You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257707482

Achieving High Strength and High Ductility in Friction Stir-Processed Cast


Magnesium Alloy

Article in Metallurgical and Materials Transactions A · August 2013


DOI: 10.1007/s11661-013-1744-5

CITATIONS READS

43 506

3 authors:

Wei Yuan Sushanta Panigrahi


Hitachi America Ltd. Indian Institute of Technology Madras
41 PUBLICATIONS 1,742 CITATIONS 102 PUBLICATIONS 3,986 CITATIONS

SEE PROFILE SEE PROFILE

Rajiv S. Mishra
University of North Texas
524 PUBLICATIONS 30,889 CITATIONS

SEE PROFILE

All content following this page was uploaded by Sushanta Panigrahi on 09 May 2015.

The user has requested enhancement of the downloaded file.


Achieving High Strength and High Ductility in Friction
Stir-Processed Cast Magnesium Alloy
WEI YUAN, SUSHANTA K. PANIGRAHI, and RAJIV S. MISHRA

Friction stir processing (FSP) is emerging as an effective tool for microstructural modification
and property enhancement. As-cast AZ91 magnesium alloy was friction stir processed with
one-pass and two-pass to examine the influence of processing conditions on microstructural
evolution and corresponding mechanical properties. Grain refinement accompanied with
development of strong basal texture was observed for both processing conditions. Ultrafine-
grained (UFG) AZ91 was achieved under two-pass FSP with fine precipitates distributed on the
grain boundary. The processed UFG AZ91 exhibited a high tensile strength of ~435 MPa
(117 pct improvement) and tensile fracture elongation of ~23 pct. The promising combination of
strength and ductility is attributed to the elimination of casting porosity, and high density of fine
precipitates in an UFG structure with quite low dislocation density. The effects of grain size,
precipitate, and texture on deformation behavior have been discussed.

DOI: 10.1007/s11661-013-1744-5
 The Minerals, Metals & Materials Society and ASM International 2013

I. INTRODUCTION reduction in ductility, especially when the grain size is


refined to an ultrafine regime, mainly because of the
MAGNESIUM alloys are attracting great interest reduced work hardening.[15–17]
from transportation industries for structural applica- Texture, which is always observed in magnesium
tions because of their high strength-to-weight ratio.[1,2] alloys during intense shear deformation, plays an
However, low ductility and room temperature formabil- important role on room-temperature deformation
ity limit the processing and application.[3] As a result, behavior in addition to grain size as a result of limited
the main fabrication route of magnesium alloys remains independent slip systems available. Texture modification
casting.[1,2] Casting defects such as microscopic shrink- of wrought AZ magnesium alloys by SPD has shown
age porosity and inclusions adversely affect the strength promising room-temperature ductility because of the
and ductility of magnesium alloy. It is expected that the formation of a strong basal texture, which orientated the
application will be expanded if the strength and ductility easy basal slip plane to the preferred orientation;
of magnesium alloys can be enhanced. however, the tensile yield strength was very low.[18–25]
Grain refinement has been a good option for enhanc- The special basal texture generation during SPD also led
ing the strength of magnesium alloys. Several severe to the anisotropy in mechanical properties, especially in
plastic deformation (SPD) tools, such as equal channel yield strength, which is directly related to the activation
angular extrusion/pressing,[4,5] high-pressure torsion,[6,7] of deformation slip or twinning systems.[14,20,26]
accumulated roll bonding,[8] alternate biaxial reverse It is desirable to enhance strength as well as ductility
corrugated pressing,[9] differential speed rolling,[10,11] in magnesium alloys. Recently, the FSP of cast alloys
and friction stir processing (FSP)[12–14] have demon- has shown effective elimination of cast porosity, exten-
strated capability in refining grain structure of magne- sive grain refinement, dissolution, and breakup of coarse
sium alloys. However, grain refinement-induced second phase.[12,27–32] Precipitation has been shown to
strength enhancement is generally accompanied by the be an effective additive strengthening mechanism for
magnesium alloys in addition to the grain refinement-
related strengthening. Feng and Ma[28] reported a high
tensile strength of 337 MPa in AZ91 magnesium alloy
WEI YUAN, Researcher, is with the Department of Materials
Science & Engineering, Missouri University of Science and Technol-
via FSP with average grain size refined to ~15 lm and
ogy, Rolla, MO 65409, and also with the Research & Development fine precipitates decorating the grain boundaries. By
Division, Hitachi America Ltd., 34500 Grand River Avenue, Farm- introducing nanosized precipitates in a fine-grained rare
ington Hills, MI 48335. SUSHANTA K. PANIGRAHI, Assistant earth magnesium alloy EV31A through FSP, Freeney
Professor, is with the Department of Materials Science & Engineering, and Mishra[30] obtained significant enhancement in
Missouri University of Science and Technology, and also with the
Indian Institute of Technology Madras, Chennai 600036, India. tensile strength and ductility. Xiao et al.[32] presented
RAJIV S. MISHRA, Professor, is with the Department of Materials similar findings on a FSP fine-grained Mg-Gd-Y-Zr
Science & Engineering, Missouri University of Science and Technol- alloy with even higher tensile strength of 439 MPa, but a
ogy, and also with the Center for Friction Stir Processing, Department noticeable loss of elongation to 3.4 pct. Our previous
of Materials Science and Engineering, University of North Texas,
Denton, TX 76203. Contact e-mail: Rajiv.Mishra@unt.edu
study has achieved highly textured ultrafine-grained
Manuscript submitted September 9, 2012. (UFG) AZ31 using FSP, which exhibited high tensile
Article published online April 20, 2013 strength and ductility because of restricted basal slip.[26]

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, AUGUST 2013—3675


The current study aims to (a) achieve high strength and grain structure and morphology and distribution of
high ductility in commercial grade cast magnesium alloy precipitates under an operating voltage of 200 kV.
through microstructural modification, (b) verify whether Tensile property characterization in the current study
strong texture and strength anisotropy exist in FSP of was conducted using a computer-controlled, screw-driven
high aluminum content precipitation strengthened mag- tensile testing machine at a constant crosshead velocity.
nesium alloy, and (c) explore the influence of grain size, Tensile specimens have a gage length of 1.3 mm, a width
precipitates, and texture on deformation behavior of of 1.0 mm (to eliminate the effect of texture variation)
processed magnesium alloy. and a thickness of 0.6 mm. Specimens were machined
along the PD and transverse direction (TD) from the
center of the processed region and were polished with
II. EXPERIMENTAL PROCEDURE 1-lm finish and tested at room temperature at an initial
strain rate of 1 9 103 s1. Aging treatment at 441 K
A high-pressure die-casting AZ91 magnesium alloy (168 C) for 16 h was applied to two-pass AZ91 to
having a thickness of 4.2 mm was used. The nominal characterize the effect of post aging on tensile property.
composition of this alloy was 9 wt pct aluminum and At least three specimens for each condition were tested to
1 wt pct zinc. To make friction stir passes, a tool with evaluate the average property values.
12-mm diameter concave shoulder and 2.1-mm-stepped
spiral pin was employed. Two types of FSP runs were
made on the as-cast AZ91 with one-pass and two-pass III. RESULTS
runs. A tool rotation rate of 600 revolutions per minute
(rpm) and a tool traverse speed of 1.7 mm/s were applied Figure 1 shows optical and SEM micrographs of
to the one-pass run. Two-pass FSP was carried out with AZ91 alloy in the as-cast condition. The low magnifi-
300 rpm and 1.7 mm/s overlapping the first pass. A tool cation of optical micrograph exhibits typical casting
tilt of 2.5 deg was adopted for all the FSP passes. features with dendritic pattern of a-Mg matrix in white
Microstructures of the friction stir-processed samples, and second phase in gray.[29] The casting porosities in
with cross section being perpendicular to the processing black were visible in certain locations with an area
direction (PD), were examined using optical microscopy fraction of ~4.2 pct after averaging six randomly
(OM), scanning electron microscope (SEM), and elec- selected locations with same magnification using ImageJ
tron back-scattered diffraction (EBSD). For OM and software. The as-cast AZ91 varies in grain size with
SEM observations, samples were mechanically polished small grains of about 10 lm and large dendrites of
to 1 lm with alcohol-based diamond polishing com- about 120 lm. The secondary electron imaging at low
pounds and etched using an acetic picral solution (4.5 g magnification shows high density of coarse second-
picric acid, 10 ml acetic acid, 10 ml water, and 70 ml phase network distributed at grain boundaries. This
ethanol). The specimens for EBSD characterization second phase was identified as b-Mg17Al12 phase
were further fine polished with 0.02 lm colloidal silica through energy dispersive X-ray (EDX) analysis. One
solution. The EBSD characterization was performed at such b-Mg17Al12 phase is shown in Figure 2 with a size
the center of the processed AZ91 in a cross section of about 10 lm. The fraction of b-Mg17Al12 phase in as-
perpendicular to the PD. The linear intercept method cast AZ91 was ~18.6 pct.
was used to obtain an average grain size. Transmission The friction stir-processed AZ91 was sectioned in the
electron microscope (TEM) samples were prepared by TD for evaluation of the microstructure evolution.
focused ion milling in Helio Nano Lab 600 FIB/ Figure 3 presents the typical cross sections of friction
FESEM. A TECNAI F20 TEM was used to examine stir-processed AZ91 in one-pass and two-pass conditions.

Fig. 1—Optical micrograph and SEM image of as-cast AZ91.

3676—VOLUME 44A, AUGUST 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


In the current study, different material flow behavior microscope (STEM) image at the same location. The
can be observed for material processed with one-pass two-pass FSP caused a significant grain refinement
and two-pass. The material processed with two-pass compared with the one-pass run with much finer and
shows a uniform macrostructure. However, a nonuni- more equiaxed grain structure. As shown in Figure 6(a),
form structure can be observed for one-pass run, which the average grain size of two-pass friction stir-processed
separates the processed region into top and bottom AZ91 was 0.54 ± 0.20 lm computed based on STEM
parts. images at six different locations. The grain boundaries
Figure 4 presents the microstructures of one-pass under bright field imaging are not distinct, which might
friction stir-processed AZ91 at top and bottom regions. be related to the low misorienation angle nature of
Figure 4(a) is a low-magnification view of the top region boundaries. The round-edged b-Mg17Al12 phase parti-
and shows layer structure with alternate white and black cles were mainly observed on the grain boundaries and
contrast. A high-magnification examination indicates grain triple junctions. Few round shaped finer precip-
b-phase particle-rich layers, with some horizontally itates were noticed in the grain interiors. The relatively
elongated, incompletely dissolved coarse b-phase being large precipitates have ellipse shape; however, the fine
distributed alternately with a-Mg matrix. The cast precipitates retain a round shape. The size of precipi-
dendritic structure was completely eliminated and tates was computed in terms of equivalent circular
substituted by much finer equiaxed grain structure. diameter (square root of the product of lengths in major
Furthermore, examination of the processed region and minor axes of the ellipse) assuming ellipse shape of
showed complete elimination of cast porosity. A com- precipitates. The size distribution of precipitates in two-
parison between the top and bottom regions pass FSP is shown in Figure 6(b) with a phase fraction
(Figures 4(b) and (c)) indicates similar grain size with of 12.8 pct. The average size of precipitates was
an average grain size of about 2.8 ± 1.2 lm but distinct 113 ± 98 nm. Few dislocations can be detected in the
precipitate distribution. As shown in Figure 4(d), the processed region. Table I lists the EDX analysis of
initial coarse network type of Mg17Al12 phase were contents at various locations, the grain interior con-
dissolved or broken into much smaller and almost tained ~3.8 wt pct Al, and Zn was mainly concentrated
equiaxed ones at the bottom region with majority of on precipitates or adjacent locations.
them distributed at the grain boundaries. Figure 7 shows the {0002} and {10-10} pole figures of
Grain structure and precipitate morphology and friction stir-processed AZ91 in as-cast and friction stir-
distribution of two-pass friction stir-processed AZ91 processed conditions. The location for friction stir-
were examined using TEM. Figures 5(a) and (b) show a processed specimens was at the centerline of the pro-
bright-field image and scanning transmission electron cessed region at a depth of 1.0 mm. Pole figures of
{0002} and {10-10} of as-cast AZ91 indicate a relatively
random texture with maximum intensity of about 3.
This random texture of cast AZ91 was modified to a
strong basal fiber texture with alignment of the c-axes of
the grains tilting about 25 to 30 deg away from the ND
after FSP. The difference in degree of tilt between one-
pass and two-pass was negligible. The observation of
texture evolution in AZ91 is similar to that has been
reported in FSP of rolled AZ31 with a completely
different starting textures.[26]
Figure 8 presents the representative tensile engineer-
ing stress–strain curves of AZ91 in as-cast and friction
stir-processed conditions tested at room temperature
with an initial strain rate of 1 9 103 s1. At least three
(six for as-cast AZ91) specimens for each condition were
tested. The statistical value of mechanical properties,
including tensile yield strength (YS), ultimate tensile
strength (UTS), uniform elongation and total elonga-
Fig. 2—Presence of b-Mg17Al12 phase and a-magnesium matrix in tion to failure, are summarized in Table II. The as-
the as-cast AZ91. received AZ91 shows low tensile strength and poor

Fig. 3—Cross-sectional macrographs of friction stir-processed AZ91 with (a) single pass at 600 rpm/1.7 mm/s and (b) two-pass with 600 rpm/
1.7 mm/s for the first pass and 300 rpm/1.7 mm/s for the second pass.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, AUGUST 2013—3677


room-temperature ductility of less than 3 pct total processed AZ91 tested in TD, after one-pass, the tensile
elongation to failure for average. Majority of the as-cast YS and total elongation increased from ~160 MPa to
AZ91 specimens failed abruptly. For friction stir- 302 MPa and ~3 pct to 19 pct, respectively. The tensile

Fig. 4—Micrographs show the grain structure and precipitate in one-pass friction stir-processed AZ91, (a, b) top region, and (c, d) bottom region.

Fig. 5—Grain structure and precipitates of two-pass friction stir-processed AZ91 in (a) TEM bright field image and (b) STEM image, location
‘A’ was marked in both images.

3678—VOLUME 44A, AUGUST 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


Fig. 6—Frequency distribution plots of (a) grain size and (b) precipitate size in two-pass friction stir-processed AZ91.

Table I. Results of EDX Analysis of Fig. 5 (Weight Percent)

Location Al Zn Mg
A 3.8 — 96.2
B 39.6 3.5 56.9
C 3.6 0.2 96.2
D 19.4 3.3 77.3

Fig. 8—Stress–strain curves of as-received and friction stir-processed


AZ91.

strength was further enhanced after two-pass FSP, with


average tensile YS of ~390 MPa and UTS of ~435 MPa.
The total elongation to failure for two-pass friction stir-
processed AZ91 was also improved to ~23 pct. Com-
parison of one-pass specimen with two-pass specimen
shows a decrease in uniform elongation from 13.6 pct to
8.8 pct, but an increase in total elongation as grain size
becomes finer. This observation is consistent with
previous findings on friction stir-processed AZ31.[23]
A significant anisotropy was also observed for friction
stir-processed AZ91 in PD and TD. For processed AZ91
tested in PD, all stress–strain curves exhibit a common
feature similar to that of the as-cast AZ91 specimens
that fractured without developing a macroscopic neck-
ing, i.e., there is no pronounced post-uniform elonga-
tion. However, unlike as-cast AZ91, the tensile strength
and total elongation were improved significantly after
FSP with more than 10 pct total elongation. Contrary to
Fig. 7—{0002} and {10-10} pole figures of AZ91 in (a) as-cast, the previous results on fine-grained and UFG AZ31,[23]
(b) one-pass FSP and (c) two-pass FSP. the ductility in PD for AZ91 increased as the grain grew

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, AUGUST 2013—3679


Table II. A Summary of Tensile Properties in Different Conditions

Condition YS (MPa) UTS (MPa) Uniform Elongation (pct) Total Elongation (pct)
As-Cast 162 ± 12 201 ± 15 1.9 ± 1.4 2.7 ± 1.1
One-pass_TD 302 ± 3 379 ± 4 13.6 ± 0.8 19.1 ± 1.4
One-pass_PD 203 ± 10 285 ± 8 10.5 ± 1.3 10.5 ± 1.3
Two-pass_TD 390 ± 5 435 ± 1 8.8 ± 0.3 23.4 ± 0.2
Two-pass_PD 258 ± 6 351 ± 6 15.4 ± 0.4 15.6 ± 0.5
Two-pass+ aging_TD 381 ± 4 422 ± 7 7.5 ± 0.4 21.2 ± 0.2
Two-pass + aging_PD 234 ± 8 329 ± 6 8.1 ± 0.5 8.1 ± 0.5

finer. The discrepancy is likely to be due to the better between the as-cast and the recrystallized microstruc-
microstructural homogenization in two-pass AZ91 com- tures, the transported cast material from the front of the
pared with the one-pass. The observation that TD has pin tool might not undergo full dissolution of precipi-
higher tensile strength as well as higher ductility than tates and exhibited the difference between top and
those of PD for the current fine-grained and UFG AZ91 bottom regions. For the two-pass FSP, the starting
is similar to the findings of friction stir-processed microstructure for the second pass is recrystallized grain
AZ31.[23] Aging of two-pass FSP did not enhance the structure with remaining b-phase particles, which elim-
strength in either PD or TD, on the contrary, reduced inated the variation in structure in the processed region
the ductility, especially in PD. of the two-pass run. Pre-solution treatment of cast
magnesium alloy to homogenize the microstructure has
been proven to result in uniform microstructure in single
pass FSP.[29,31]
IV. DISCUSSION As shown in Figures 4 and 5, FSP eliminated cast
porosity, network type of coarse b-phase, and signifi-
A. Microstructural and Texture Evolution
cantly refined the grain structure as a result of dynamic
The as-cast structure was modified by FSP with the recrystallization.[12,27,36–39] Average grain sizes of
complete elimination of coarse dendritic grains and ~2.8 and 0.5 lm were achieved in one-pass and two-
casting porosity. Accompanying the grain refinement, pass runs, respectively. UFG structure has been
the coarse network type of precipitates was dissolved reported by FSP of various metals and alloys, which
and refined into much finer scale. However, distinct generally requires low frictional heat input, high strain
difference in processed region was observed between (rate) and extra cooling to recrystallize the grain
one-pass and two-pass FSP of as-cast AZ91, with structure and suppress grain growth during post-
heterogeneous layer structure for one-pass and more FSP.[13,40,41] The very low dislocation density in
uniform structure for two-pass. Similar observation of two-pass friction stir-processed AZ91 might indicate
b-phase particles rich bands has been reported during discontinuous dynamic recrystallization which sweeps
the FSPs of cast AZ91 and AZ80 magnesium alloys, off dislocation during grain boundary migration, result-
which was attributed to the lower strain rate/strain and ing in very low dislocation density in the recrystallized
temperature conditions.[27,31] A high tool rotation rate grains. The current UFG structure via two-pass FSP
was suggested to dissolve b-phase and in turn achieve was mainly attributed to low heat input and homoge-
uniform microstructure.[27] Processing parameters have neous distribution of fine b-phase precipitates, which pin
shown direct influence on the thickness of heterogeneous grain boundary migration and inhibit grain growth.
layer during the FSP of cast AZ91. Previous observation The as-cast AZ91 had a random texture as expected
indicated that the top layer was thinner as the tool (Figure 7(a)). However, a strong basal texture devel-
rotation rate increased for a constant tool traverse speed oped during one-pass and two-pass FSPs. Similar to the
(images not shown). previous results in FSP of AZ31, a basal fiber texture for
Another factor causing the difference in the macro- both one-pass and two-pass runs shows a tilt of c-axis
structures between one-pass and two-pass FSPs was the toward the PD by about 25 through 30 deg at the
characteristic of starting microstructure for processing. centerline of the processed region.[23,42] The formation
During the FSP, material in front of the tool is driven of basal texture has been well documented as a result of
upward. The upward-moving material on the leading shear deformation during FSP, which aligns the easy
side then sweeps around the pin in the rotation basal slip planes roughly parallel to pin column sur-
direction;[33–35] however, on the trailing side, the layer face.[24–26,43] It has been suggested that rare earth
rotating with the tool is decelerated and is deposited in elements can weaken the deformation texture[44,45]
the void in the wake of the pin with the assistance of tool through particle-stimulated nucleation of recrystalliza-
shoulder.[34] The heterogeneous layer structure is likely tion.[46] However, from the current results, it can be
related to the sweeping action of the shoulder as it concluded that the b-phase particles have limited effect
comes into contact with the as-cast material first and on texture weakening with the evidence of much similar
transports it toward the trailing edge around the strong basal texture between AZ91 and AZ31 both
rotating pin. Because of the significant difference obtained with FSP.

3680—VOLUME 44A, AUGUST 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


B. Strengthening in Friction Stir-Processed AZ91 recrystallization. The Hall–Petch parameters (friction
Magnesium Alloy stress r0 and stress concentration factor ky) for grain-
The strength of high aluminum content magnesium boundary strengthening of friction stir-processed AZ31
alloy is influenced by the effect of alloy content on solid with grain size between 0.7 and 10.2 lm is presented in
solution strengthening, the size and volume fraction of Table III.
b-phase particles, as well as the grain size and existing The theoretically computed tensile YS values without
microstructural defects. The as-cast AZ91 exhibited considering particle strengthening for two-pass friction
quite low tensile strength, poor elongation and large stir-processed UFG AZ91 were 163 MPa and 367 MPa
variation in these properties. These are attributed to the in PD and TD, respectively. The strengthening contri-
presence of cast porosities and network type coarse bution from interaction of b-phase particles with basal
Mg17Al12 phase at the grain boundaries that tend to dislocation through Orowan looping was developed by
nucleate cracks or undergo debonding from the matrix Nie[48] and can be expressed as:
 
during deformation.[28] Gb d
FSP resulted in a significant improvement in both DsOrowan ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  ln p ; ½1
b
tensile strength and ductility. This is mainly because of 2p ð1  mÞ 0:953 p ffi  1 dp
f
porosity elimination, grain refinement, and significant
breakup and dissolution of the coarse Mg17Al12 where G is the shear modulus (17 GPa for magnesium),
phase.[28,30,31] Different from those reported by Feng b is the Burgers vector (3.2 9 1010 m), v is the
et al.[31] on AZ80 that one-pass FSP led to even less Poisson’s ratio (0.35),[49] f is the phase fraction (0.128)
ductility of ~3 pct (in PD) than that of as-cast, the of b-phase, dp is the mean planar diameter of the point
current one-pass FSP enhanced not only the tensile obstacles (1.13 9 107 m). The computed contribution
strengths but also the ductility, although heterogeneous from particle strengthening to resolved shear stress is
microstructure was observed in both cases. The discrep- 34 MPa. A Taylor factor of 2.3 for basal slip dominated
ancy results from different thermal cycles which resulted deformation in PD was achieved from EBSD data by
in different grain sizes and fractions of remaining assuming Sachs polycrystal deformation model (inter-
undissolved precipitates. The finer grain size and lower granular stress equilibrium). Therefore, the increase in
amount of remaining coarse precipitates observed in the tensile YS due to b-phase particles was ~78 MPa. The
one-pass specimens of the current study contributed to overall computed tensile YS in PD was 241 MPa,
the better tensile properties. The two-pass FSP in PD slightly less than the experimentally determined value
further increased the tensile strength and ductility of 258 MPa. Assuming similar particle–dislocation
compared with the one-pass FSP (as shown in Figure 8 interaction applies to nonbasal dislocation dominated
and Table II) because of the further refinement of deformation in TD, by simply replacing the Taylor
grains, precipitates, and elimination of heterogeneous factor to 2.2 for nonbasal slip according to EBSD data
microstructure as evidenced in Figures 4 and 5. analysis, an increase in tensile YS due to b-phase
The TD testing exhibited much higher tensile strength particles was found to be ~75 MPa. This overestimates
than that in PD for both one-pass and two-pass FSPs, the tensile YS to 442 MPa. These results indirectly
with extremely high tensile YS of ~390 MPa and UTS of indicate that b-phase particles in two-pass friction stir-
~435 MPa in two-pass FSP. The strength anisotropy processed UFG AZ91 have more effect on inhibiting
was attributed to the strong basal texture generated basal dislocation slip than nonbasal dislocation slip, at
during the FSP, which tilted the c-axis toward PD, and least in the current study with this distribution of
in turn promoted easy basal slip in PD and difficult precipitates. The slight variation in tensile YS and
nonbasal slip in TD.[26] High-strength AZ91 has also significant reduction in ductility in PD, but hardly any
been reported during high ratio differential speed change in TD after aging might support this hypothesis.
rolling, which appears difficult to interpret from The negligible effect of aging on property in current
strengthening mechanisms point of view because of the UFG AZ91 compared with the significant strengthening
extremely fine grains of 280 nm and distribution of fine reported in fine-grained AZ91 by Feng et al.[28] was due
b-phase particles (88 nm) on the grain boundaries.[47] to the difference in Al contents in matrix, about 3.8 pct
For the currently obtained high-strength AZ91, EDX for current study and 7.5 pct in Reference 28. It may be
analysis (Table I) of grain interior of two-pass AZ91 speculated that grain boundary strengthening, solid
showed low aluminum supersaturating with a content of solution strengthening (initially embedded in Hall–Petch
~3.8 wt pct, which was slightly higher than the content formula for friction stir-processed AZ31) and particle
in AZ31 magnesium alloy. Considering the much similar strengthening contributed to the overall strength at
c-axis tilts in friction stir-processed AZ91 and AZ31,
and the fact that the variation in basal c-axis tilt was
about 5 to 10 deg, it might be rational to compute the Table III. Hall–Petch Parameters Developed for Friction
tensile YS of the current friction stir-processed AZ91 by Stir-Processed AZ31[42]
adopting the empirical grain boundary strengthening
formula previously developed for friction stir-processed Loading r0 (MPa) ky (MPa lm1/2)
AZ31 with additional strengthening from b-phase par- PD 39 88
ticles,[42] since the dislocation density was low for both TD 145 157
friction stir-processed AZ31 and AZ91 during dynamic

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, AUGUST 2013—3681


yielding in PD, and mainly grain boundary strengthen- stir-processed AZ91 in both PD and TD, as well as
ing and solid solution strengthening in TD. friction stir-processed UFG AZ31(having ~2 pct uni-
form elongation in TD and ~10 pct uniform elongation
in PD).[23] After a short period of elastoplastic transi-
C. Effects of Grain Size and Precipitate on Work
tion, the work hardening rate decreases linearly and
Hardening and Ductility
steadily for all testing conditions because of the balance
The ductilities were significantly improved after FSPs between dislocation storage and annihilation, which is
in both PD and TD compared with the as-cast condition the stage III hardening range.[53] Stage III can be
as shown in Figure 8 and Table II. The lower ductility expressed using Voce equation, which is given as
in the as-cast AZ91 was due to the presence of cast
porosity and network type of coarse b-phase, which dr @Gbk1 k2 r
could be the stress concentration sites during loading: as ¼  ½2
de 2 2
a consequence, crack nucleation, void coalescence, or
debonding between a-Mg matrix and b-phase led to where k1 is a constant and k2 is twice of the slope of the
premature fracture. FSP eliminated casting defects and strain hardening rate versus true stress plot and is an
coarse b-phase, and thus enhanced the tensile ductility. indication of the rate of recovery. The first term on the
The increase in ductility from one-pass to two-pass was right-hand side is associated with the dislocation storage
related to the particle refinement and microstructure discussed above and the second term is associated with
homogenization. Grain refinement to ultrafine scale annihilation of dislocations. The stage III is different for
generally leads to reduction in ductility as a result of one-pass and two-pass FSP in PD and TD, evidenced
reduced work hardening in finer grain size. It is also from the various slopes of stage III. The two-pass FSP
reported that tensile ductility could be enhanced in the in TD shows the largest slope, followed by one-pass in
UFG materials with lower dislocation density, which PD and one-pass in TD. The two-pass FSP in PD
can provide more room for dislocation storage.[50] The exhibits lowest slope, i.e., the lowest rate of dislocation
current two-pass UFG AZ91 shows extremely high recovery. As shown in Figure 9(b), the degree of work
tensile strength as well as high ductility, with a relatively hardening was smallest for two-pass in TD, and largest
low uniform elongation but high post-uniform elonga- for one-pass in TD and two-pass in PD, with one-pass in
tion in TD. The high post-uniform elongation was likely PD in between. The largest degree of work hardening
related to the enhanced strain rate sensitivity as grain and the lowest rate of dislocation recovery in PD for
size becomes finer.[51,52] two-pass AZ91 contribute to the highest uniform
Although the uniform elongation in two-pass friction elongation. On the contrary, the lowest in TD for two-
stir-processed UFG AZ91 in TD was relatively low pass AZ91.
compared with the PD or one-pass FSP, it is much Comparing the UFG AZ91 with UFG AZ31, we find
higher than that generally reported in UFG magnesium that both have submicron grain sizes (0.5 lm for AZ91
alloy,[15,23] especially with such a high tensile strength. and 0.8 lm for AZ31), similar inclined c-axis tilts of
This observation suggests that plastic instability was about 25 through 30 deg, and low dislocation densities,
delayed in current UFG AZ91 through either enhanced but completely different work hardening behaviors. The
work hardening ability or reduced rate of dislocation initial work hardening rate was much higher for UFG
recovery. Figure 9 plots the work hardening rate h AZ31, but it quickly dropped below that of UFG AZ91
(=dr/de) derived from true stress verses true strain as a in a short net flow stress region. The recovery rate
function of true stress/net flow stress for friction determined by linear fit in stage III was much higher for

Fig. 9—Work hardening rate as a function of true stress and net flow stress for friction stir-processed AZ91 in PD and TD, work hardening
behavior of UFG AZ31 was also included.[23]

3682—VOLUME 44A, AUGUST 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A


UFG AZ31 than that for UFG AZ91 in both PD and 12. R.S. Mishra and Z.Y. Ma: Mater. Sci. Eng. R, 2005, vol. 50, pp. 1–78.
TD. It is quite possible that the fine b-phase precipitates 13. C.I. Chang, X.H. Du, and J.C. Huang: Scripta Mater., 2007,
vol. 57, pp. 209–212.
enhanced effective dislocation storage, which retarded 14. G. Bhargava, W. Yuan, S.S. Webb, and R.S. Mishra: Metall.
the drop of work hardening rate and resulted in the Mater. Trans. A, 2010, vol. 41A, pp. 13–17.
improved uniform elongation in AZ91, although the 15. S.X. Ding, W.T. Lee, C.P. Chang, L.W. Chang, and P.W. Kao:
grain size was smaller than that of AZ31. Scripta Mater., 2008, vol. 59, pp. 1006–09.
16. W.J. Kim and Y.G. Lee: Mater. Sci. Eng. A, 2011, vol. 528,
pp. 2062–66.
17. W.J. Kim, Y.G. Lee, M.J. Lee, J.Y. Wang, and Y.B. Park: Scripta
V. CONCLUSIONS Mater., 2011, vol. 65, pp. 1105–08.
18. T. Mukai, M. Yamanoi, H. Watanabe, and K. Higashi: Scripta
The as-cast AZ91 magnesium alloy was friction stir Mater., 2001, vol. 45, pp. 89–94.
19. J. Koike, T. Kobayashi, T. Mukai, H. Watanabe, M. Suzuki, K.
processed with one-pass and two-pass runs. FSP elim- Maruyama, and K. Higashi: Acta Mater., 2003, vol. 51, pp. 2055–65.
inated porosity and significantly refined the average 20. S.R. Agnew, J.A. Horton, T.M. Lillo, and D.W. Brown: Scripta
grain sizes to 2.8 and 0.5 lm for one-pass and two-pass Mater., 2004, vol. 50, pp. 377–81.
runs, respectively, but also introduced strong basal fiber 21. W.J. Kim, C.W. An, Y.S. Kim, and S.I. Hong: Scripta Mater.,
2002, vol. 47, pp. 39–44.
texture. The network type of coarse precipitates was 22. H.W. Lee, T.S. Lui, and L.H. Chen: J. Alloys Compd., 2009,
dissolved and broken into fine particles mainly distrib- vol. 475, pp. 139–44.
uted at the grain boundaries. As the grain size reduced 23. W. Yuan and R.S. Mishra: Mater. Sci. Eng. A, 2012, vol. 558,
during second FSP pass, the precipitates also became pp. 716–24.
finer. FSP remarkably enhanced the tensile property of 24. Y.N. Wang, C.I. Chang, C.J. Lee, H.K. Lin, and J.C. Huang:
Scripta Mater., 2006, vol. 55, pp. 637–40.
the as-cast AZ91. Two-pass FSP improved the tensile 25. W. Woo, H. Choo, D.W. Brown, P.K. Liaw, and Z. Feng: Scripta
YS to ~390 MPa, UTS to ~435 MPa (117 pct improve- Mater., 2006, vol. 54, pp. 1859–64.
ment) in TD; and YS to ~258 MPa, and UTS to 26. W. Yuan, R.S. Mishra, B. Carlson, R.K. Mishra, R. Verma, and
~340 MPa (62 pct improvement) in PD. The tensile R. Kubic: Scripta Mater., 2011, vol. 64, pp. 580–83.
27. Z.Y. Ma, A.L. Pilchak, M.C. Juhas, and J.C. Williams: Scripta
failure elongations were also enhanced from 3.5 pct to Mater., 2008, vol. 58, pp. 361–66.
23 pct and to 16 pct in TD and PD, respectively. The 28. A.H. Feng and Z.Y. Ma: Scripta Mater., 2007, vol. 56, pp. 397–
promising combination of strength and ductility was 400.
attributed to the elimination of porosity in UFG 29. P. Cavaliere and P.P. De Marco: J. Mater. Process. Technol., 2007,
structure with low dislocation density and fine precip- vol. 184, pp. 77–83.
30. T.A. Freeney and R.S. Mishra: Metall. Mater. Trans. A, 2009,
itates. Precipitates appear to influence basal dislocations vol. 41A, pp. 73–84.
more than nonbasal dislocations in current UFG AZ91. 31. A.H. Feng, B.L. Xiao, Z.Y. Ma, and R.S. Chen: Metall. Mater.
Trans. A, 2009, vol. 40A, pp. 2447–56.
32. B.L. Xiao, Q. Yang, J. Yang, W.G. Wang, G.M. Xie, and Z.Y.
Ma: J. Alloys Compd., 2011, vol. 509, pp. 2879–84.
33. H.N.B. Schmidt, T.L. Dickerson, and J.H. Hattel: Acta Mater.,
ACKNOWLEDGMENTS 2006, vol. 54, pp. 1199–1209.
34. K. Colligan: Welding J., 1999, vol. 78, pp. 229–37s.
The authors gratefully acknowledge the support 35. T.U. Seidel and A.P. Reynolds: Metall. Mater. Trans. A, 2001,
vol. 32A, pp. 2879–84.
provided by the National Science Foundation through 36. J.A. Esparza, W.C. Davis, E.A. Trillo, and L.E. Murr: J. Mater.
grants NSF-EEC-0531019 and NSF-IIP-1157754. Sci. Lett., 2002, vol. 21, pp. 917–20.
37. S.H.C. Park, Y.S. Sato, and H. Kokawa: J. Mater. Sci., 2003,
vol. 38, pp. 4379–83.
38. W.B. Lee, Y.M. Yeon, and S.B. Jung: J. Mater. Sci. Technol.,
REFERENCES 2003, vol. 19, pp. 785–90.
39. U.F.H.R. Suhuddin, S. Mironov, Y.S. Sato, H. Kokawa, and
1. B.L. Mordike and T. Ebert: Mater. Sci. Eng. A, 2001, vol. 302, C.W. Lee: Acta Mater., 2009, vol. 57, pp. 5406–18.
pp. 37–45. 40. C.I. Chang, X.H. Du, and J.C. Huang: Scripta Mater., 2008,
2. B.B. Clow: Adv. Mater. Process., 1996, vol. 150, p. 33. vol. 59, pp. 356–59.
3. S.R. Agnew, J.W. Senn, and J.A. Horton: JOM, 2006, vol. 58, 41. J.Q. Su, T.W. Nelson, and C.J. Sterling: Scripta Mater., 2005,
pp. 62–69. vol. 52, pp. 135–40.
4. M. Mabuchi, H. Iwasaki, K. Yanase, and K. Higashi: Scripta 42. W. Yuan, S.K. Panigrahi, J.Q. Su, and R.S. Mishra: Scripta
Mater., 1997, vol. 36, pp. 681–86. Mater., 2011, vol. 65, pp. 994–97.
5. A. Yamashita, Z. Horita, and T.G. Langdon: Mater. Sci. Eng. A, 43. S.H.C. Park, Y.S. Sato and H. Kokawa: Metall. Mater. Trans. A,
2001, vol. 300, pp. 142–47. 2003, vol. 34A, pp. 987–94.
6. A. Galiyev and R. Kaibyshev: Mater. Trans., 2001, vol. 42, 44. M.R. Barnett, M.D. Nave, and C.J. Bettles: Mater. Sci. Eng. A,
pp. 1190–99. 2004, vol. 386, pp. 205–11.
7. M. Kai, Z. Horita, and T.G. Langdon: Mater. Sci. Eng. A, 2008, 45. J. Bohlen, M.R. Nürnberg, J.W. Senn, D. Letzig, and S.R. Agnew:
vol. 488, pp. 117–24. Acta Mater., 2007, vol. 55, pp. 2101–12.
8. M.T. Pérez-Prado, J.A. Del Valle, and O.A. Ruano: Scripta 46. E.A. Ball and P.B. Prangnell: Scripta Metall., 1994, vol. 31,
Mater., 2004, vol. 51, pp. 1093–97. pp. 111–16.
9. Q. Yang and A.K. Ghosh: Acta Mater., 2006, vol. 54, pp. 5147–58. 47. W.J. Kim, H.G. Jeong, and H.T. Jeong: Scripta Mater., 2009,
10. H. Watanabe, T. Mukai, and K. Ishikawa: J. Mater. Sci., 2004, vol. 61, pp. 1040–43.
vol. 39, pp. 1477–80. 48. J.F. Nie: Scripta Mater., 2003, vol. 48, pp. 1009–15.
11. W.J. Kim, J.D. Park, J.Y. Wang, and W.S. Yoon: Scripta Mater., 49. H.J. Frost and M.F. Ashby: Deformation-Mechanism Maps,
2007, vol. 57, pp. 755–58. Pergamon Press, Oxford, 1982, p. 44.

METALLURGICAL AND MATERIALS TRANSACTIONS A VOLUME 44A, AUGUST 2013—3683


50. Y.H. Zhao, J.F. Bingert, Y.T. Zhu, X.Z. Liao, R.Z. Valiev, Z. 52. W.J. Kim, H.W. Lee, S.J. Yoo, and Y.B. Park: Mater. Sci. Eng. A,
Horita, T.G. Langdon, Y.Z. Zhou, and E.J. Lavernia: Appl. Phys. 2011, vol. 528, pp. 874–79.
Lett., 2008, vol. 92, p. 081903. 53. U.F. Kocks and H. Mecking: Prog. Mater Sci., 2003, vol. 48,
51. Q. Yang and A.K. Ghosh: Acta Mater., 2006, vol. 54, pp. 5159–70. pp. 171–273.

3684—VOLUME 44A, AUGUST 2013 METALLURGICAL AND MATERIALS TRANSACTIONS A

View publication stats

You might also like