Vous êtes sur la page 1sur 6

Journal of Alloys and Compounds 487 (2009) 263268

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Spherical shape of -Mg17 Al12 precipitates in AZ91 magnesium alloy processed by


equal-channel angular pressing

K.N. Braszczynska-Malik
Czestochowa University of Technology, Institute of Materials Engineering, Al. Armii Krajowej 19, 42-200 Czestochowa, Poland

a r t i c l e

i n f o

Article history:
Received 5 May 2009
Received in revised form 15 July 2009
Accepted 17 July 2009
Available online 25 July 2009
Keywords:
Magnesium alloy
Precipitation
Microstructure
Equal-channel angular pressing

a b s t r a c t
The microstructure investigations of the AZ91 alloy after equal-channel angular pressing (ECAP) were
presented. Solution annealed billets were processed at a 0.16 mm s1 pressing rate and temperatures of
623 and 553 K. The average grain size was reduced from 150 m initially to a nal value of 10 m. The
microstructure analyses revealed -Mg17 Al12 precipitates with a spherical morphology located particularly inside the equiaxal grains of an -solid solution. Additionally, different shapes of precipitates,
plate-like or rod-like (typical for heat treatment processes) were not observed. The spherical shape of
precipitates was obtained due to correlate magnesium matrix alloy deformation during pressing with
precipitation process.
2009 Elsevier B.V. All rights reserved.

1. Introduction
The AZ91 (MgAlZn) alloy is the most widely used magnesium
alloy exhibiting a good combination of high strength at room temperature, good castability and excellent corrosion resistance [14].
The microstructure of as-cast AZ91 alloy is generally characterized
by a solid solution of aluminium in magnesium (an -Mg phase
with a hexagonal closely-packed, hcp structure) and an + eutectic. In MgAl type alloys, the phase (called also phase [59])
is an intermetallic compound with a stoichiometric composition
of Mg17 Al12 (at 43.95 wt.% Al) and an -Mn-type cubic unit cell.
The zinc that is present in the AZ91 alloy does not create new
phases but substitutes aluminium in the -Mg17 Al12 phase, forming
a ternary intermetallic compound Mg17 Al11.5 Zn0.5 or Mg17 (Al,Zn)12
type [9,10].
Magnesiumaluminium alloys are susceptible to heat treatment
due to the variable solubility of aluminium in a solid state from
12.9 wt.% Al at an eutectic temperature of 710 K to about 2.9 wt.%
Al at 473 K. During conventional heat treatment of the AZ91 alloy,
involving solution annealing at about 690 K for a minimum of 24 h
(Fig. 1a), followed by ageing at about 430 K for 16 h (T6 conditions),
the discontinuous precipitates of the phase appear [4,10]. On the
other hand, during the ageing of a supersaturated solid solution at
a higher temperature, continuous precipitates of the phase can
also occur simultaneously [4,10]. Fig. 1b and c presents a discontinuous and continuous precipitates of the phase after ageing

Tel.: +48 34 3250 721; fax: +48 34 3250 721.


E-mail address: kacha@wip.pcz.pl.
0925-8388/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2009.07.100

a supersaturated AZ91 alloy at temperatures of 423 and 623 K,


respectively. At intermediate ageing temperatures both discontinuous and continuous precipitates can occur. It should also be noted
that both discontinuous and continuous precipitates in MgAl type
alloys have plate-like morphology with an accurate orientation
relationship (OR) with a matrix phase. For both precipitates, the
predominant orientation relationship is the Burgers OR, namely:
(0 0 0 1) || (0 1 0) and [2 1 1 0] || [1 1 1] [7,1114]. Additionally,
other ORs were also reported in MgAl-based alloys, i.e. the Crawley OR [7,13,14], the Porter OR [7], the Gjmmes-strmoe OR [13,14]
or the Potter OR [14,15].
The AZ91 alloy (and other MgAl type alloys) is characterized
by a large grain size after heat treatment due to the long time
required to obtain structure changes (slow volume diffusion of aluminium in magnesium) [4]. On the other hand, grain renement
is very important in magnesium alloys because they have poor
formability and limited ductility at room temperature rooted in
their hexagonal closely-packed crystal structure. In the past decade,
efforts have been concentrated on thermomechanical processing
for grain size renement using methods of severe plastic deformation (SPD) [16]. SPD techniques, such as equal-channel angular
pressing (ECAP) [16,17], accumulative roll bonding (ARB) [18,19]
or high-pressure torsion (HCP) [20,21], have been applied to the
grain renement of magnesium alloys on bulk materials. Recent
results have shown that ECAP processing can be used to achieve an
ultrane grain size in Mg-based alloys [2230]. The microstructure
and properties of materials pressed by ECAP are strongly dependent on the plastic deformation behaviour during pressing which
is governed mainly by a die geometry and process variables like
temperature, pressing speed and rotation of the billet around its

264

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 487 (2009) 263268

ECAP deformation carried out at 498 (four passes) and 453 K (two
passes) at a pressing speed of 25.2 mm/min and via route BC (before
ECAP samples were hot-rolled). Mthis et al. [32] observed rod-like
precipitates in AZ91 alloy pressed four times at 543 K using route
C with a pressing rate of 5 mm/min. They also concluded that long
rod-like precipitates were broken into smaller parts during ECAP
carried out in up to eight passes. Miyahara et al. [26] presented
small phase particles lying preferentially on the grain boundary
in the AZ61 alloy after pressing in four passes at 473 and 523 K via
route BC .
In the present work, ECAP processing was applied to obtain
a new shape of phase precipitates in AZ91 alloy thanks to
correlate plastic deformation behaviour during pressing with a precipitation process. Investigations of the individual microstructure
development were conducted using light, scanning and transmission electron microscopy.
2. Experimental procedures

Fig. 1. Microstructure of AZ91 alloy after: (a) solution annealing at 693 K for 24 h,
supersaturated solid solution; light microscopy, (b) ageing at 423 K for 16 h, discontinuous precipitates of phase, (c) ageing at 623 K for 8 h, continuous precipitates
of phase; SEM.

longitudinal axis between adjacent passes. In practice, there are


four separate processing routes in ECAP: route A in which the sample is not rotated between passes, route BA in which the sample
is rotated through 90 in alternate directions between each pass,
route BC in which the sample is rotated by 90 in the same direction (clockwise) between each pass and route C where the sample
is rotated by 180 between passes [16,17]. There are several reports
describing the microstructure and properties of MgAl type alloys
processed by ECAP [2230] but these investigations were not concerned with precipitation manner. Precipitates of the phase have
been reported sporadically in the microstructure of the AZ91 alloy
processed by ECAP. Chen et al. [31] observed formation of phase
precipitates at grain boundaries of the AZ91 alloy during two-step

The commercial as-cast AZ91 magnesium alloy with a nominal chemical composition of 9 wt.% Al, 1 wt.% Zn, 0.5 wt.% Mn was used in this study. The investigated
AZ91 alloy was cut into rods of 50 mm in length and 11.8 mm in diameter. Before
the ECAP process, the samples were heat treated in order to obtain a homogeneous
microstructure. Solution annealing was carried out at 693 K for 26 h in a protective
argon atmosphere followed by water quenching (at approximately 287 K) for all the
samples. The microstructure obtained after this heat treatment was characterized
by large grains of supersaturated solid solution with an average grain size of about
150 m (Fig. 1a).
The ECAP die used in this investigation was designed to obtain a maximum shear
strain of about 1.15 during each pass. It contained an inner contact angle equal
to 90 and corner angle  of 0 . The billets were processed at a pressing rate of
0.16 mm s1 using a plunger attached to a hydraulic press on an Instron machine. All
the pressings were conducted using route BC . This procedure was adopted because it
is evident from many experiments [16,17,2330] that the BC route leads most effectively to an array of equiaxed, ne grains separated by high-angle grain boundaries.
For each separate pressing, the samples were coated with molybdenum disulphide
(MoS2 ) as a lubricant. The temperature of the process was controlled using a thermocouple in a die. The process was conducted at 623 and 553 K, i.e. below the solvus
curve (in the range of ageing temperatures).
The samples were subjected up to four pressings and then sectioned for
microstructure examination. All microstructure analyses were carried out from both
transverse and longitudinal sections of the as-pressed billets using light and scanning electron microscopy (SEM) techniques. A standard metallographic technique
was applied for sample preparation including wet prepolishing and polishing with
different diamond pastes without contact with water. To reveal the microstructure, samples were etched in a 1% solution of HNO3 in C2 H5 OH for about 60 s. An
XL30ESEM FEG (Philips, Eindhoven, The Netherlands) scanning electron microscope
was used. The microstructure of the deformed samples was also studied using transmission electron microscopy (TEM) on a Philips CM20 instrument equipped with
an energy dispersive X-ray (EDX) spectrometer operating at 200 kV (Philips, Eindhoven, The Netherlands). For TEM analyses, the thin foils for transmission electron
microscopy study were polished electrolytically. The slices of material from both
the transverse and longitudinal sections of the as-pressed billets were mechanically thinned down, followed by punching of 3-mm diameter discs. Finally, the discs
were thinned to perforation using a Fischione double-jet electropolisher (Fischione,
Export, PA) with a solution of 33% nitric acid and 67% methyl alcohol at a temperature
of 245 K and a voltage of 20 V. The brighteld technique was used for microstructure
observations and selected area electron diffraction (SAED) patterns were employed
to dene structural constituents.

3. Results
Fig. 2 shows a representative microstructure of the AZ91 alloy
observed after four passes of the ECAP process conducted at 623 K.
The initial average grain size for the annealed condition was about
150 m as shown in Fig. 1a. The grain structure shown in Fig. 2
was recorded after four passes via route BC using the die with
contact angle equal to 90 (which allowed attainment of a maximum shear strain of about 1.15 during each pass). The average grain
size was reduced to a nal value of about 10 m. This result conrms that ECAP is a powerful technique to obtain a fast decrease of
grain size in the AZ91 magnesium alloy. Additionally, it can also be
seen that the microstructure was characterized by the presence of

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 487 (2009) 263268

265

Fig. 2. Microstructure of AZ91 alloy after four ECAP passes at 623 K (a) transverse
section (b) longitudinal section; light microscopy.

equiaxal grains in both the transverse and the longitudinal sections


of the pressed billets.
The microstructure observation of the AZ91 alloy processed by
ECAP at 623 and 553 K also revealed the presence of ne, spherical precipitates located especially inside the grains. Fig. 3 shows
light microscopy images of the investigated alloy pressed at 553 K.
Fine precipitates with a new spherical shape were observed in both
the transverse and the longitudinal sections of the pressed billets.
The same result was obtained after the ECAP process conducted at
623 K. Fig. 4 presents SEM images with visible spherical precipitates of the AZ91 alloy after ECAP processing at 623 K. It should also
be noted, that different types of precipitates, i.e. with plate-like
or rod-like morphology were not observed. In the microstructure of the AZ91 alloy after the conducted ECAP process, typical
(for heat treatment) continuous or discontinuous precipitates of
the phase (Fig. 1b and c) with characteristic morphology and
marked anisotropy of growth were not present. Identication of
the observed precipitates with a spherical shape was carried out by
means of the transmission electron microscopy technique.
Figs. 57 show typical TEM images of the analysed material.
Precipitates with spherical morphology are clearly visible in both
perpendicular sections of billets pressed at 553 and 623 K. Energy
dispersive X-ray spectrometry analyses obtained from the precipitate marked as X in Fig. 6a revealed a high concentration of
aluminium, i.e. 37 wt.% Al and raised content of zinc, i.e. 1.9 wt.%
Zn. Although the Mg:Al ratio recorded exceeds the expected stoichiometry of Mg17 (Al,Zn)12 phase, probably because of inuences
from the matrix, presented result indicates that the observed precipitates are the phase and it also provides direct evidence of
zinc presence in the phase. Spherical precipitates were identied as the phase from the analysis of the selected area diffraction

Fig. 3. Microstructure of AZ91 alloy after two ECAP passes at 553 K (a and b) transverse section (c) longitudinal section; light microscopy.

pattern. The associated SAED patterns obtained from the spherical


precipitates, marked as Y and Z in Fig. 7a and b, respectively, conrm
the identication of the phase. Additionally, the conducted investigations did not disclose any difference in structure parameters
(like unit cell parameters) between the newly obtained spherical
precipitates and typical plate-like precipitates of the phase analysed earlier [33]. It should also be noted that the maximum size of
the obtained spherical precipitates was about 1 m.
4. Discussion
The presented results indicated that the microstructure of the
AZ91 alloy processed by ECAP at 553 and 623 K via the BC pressing route at a rate of 0.16 mm s1 consisted of matrix phase
equiaxal grains of about 10 m in size (Figs. 2 and 3) and spherical
precipitates located particularly inside the grains (Figs. 3 and 6).

266

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 487 (2009) 263268

Fig. 6. TEM image of AZ91 alloy after two ECAP passes at 623 K; longitudinal section
(a) and result of EDX analysis obtained from precipitate marked as X (b).
Fig. 4. Microstructure of AZ91 alloy after four ECAP passes at 623 K (a) longitudinal
section (b) transverse section; SEM.

The observed precipitates were identied as the phase. The


microstructure observations carried out on both the transverse and
the longitudinal sections of the pressed samples provide direct evidence of the spherical shape of precipitates with a maximum size
of about 1 m. It should also be noted that in deformed samples,
different shapes of precipitates (plate-like or rod-like) were not
observed. It is well known that equilibrium morphology is a shape
that minimizes the total energy which is composed of two parts:
elastic strain energy and interfacial energy.

Fig. 5. TEM image of AZ91 alloy after two ECAP passes at 623 K; longitudinal section.

In the AZ91 alloy, discontinuous or continuous precipitates of


the phase growing during the ageing of a supersaturated solid
solution have characteristic plate-like morphology with marked
anisotropy of growth, presented in Fig. 1a and b, respectively. Precipitates with the Burgers OR are parallel to the basal plane of
the matrix, i.e. (0 0 01) whereas precipitates with optional ORs
lie on the prism plane of the magnesium, i.e. (11 00) and they
are perpendicular to the basal plane of the matrix. The analyses of
ORs [5,1015,33] suggest the presence of coherent boundaries with
slight lattice deformation between the magnesium matrix and
precipitates that explain both plate-like morphology of precipitates
and its visible anisotropy of growth.
On the other hand, the analysed planes (basal and prism planes)
in hcp magnesium are simultaneously the main slip planes. For
magnesiumwith a c/a ratio equal to 1.624 the main slip system
is {0 0 0 1} 1 1 2 0 [34,35]. In case of unfavorable orientation of
the main slip system to external stress
 or in
 higher temperatures,


different slip systems, e.g. {1 1 0 0} 1 1 2 0 and {1 1 0 1} 1 1 2 0
can also operate. In works concerning hot working of magnesium
alloys [36,37] however, an a cross-slip and energetically favorable junction between glissile a and sensile c dislocations on a
{1 1 0 0} prism plane seem to be predominant. The dissolution of
an a dislocation on the basal plane to Shockley partial dislocations
connected with a single stacking fault is also possible. However,
plastic deformation during SPD processes especially of magnesium alloys at high temperatures appears to be more composite
and complicated. For example, in the present case, the supersaturated solid solution exhibited higher than equilibrium solute atoms
concentration, which could form atmospheres generating dislocation locking. It should also be noted that the deformation of hcp
magnesium alloys caused the formation of twins, especially the
{1 1 0 2}matrix {0 1 1 2}twin type [38,39].

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 487 (2009) 263268

267

also, spherical precipitates were observed especially inside matrix


grains which is seen in Fig. 3.
In the present case, a new spherical shape of precipitates
was obtained in the AZ91 alloy processed by ECAP at a maximum
shear strain of about 1.15 during each pass and at a pressing rate
of 0.16 mm/s. The applied ECAP parameters allowed the attainment
of severe plastic deformation introducing strong disordering of the
magnesium lattice. On the other hand, the temperature and time
were suitable for the growth of precipitates during deformation.
Thus, the selected process parameters allowed the attainment of
new spherical precipitates of the phase. The obtained results are
different of those described earlier [31,32,26], probably due to different process parameters, like for example pressing speed. On the
other hand, very important factor could be a time of samples heating between passes. For example, prolonged holding of billets in a
die between consecutive passes may causes static recrystallization
processes of matrix and different conditions inside grains during
nucleation and growth of precipitates, which may result in different
form of precipitates.
5. Conclusions
1. Equal-channel angular pressing is a powerful technique for the
processing of magnesium alloys with a ne grain structure. The
homogeneous microstructure of the AZ91 alloy with an average
10 m equiaxial grains was obtained after four passes via route
BC .
2. The spherical shape of precipitates was obtained due to correlate precipitation and magnesium matrix deformation processes.
The high density of crystal defects (dislocations especially) introduced during severe plastic deformation precluded growth of
plate-like precipitates with a coherent boundary with the matrix
and enforced a new energetically favorable spherical shape.
References

Fig. 7. TEM images of AZ91 alloy after two (a) and four (b) ECAP passes at 623 K; longitudinal sections (SAED patterns obtained from the spherical precipitates marked
as Y and Z).

In most published works [2325,30,39] the microstructure of


magnesium alloys processed by the ECAP technique was characterized by strong deformation and high density of dislocations and
twins. Introducing a high dislocation density and strong disordering
of the matrix lattice during severe plastic deformation can preclude
formation of a coherent boundary between the matrix and growing precipitates. If precipitate growth proceeds during continuous
plastic deformation, then the plate-like shape of precipitates is
unfavorable. In this case the spherical shape of precipitates can
be the most energetically favorable. On the other hand, the high
density of crystal defects introduced inside phase grains during
plastic deformation act as privileged sites for the heterogeneous
nucleation of precipitates. Additionally, high dislocation density
(and other defects like vacancies) caused faster diffusion of solute
atoms in the magnesium matrix which were effective in reducing
the time necessary for the nucleation and growth of precipitates (in
comparison to precipitation during heat treatment). For this reason

[1] H. Friedrich, S. Schumann, J. Mater. Process. Technol. 117 (2001) 76.


[2] B.L. Mordike, T. Ebert, Mater. Sci. Eng. A 302 (2001) 37.
[3] B. Smola, I. Stulikova, F. von Buch, B.L. Mordike, Mater. Sci. Eng. A 324 (2002)
113.
[4] S. Kleiner, E. Ogris, O. Beffort, P.J. Uggowitzer, Adv. Eng. Mater. 9 (2003) 653658.

[5] K.N. Braszczynska-Malik,


The Study on the Shaping of the Microstructure of
MagnesiumAluminium Alloys, WIPMiFS PCz., Czestochowa, 2005 (in Polish).
[6] J. Bursik, M. Svoboda, Microchim. Acta 139 (2002) 39.
[7] S. Celotto, Acta Mater. 48 (2000) 1775.
[8] M.A. Gharghouri, G.C. Weatherly, D.J. Embury, Philos. Mag. 78 (1998) 1137.
[9] S. Celotto, T.J. Bastow, Acta Mater. 49 (2001) 41.

[10] K.N. Braszczynska-Malik,


J. Alloys Compd. 477 (2009) 870.
[11] T.J. Brastow, M.E. Smith, J. Phys. Condens. Matter. 7 (1995) 49294937.
[12] E. Cerri, S. Barbagallo, Mater. Lett. 56 (2002) 716720.
[13] M.X. Zhang, P.M. Kelly, Scripta Mater. 48 (2003) 647652.
[14] J.F. Nie, X.L. Xiao, C.P. Luo, B.C. Muddle, Micron 32 (2001) 857863.
[15] D. Duly, M.C. Cheynet, Y. Brechet, Acta Metall. Mater. 42 (1994) 38433847.
[16] R.Z. Valiev, R.K. Islamgaliev, I.V. Alexandrov, Prog. Mater. Sci. 45 (2000) 103.
[17] R.Z. Valiev, T.G. Langdon, Prog. Mater. Sci. 51 (2006) 881.
[18] L. Jiang, M.T. Perez-Prado, P.A. Gruber, E. Arzt, O.A. Ruano, M.E. Kassner, Acta
Mater. 56 (2008) 1228.
[19] Y. Saito, H. Utsonomiya, N. Tsuji, T. Sakai, Acta Mater. 47 (1999) 579.
[20] A.P. Zhilyaev, T.G. Langdon, Prog. Mater. Sci. 53 (2008) 893.
[21] A.P. Zhilyaev, G.V. Nurislamova, B.K. Kim, M.D. Baro, J.A. Szpunar, T.G. Langdon,
Acta Mater. 51 (2003) 753.
[22] N.V. Ravi Kumar, J.J. Blandin, C. Desrayaund, F. Montheillet, M. Surey, Mater. Sci.
Eng. A 359 (2003) 150.
[23] M. Janecek, M. Popov, M.G. Krieger, R.J. Hellmig, Y. Estrin, Mater. Sci. Eng. A 462
(2007) 116.
[24] C.W. Su, L. Lu, M.O. Lai, Mater. Sci. Eng. A 434 (2006) 227.
[25] O. Kulyasova, R. Islamgaliev, B. Mingler, M. Zehetbauer, Mater. Sci. Eng. A 503
(2009) 176.
[26] Y. Miyahara, Z. Horita, T.G. Langdon, Mater. Sci. Eng. A 420 (2006) 240.
[27] X.-S. Wang, L. Jin, Y. Li, X.-W. Guo, Mater. Lett. 62 (2008) 1856.
[28] S.H. Kang, Y.S. Lee, J.H. Lee, J. Mater. Proc. Technol. 201 (2008) 436.
[29] R.B. Figueiredo, T.G. Langdon, Mater. Sci. Eng. A 501 (2009) 105.

[30] K.N. Braszczynska-Malik,


L. Froyen, Zeit. Metall. 96 (2005) 913.
[31] B. Chen, D.-L. Lin, L. Jina, X.-Q. Zeng, L. Chen, Mater. Sci. Eng. A 483484 (2008)
113.

268
[32]
[33]
[34]
[35]
[36]

K.N. Braszczy
nska-Malik / Journal of Alloys and Compounds 487 (2009) 263268
K. Mthis, J. Gubicza, N.H. Nam, J. Alloys Compd. 394 (2005) 194.

K.N. Braszczynska-Malik,
Zeit. Metall. 93 (2002) 845.
K. Mthis, K. Nyilas, A. Axt, Acta Mater. 52 (2004) 2889.
C.H. Woo, J. Nucl. Mater. 276 (2000) 90.
J. Koike, T. Kobayashi, T. Mukai, H. Wanatabe, M. Suzuki, K. Maruyama, K.
Higashi, Acta Mater. 51 (2003) 2055.

[37] M.H. Yoo, S.R. Agnew, J.R. Morris, K.M. Ho, Mater. Sci. Eng. A 319321 (2001)
87.
[38] J.W. Christian, S. Mahajan, Prog. Mater. Sci. 39 (1995) 1.

[39] K.N. Braszczynska-Malik,


L. Litynska,
A. Zyska, W. Baliga, Mater. Chem. Phys. 81
(2003) 26.

Vous aimerez peut-être aussi