Sie sind auf Seite 1von 7

Journal of Alloys and Compounds 476 (2009) 118124

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jallcom

Aging behaviour and precipitate morphologies in Mg7.7Al0.5Zn0.3Mn


(wt.%) alloy
Wei-Jen Lai a, , Yi-Yun Li a , Yung-Fu Hsu b , Shan Trong c , Wen-Hsiung Wang a
a

Department of Materials Science and Engineering, National Taiwan University, Taipei 106, Taiwan, ROC
Department of Materials Science and Mineral Resources Engineering, National Taipei University of Technology, Taipei 106, Taiwan, ROC
c
Materials and Electro-Optics Research Division, Chung-Shan Institute of Science and Technology, Lung-Tan 325, Taiwan, ROC
b

a r t i c l e

i n f o

Article history:
Received 26 July 2008
Accepted 25 August 2008
Available online 16 October 2008
Keywords:
Metals
Mechanical alloying
Scanning and transmission electron
microscopy

a b s t r a c t
All the precipitate morphologies of Mg17 Al12 in AZ80 for a range of aging temperatures are investigated
in detail using TEM and SEM. The results show that Mg17 Al12 is the dominant precipitate in AZ80 and
can be divided into different types which can be discriminated by their morphologies. To date there have
been few papers in the literature focused on the relationship between the aging behaviour and these
morphologies. This study elaborates on the sequence of morphological evolution of Mg17 Al12 precipitates
as a function of temperature and investigates the associated age-hardening response.
2008 Elsevier B.V. All rights reserved.

1. Introduction
In recent years magnesium alloys have become one of the most
important commercial alloys. The major advantage of magnesium
alloys is their high strength to weight ratio, which makes them
popular with mobile device manufacturers, whose main goal is to
reduce the weight of their products. Conventionally, magnesium
alloys are fabricated largely by casting and in particular die-casting.
However, the major problem with this processing route is the low
yield rate. Therefore, manufacturers are striving to develop better
wrought magnesium alloys to address aspects of this problem. Thus
AZ80 has become one of the prominent newly developed wrought
magnesium alloys.
There are several kinds of magnesium alloy systems. The AZ
series is primarily based on the MgAl binary alloy system and
dominates most of the magnesium alloys. Aluminium, which can
reduce the grain size and greatly enhance the mechanical properties of magnesium alloys, is the most signicant elemental addition
to magnesium alloys. Besides, when the addition of Al exceeds the
critical limit (6 wt.%) [1], Mg17 Al12 intermetallic compounds will
precipitate and the mechanical properties of the AZ series alloy are
further enhanced.
AZ80 is based on the Mg8 wt.% Al with the addition of about
0.5 wt.% Zn and a small amount of Mn. Manganese can improve

Corresponding author. Tel.: +886 22 5915343.


E-mail address: william721225@hotmail.com (W.-J. Lai).
0925-8388/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2008.08.043

the corrosion resistance by adding MnCl2 to the melt to precipitate any dissolved iron as a complex compound, thus removing the
iron and enhancing the corrosion resistance of the resulting alloy.
Besides, manganese will also form intermetallic compounds with
aluminium and magnesium [2].
A small amount of zinc in magnesium alloys will not form any
intermetallic compound with either magnesium or aluminium,
but will reduce the solid solubility of aluminium in magnesium
and increase the amount of Mg17 Al12 precipitate. It will therefore
improve the strength of the alloy [2]. Besides, Zn is also a potent
solid solution strengthener.
This study will focus on the identication of the diverse range
of Mg17 Al12 precipitate morphologies formed under different aging
conditions and the determination of the relationship between these
morphologies and the associated aging behaviour.
2. Experimental procedure
The as-extruded AZ80 alloy was received from Chung-Shan Institute of Science
and Technology (Taiwan). An ingot of 150 mm in diameter was produced by verticaldirect-chill-casting and was extruded at 340 C. The composition was measured by
EPMA and is shown in Table 1. The as-extruded material was in the form of a slab of
100 mm in width, 6 mm in thickness and 1000 mm in length. The thickness of the
material was then further reduced to 2 mm by hot rolling at 400 C. The rolled plate
was then solution-treated at 420 C for 1 h and water quenched to room temperature.
Samples were cut from the plate and then aged at 125, 150, 175, 200, 250, and
300 C, respectively. Samples aged between 125 and 200 C were heat treated in a
silicone oil bath and for higher temperatures in an air furnace. All aging temperatures
were calibrated within 2 C. The aged samples were then water quenched to room
temperature.

W.-J. Lai et al. / Journal of Alloys and Compounds 476 (2009) 118124

119

Table 1
The composition (wt.%) of the as-received AZ80 alloy
Al
Zn
Mn
Mg

7.7
0.5
0.3
balance

Hardness was measured using a micro-Vickers machine. The microstructure was


observed after polishing and etching with a solution comprising of 3 g picric acid,
3 ml acetic acid, 5 ml water and 50 ml ethanol. Materal for TEM samples was rst
ground to about 100 m thickness and then punched to a disk of 3 mm in diameter.
The disks were subsequent twin-jet electro-polished at 15 C and 20 V with a 10
vol.% perchloric acid90 vol.% ethanol electrolyte. TEM observations were conducted
with a JEOL100CXII electron microscope operated at 100 kV.

3. Results and discussion


Fig. 2. Discontinuous lamellar structure in AZ80 aged at 200 C for 16 h.

3.1. Hardness
The hardness curves of the AZ80 aged at different temperatures for various periods of time are shown in Fig. 1. No signicant
increase in hardness was observed for AZ80 aged at 125 C up
to 128 h. The hardness curves are obviously composed of three
regions, which is similar to most of the precipitation-hardenable
alloys. First there is an incubation period, with the higher the aging
temperature the shorter the length of time. The second region is
dened by a steady increase to the peak hardness, and the value of
which also depends on the aging temperature. After the peak hardness, the hardness became steady and does not fall abruptly, which
is contrast to the behaviour of precipitation-hardenable aluminium
alloys.
At aging temperatures above 150 C, the alloy started to show
signicant increase in hardness. As can be seen from Fig. 1, when
aging at 150 C, the hardness began to increase signicantly after
32 h and did not reach its peak hardness during the time period
examined, 256 h. When aging at 175 C, a peak hardness of HV82
at about 240 h was observed. When aging at 200 C, the hardness
plateaued at about HV76 after 64 h. The peak hardness drops to
HV70 at 32 h for aging at 250 C. Aging at 300 C has no signicant
hardening effect.
The difference in peak hardness observed at different aging
temperatures is in part associated with the difference in volume
fraction of the precipitates formed. But the other important reason

Fig. 1. The aging curve of AZ80 magnesium alloy.

for this difference is the morphology of the corresponding precipitates.


3.2. Morphologies of Mg17 Al12 precipitates in AZ80 alloy
The Mg17 Al12 phase has a BCC structure with a lattice parameter of 1.056 nm [3]. The Mg17 Al12 precipitates can be divided into
two distinct categories according to how they are formed, being
either discontinuous or continuous. Both of these categories precipitate from the magnesium matrix directly to the -phase (Mg17 Al12 )
without any intermediate phases or GP zones [4]. The discontinuous precipitates can subsequently be divided into three different
groups according to their growth morphology. One is a cellular
structure which consists of lamellae. The lamellar structure always
initiates at grain boundaries and grows perpendicular to the boundary, as shown in Fig. 2. The growth will stop only if the grain
is full of the precipitates or the continuous precipitates start to
increase signicantly and impede the growth of the discontinuous precipitates. According to the literature, the growth region of
the cellular structure can be treated as a part of the grain in the
opposite side of the grain it grows into [5]. The cellular structure
region and the grain which it grows from have the same matrix
structures.
Another kind of discontinuous precipitate is an oval-shaped
ellipsoid, hereafter referred to as the elliptical structure. The elliptical structure and the cellular structure usually formed at the same

Fig. 3. Discontinuous elliptical structure in AZ80 aged at 150 C for 64 h.

120

W.-J. Lai et al. / Journal of Alloys and Compounds 476 (2009) 118124

Fig. 4. Discontinuous intergranular structure in AZ80 aged at 300 C for 1 h.

time, as shown in Fig. 3. The size of the elliptical structure is about


0.11.5 m and rarely exceeds 1.5 m.
These two kinds of precipitate constitute about 90% of the
discontinuous precipitates observed. Once these precipitates are
formed, they do not change in their morphology signicantly. They
only grow outward to consume regions which are not occupied by
other precipitates.
The last minor kind of discontinuous precipitate is the intergranular precipitate, as shown in Fig. 4. This kind of precipitate
is detrimental to mechanical properties of AZ80 and causes the
material to fracture intergranularly.
The continuous precipitates occur in two distinct morphologies.
One is the Widmansttten structure (Fig. 5), which has some different precipitation orientations with the matrix [3,611], and the
other being an irregular slab structure (Fig. 6), which is highly oriented to one direction within a grain. Sometimes the two structures
will also mix with each other, as can be seen in Fig. 7.
Table 2 contains data on the size of precipitates as a function of
time at different aging temperatures. Because the observed precipitate size varies widely, a range of size is given instead of an average
value. The range is calculated by counting at least 20 precipitates
and eliminating the extreme cases. The thickness of the lamellae in
the lamellar structures showed no signicant change with different aging times or temperatures and therefore is omitted from the
table. From Table 2 it can be seen that continuous and discontinuous
precipitates appear at almost the same time.
Discontinuous precipitates stop growing relatively earlier than
continuous ones. But continuous precipitate will still keep grow-

Fig. 5. Dark eld image of continuous Widmansttten structure in AZ80 aged at


200 C for 8 h.

Fig. 6. Continuous irregular slab structure in AZ80 aged at 250 C for 32 h.

ing and occupy the rest of the unoccupied areas. The size of these
two precipitates will not change signicantly during the overaging
process.
3.3. Variation of discontinuous precipitate morphology
The lamellar and elliptical structures appear mostly under
200 C, with the amount of these two kinds of discontinuous precipitate decreasing as the aging temperature increases. It is hard
to see them above 250 C. Table 3 shows the different precipitate morphologies associated with different aging temperatures. In
addition, different aging temperatures will result in different elliptical structures, not in the shape of the precipitate, but in the size
and number density. Comparing the 150 C-256 h, 200 C-64 h (T6)
and 250 C-32 h (T6) samples, it was observed that the elliptical
precipitate density at 150 C was higher than that at both 200 and
250 C, as indicated in Fig. 8(a), (b) and (c) respectively. The lower
number density translates into a larger interspacing between the
precipitates. The larger the interspacing will result in the lower the
ability of the precipitate to resist dislocation movement, and the
hardness will then decrease [12]. In the 150 C-256 h sample the
elliptical structure is rather smaller in size and has a larger number
density, which corresponds to a higher peak hardness. The result
is from lots of observations of the SEM and TEM samples, not just
from particular sites.

Fig. 7. Irregular slab and Widmansttten structure intermix with each other. Specimen aged at 250 C for 32 h.

W.-J. Lai et al. / Journal of Alloys and Compounds 476 (2009) 118124

121

Table 2
Dimensions (m) of the precipitates in AZ80 aged at various temperatures ( C) and times (h)
Aging condition

Temperature ( C)

Continuous precipitate
Time (h)

Widmansttten

Discontinuous precipitate
Irregular slab

Length

Width

Length

Elliptical
Width

Intergranular

Diameter

Diameter

150

64
256

0.20.4
0.51

0.030.05
0.10.15

0.10.3
0.10.3

0.10.5
0.30.6

200

4
8
16
32
64
256

0.20.5
0.20.7
0.41.0
0.71.5
1.01.7
1.52.0

0.020.05
0.050.07
0.050.1
0.070.1
0.150.3
0.30.4

0.20.4
0.20.5
0.20.5
0.20.5
0.20.5
0.20.7

0.10.3
0.10.3
0.20.4
0.50.7
0.60.9
1.01.4

250

2
32

0.61.0
1.53.0

0.10.15
0.6

0.51.5
0.51.5

0.30.5
1.02.0

1
16

0.20.7
1.52.5

0.050.1
0.5

300

From Table 2 it can be seen that the size ranges for the various
precipitate morphologies observed in the T6 condition for the different aging temperatures. The size may contribute a little benet
to the mechanical properties because smaller and more numerous
precipitation particles will have better abilities to resist dislocation slip. Despite a large amount of discontinuous precipitation
in the material, the hardening effect is relatively poor because of
the large precipitate size compared with precipitation-hardenable
aluminium alloys.
The last kind of discontinuous precipitation is the intergranular precipitate. It appears for a wide temperature range and forms
simultaneously with the other two kinds of discontinuous precipi-

0.52.5

0.30.6

0.30.5

tates. The size of the intergranular precipitate changes widely with


both aging time and temperature as indicated in Table 2. This kind of
precipitate prefers to nucleate at grain boundary triple points, but
will also nucleate discretely along the grain boundaries, as shown
in Fig. 4. This type of precipitate grows into large and irregular
shape which is undesired and therefore should be avoided as far
as possible. When the aging temperature is raised up to 300 C, the
intergranular precipitates become the dominant precipitate and
grow extensively. The discrete precipitation particles along grain
boundaries also grow, coalesce and form a massive bulk structure
which is also detrimental to the material, as shown in Fig. 9. This
may be the primary factor that AZ80 at this aging temperature

Fig. 8. The morphologies of elliptical precipitates: (a) 150 C-256 h, (b) 200 C-64 h (T6), and (c) 250 C-32 h (T6).

122

W.-J. Lai et al. / Journal of Alloys and Compounds 476 (2009) 118124
Table 3
The temperature ranges ( C) of various precipitates
Widmanstatten
Slab-shaped
Lamellar
Elliptical
Intergranular

Fig. 9. Coalescent intergranular precipitates of specimen aged at 300 C for 16 h.

yields no increase in hardness after a signicant aging time. The


intergranular precipitates also consume a signicant proportion of
the solute content and this results in a depletion of other precipitation like continuous precipitates, as shown in Fig. 9.
3.4. Variation of continuous precipitate morphology
Continuous precipitation precipitates within the whole material
uniformly but does not form on the grain boundaries. Its growth is
a continuous process which the composition of the matrix is continuously changing with the growth of the continuous precipitates.

150300
250
150250
150250
150300

The Widmansttten structure is the most dominant continuous precipitate morphology in the aging temperatures between
150 and 250 C. From the work of Celotto [3], it is stated that precipitates of the Widmansttten structure are the major hardening
factor in AZ91 and different morphologies of the Widmansttten
structure were also observed. The size changes of the precipitates
with aging temperatures are listed in Table 2, but the distribution densities of the precipitate are different as well. From the
observations of 150 C-256 h, 200 C-64 h (T6), and 250 C-32 h (T6)
samples (Fig. 10(a)(c)), it was found that the distribution density of
precipitates in the 150 C-256 h sample was much higher and yields
higher hardness than for the other two conditions. Lower aging
temperatures also yield smaller precipitates, which are benecial
to the mechanical properties.
The morphologies of Widmansttten structures which change
with aging time were also investigated in Celottos work [3]. Taking 200 C aging for example, initially the precipitate nucleate as a
lozenge shape about 0.20.5 m in length (Fig. 5), and then grows
along a particular direction while keeping the other dimensions
unchanged. The precipitate is now called an asymmetric lozenge
shape (Fig. 10(b)). The precipitate keeps growing to its maximum
length, which corresponds to the peak hardness. After the peak
hardness, the precipitate grows in thickness. The thickness at this

Fig. 10. Widmansttten structures of (a) 150 C-256 h, (b) 200 C-64 h (T6), and (c) 250 C-32 h (T6) specimens.

W.-J. Lai et al. / Journal of Alloys and Compounds 476 (2009) 118124

Fig. 11. Specimen aged at 200 C for 256 h shows an asymmetric hexagon structure.

123

stage is about 0.30.4 m while the length varies from 1.52.0 m.


After sufcient time of aging, the sharp corners of the Widmansttten structures will become blunt and are shaped like asymmetric
hexagons (Fig. 11).
According to the literature [611], the predominant orientation relationship between Widmansttten structure and the
magnesium matrix is (0 0 0 1)m //(1 1 0)p and [1 2 10]//[1 1 1]. Some
precipitates have precipitation orientations perpendicular to the
basal plane. The orientation relationship of the precipitate is
p , which is reported by
(0 0 0 1)m //(1 1 1)p and [1 2 10]m //[1 1 2]
Crawley and Lagowski [6] and Crawley and Milliken [7]. And
some precipitates have its growing direction lie an angle to the
basal plane. The orientation relationship of the precipitate is
(1 2 11)m //(1 1 0)p and [10 1 0]m //[1 1 0]p [8].
Fig. 12(a) and (b) is the bright eld image and corresponding diffraction pattern from the Widmansttten structure aged at
250 C for 32 h. The diffraction pattern is from the precipitate indicated in Fig. 12(a). The spots from matrix and precipitate are easily
distinguished. The zone axis of the matrix diffraction pattern is

Fig. 12. TEM image and the corresponding diffraction pattern of Widmansttten structure of the alloy aged at 250 C for 32 h. (a) Bright eld image, (b) corresponding
diffraction pattern of the Widmansttten precipitate lying in (0 0 0 1) zone axis, (c) the reconstructed pattern, (d) the computed symmetric pattern and (e) the computed six
variants of the precipitate.

124

W.-J. Lai et al. / Journal of Alloys and Compounds 476 (2009) 118124

[0 0 0 1]m and the zone axis of the precipitate diffraction pattern


is [1 1 0]p . Because of the two-fold symmetry of the (1 1 0)p precipitate diffraction pattern, the above diffraction pattern has two
symmetric forms. Fig. 12(c) is the diffraction patterns reconstructed
from Fig. 12(b), and Fig. 12(d) is the symmetric pattern computed
from Fig. 12(c). Open circles indicate the matrix reections and
black spots indicate the precipitate reections. From Fig. 12(c) it
p and [10 1 0]m is
can be seen that [1 2 10]m is parallel to [1 1 1]
p . On the other hand, the symmetric pattern in
parallel to [1 1 2]
Fig. 12(d) has [1 2 10]m parallel to [1 1 1]p and [10 1 0]m parallel to
p . Fig. 12(e) shows the relative position of the precipitate in
[1 1 2]
Fig. 12(a). The white one in the middle represents the precipitate
indicated by an arrow in Fig. 12(a). The other white ones represent
the precipitates resulted from the six-fold symmetry of the (0 0 0 1)
matrix diffraction pattern. The black ones represent the precipitates resulted from the two-fold symmetry of (1 1 0)p stated above.
Therefore, the Widmansttten structure has totally six variants [3]
as indicated in Fig. 12(e).
All the precipitates in Fig. 12(a) have their shapes and directions
exactly the same as in Fig. 12(e). This further conrms the computed
result of the precipitate orientation. Most of the Widmansttten
structures have a length 48 times greater than the width. Some
are extremely long, and some have almost equal sides.
In magnesium alloys, most of the slips occur mainly on the basal
planes (0 0 0 1) at room temperature. Only at high temperature the
high index slip systems are initiated.
Since the most Widmansttten structures precipitate parallel
to the basal planes, it is hard for precipitation plates to resist the
dislocation movement. Only the precipitates lying perpendicular to
the basal plane and having an angle to the basal plane have better
ability to impede dislocation slip.
The irregular slab structures form above 200 C and appear
extensively around 250 C, as indicated in Table 2. These precipitates have a length that varies widely, from 0.5 to 2.5 m in length,
0.3 to 0.6 m in width, as shown in Fig. 6. The special feature of
this precipitate is that it orients along a specic direction which
remains undetermined. The precipitates are usually accompanied
by Widmansttten structures.
The irregular slab structures are usually relatively large, so the
mechanical properties are not good. From the aging curve of 250 C
in Fig. 1, it was found that the hardness of the material has only 10
HV higher than that of the solution-treated state. At this aging temperature the dominant precipitates are Widmansttten and slab
structures. When irregular slab structure starts to form, the Widmansttten structure will coarsen signicantly, and resulted in poor
mechanical properties.
4. Conclusions
1. The hardness increment of AZ80 is about 38% for the T6 condition
at 175 C. This hardening ability is extremely poor compared to

2.

3.

4.

5.

6.

precipitation-hardenable aluminium alloys, which is inuenced


by the morphology, the size, and the distribution density of the
Mg17 Al12 precipitates.
All the morphologies of Mg17 Al12 precipitate in AZ80 are discussed. There are three kinds of discontinuous precipitate
including a lamellar structure, an elliptical structure, and an
intergranular precipitate. In addition, there are two kinds of continuous precipitate, a Widmansttten structure and an irregular
slab structure.
The size and distribution density of precipitates depend on
the aging temperature. Different aging temperatures also yield
different precipitate morphologies. Basically, the Widmansttten structure appears at all aging temperatures; the lamellar
and elliptical structures appear under 250 C; the irregular slab
structure appears around 250 C and the intergranular precipitates appear at all temperatures but grows extensively around
300 C.
Both continuous and discontinuous precipitates have effect on
the hardness of the alloy. At lower aging temperatures, the elliptical and Widmansttten structures have a higher number density
and smaller size, which will signicantly increase the hardness
of the alloy.
The Widmansttten structure has three different orientations
with respect to basal plane. Most of them are lying in the basal
plane.
At around 250 C both the irregular slab and Widmansttten
morphologies coarsen extensively. So the hardness is poor compared with other aging temperatures.

Acknowledgement
The authors gratefully acknowledge the nancial support for
this research by Chung-Shan Institute of Science and Technology
(Taiwan) under grant no. BV96E06P.
References
[1] C.H. Caceres, C.J. Davidson, J.R. Grifths, C.L. Newton, Mater. Sci. Eng. A 325
(2002) 344355.
[2] C.R. Brooks, Heat Treatment, Structure and Properties of Nonferrous Alloys,
ASM, Metals Park, OH, 1984.
[3] S. Celotto, Acta mater. 48 (2000) 17751787.
[4] S. Celotto, T.J. Bastow, Acta Mater. 49 (2001) 4151.
[5] D.A. Porter, K.E. Easterling, Phase Transformations in Metals and Alloys, 2nd ed.,
CRC Press, 2004, pp. 322326.
[6] A.F. Crawley, B Lagowski, Metall. Trans. 5 (1974) 949.
[7] A.F. Crawley, K.S. Milliken, Acta Metall. 22 (1974) 557.
[8] D. Duly, W.-Z. Zhang, M. Audier, Phil. Mag. A 71 (1) (1995) 187.
[9] D. Duly, Y. Brechet, Acta Metall. Mater. 42 (9) (1994) 3036.
[10] D. Duly, M.C. Cheynet, Y. Brechet, Acta Metall. Mater. 42 (11) (1994) 3843.
[11] D. Duly, J.P. Simon, Y. Brechet, Acta Metall. Mater. 43 (1) (1995) 101.
[12] R.E. Reed-Hill, R. Abbaschian, Physical Metallurgy Principles, 3rd ed., PWS Publishing Company, 1994, pp. 532534.

Das könnte Ihnen auch gefallen