JOM, 1998, vol. 50 (9),
pp. 30-34.
enhanced ductilities with m ⬇ 0.3 to 0.4. 2. K. Johnson: Adv. Mater. Proc., 2002, vol. 160 (6), pp. 62-65.
Second, it is intriguing to note that the highest elongation 3. H. Friedrich and S. Schumann: J. Mater. Proc. Technol., 2001, vol. 117,
recorded in these experiments was ⬃380 pct at 573 K for pp. 276-81.
the alloy subjected to only a single pass of ECAP, as shown 4. R.Z. Valiev, R.K. Islamgaliev, and I.V. Alexandrov: Progr. Mater.
Sci., 2000, vol. 45, pp. 103-89.
in Figure 8. At first sight, this point appears anomalous by 5. V.M. Segal, V.I. Reznikov, A.E. Drobyshevskiy, and V.I. Kopylov:
reference to the elongations achieved after processing through Russ. Metall., 1981, vol. 1, pp. 99-105.
two and four passes. However, an even higher elongation 6. V.M. Segal: Mater. Sci. Eng., 1995, vol. A197, pp. 157-64.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 35A, JUNE 2004—1743


10-03-80A-4.qxd 27/4/04 17:24 Page 1744

7. M. Furukawa, Z. Horita, M. Nemoto, and T.G. Langdon: J. Mater. 31. A. Yamashita, Z. Horita, and T.G. Langdon: Mater. Sci. Eng., 2001,
Sci., 2001, vol. 36, pp. 2835-43. vol. A300, pp. 142-47.
8. N.A. Smirnova, V.I. Levit, V.I. Pilyugin, L.S. Davydova, and V.A. 32. Z. Horita, K. Matsubara, K. Makii, and T.G. Langdon: Scripta Mater.,
Sazonova: Fiz. Metall. Metalloved., 1986, vol. 61, pp. 1170-77. 2002, vol. 47, pp. 255-60.
9. A.P. Zhilyaev, S. Lee, G.V. Nurislamova, R.Z. Valiev, and T.G. 33. K. Matusbara, Y. Miyahara, Z. Horita, and T.G. Langdon: Acta Mater.,
Langdon: Scripta Mater., 2001, vol. 44, pp. 2753-58. 2003, vol. 51, pp. 3073-84.
10. A.P. Zhilyaev, G.V. Nurislamova, B.-K. Kim, M.D. Baró, J.A. Szpunar, 34. J.A. Chapman and D.V. Wilson: J. Inst. Met., 1962–63, vol. 91, pp. 39-40.
and T.G. Langdon: Acta Mater., 2003, vol. 51, pp. 753-65. 35. K.U. Kainer: Proc. 3rd Int. Magnesium Conf., G.W. Lorimer, ed., The
11. Y. Iwahashi, Z. Horita, M. Nemoto and T.G. Langdon: Acta Mater., Institute of Materials, London, 1997, pp. 533-43.
1997, vol. 45, pp. 4733-41. 36. D.V. Wilson: J. Inst. Met., 1970, vol. 98, pp. 133-43.
12. Y. Iwahashi, Z. Horita, M. Nemoto, and T.G. Langdon: Acta Mater., 37. M. Hilpert, A. Styczynski, J. Kiese, and L. Wagner: in Magnesium
1998, vol. 46, pp. 3317-31. Alloys and Their Applications, B.L. Mordike and K.U. Kainer, eds.,
13. J. Wang, Y. Iwahashi, Z. Horita, M. Furukawa, M. Nemoto, R.Z. Wiley-VCH, Weinheim, Germany, 1998, pp. 319-24.
Valiev, and T.G. Langdon: Acta Mater., 1996, vol. 44, pp. 2973-82. 38. G.F. Weissmann and W. Babington: Proc. ASTM, 1958, vol. 58,
14. Z. Horita. D.J. Smith, M. Furukawa, M. Nemoto, R.Z. Valiev, and pp. 869-86.
T.G. Langdon: J. Mater. Res., 1996, vol. 11, pp. 1880-90. 39. A.A. Nayeb-Hashemi and J.R. Clark: Binary Alloy Phase Diagrams,
15. P.B. Berbon, N.K. Tsenev, R.Z. Valiev, M. Furukawa, Z. Horita, T. Massalski, ed., ASM, Metals Park, OH, 1986, pp. 1566-67.
M. Nemoto, and T.G. Langdon: Metall. Mater. Trans. A, 1998, 40. Y. Iwahashi, J. Wang, Z. Horita, M. Nemoto, and T.G. Langdon:
vol. 29A, pp. 2237-43. Scripta Mater., 1996, vol. 35, pp. 143-46.
16. Y. Iwahashi, Z. Horita, M. Nemoto, and T.G. Langdon: Metall. Mater. 41. C. Xu and T.G. Langdon: Scripta Mater., 2003, vol. 48, pp. 1-4.
Trans. A, 1998, vol. 29A, pp. 2503-10. 42. M. Furukawa, Y. Iwahashi, Z. Horita, M. Nemoto, and T.G. Langdon:
17. Z. Horita, T. Fujinami, and T.G. Langdon: Mater. Sci. Eng., 2001, Mater. Sci. Eng., 1998, vol. A257, pp. 328-32.
vol. A318, pp. 34-41. 43. M. Furukawa, Z. Horita, and T.G. Langdon: Mater. Sci. Eng., 2002,
18. K. Nakashima, Z. Horita, M. Nemoto, and T.G. Langdon: Mater. Sci. vol. A332, pp. 97-109.
Eng., 2000, vol. A281, pp. 82-87. 44. T.G. Langdon: Metall. Trans. A, 1982, vol. 13A, pp. 689-701.
19. Y. Nishida, H. Arime, J.-C. Kim, and T. Ando: Scripta Mater., 2001, 45. J. Wang, Z. Horita, M. Furukawa, M. Nemoto, N.K. Tsenev, R.Z.
vol. 45, pp. 261-66. Valiev, Y. Ma, and T.G. Langdon: J. Mater. Res., 1993, vol. 8,
20. A. Azushima and K. Aoki: Mater. Sci. Eng., 2002, vol. A337, pp. 45-49. pp. 2810-18.
21. Y. Saito, H. Utsunomiya, H. Suzuki, and T. Sakai: Scripta Mater., 46. Z. Horita, D.J. Smith, M. Nemoto, R.Z. Valiev, and T.G. Langdon:
2000, vol. 42, pp. 1139-44. J. Mater. Res., 1998, vol. 13, pp. 446-50.
22. J.-C. Lee, H.-K. Seok, J.-H. Han, and Y.-H. Chung: Mater. Res. Bull., 47. K. Oh-ishi, Z. Horita, D.J. Smith, and T.G. Langdon: J. Mater. Res.,
2001, vol. 36, pp. 997-1004. 2001, vol. 16, pp. 583-89.
23. J.-H. Han, H.-K. Seok, Y.-H. Chung, M.-C. Shin, and J.-C. Lee: Mater. 48. S. Komura, Z. Horita, M. Furukawa, M. Nemoto, and T.G. Langdon:
Sci. Eng., 2002, vol. A323, pp. 342-47. Metall. Mater. Trans. A, 2001, vol. 32A, pp. 707-16.
24. J.-C. Lee, H.-K. Seok, J.-Y. Suh, J.-H. Han, and Y.-H. Chung: Met- 49. S. Lee, M. Furukawa, Z. Horita, and T.G. Langdon: Mater. Sci. Eng.,
all. Mater. Trans. A, 2002, vol. 33A, pp. 665-73. 2003, vol. A342, pp. 294-301.
25. J.-C. Lee, H.-K. Seok, and J.-Y. Suh: Acta Mater., 2002, vol. 50, 50. X. Wu and Y. Liu: Scripta Mater., 2002, vol. 46, pp. 269-74.
pp. 4005-19. 51. J.C. Tan and M.J. Tan: Scripta Mater., 2002, vol. 47, pp. 101-06.
26. M. Mabuchi, H. Iwasaki, K. Yanase, and K. Higashi: Scripta Mater., 52. W.-J. Kim, S.W. Chung, C.S. Chung, and D. Kum: Acta Mater., 2001,
1997, vol. 36, pp. 681-86. vol. 49, pp. 3337-45.
27. M. Mabuchi, K. Ameyama, H. Iwasaki, and K. Higashi: Acta Mater., 53. A. Wisbey and M.W. Kearns: in Superplasticity: 60 Years after
1999, vol. 47, pp. 2047-57. Pearson, N. Ridley, ed., The Institute of Materials, London, 1995,
28. T. Mukai, M. Yamanoi, H. Watanabe, and K. Higashi: Scripta Mater., pp. 305-23.
2001, vol. 45, pp. 89-94. 54. C.F. Dressel: in Superplasticity: 60 Years after Pearson, N. Ridley,
29. H. Watanabe, T. Mukai, K. Ishikawa, and K. Higashi: Scripta Mater., ed., The Institute of Materials, London, 1995, pp. 359-76.
2002, vol. 46, pp. 851-56. 55. T.G. Langdon: Acta Metall. Mater., 1994, vol. 42, pp. 2437-43.
30. W.J. Kim, C.W. An, Y.S. Kim, and S.I. Hong: Scripta Mater., 2002, 56. S.S. Vagarali and T.G. Langdon: Acta Metall., 1982, vol. 30,
vol. 47, pp. 39-44. pp. 1157-70.

1744—VOLUME 35A, JUNE 2004 METALLURGICAL AND MATERIALS TRANSACTIONS A

You might also like