Sie sind auf Seite 1von 9

Materials Science and Engineering A 405 (2005) 6573

Age hardening of a sintered AlCuMgSi(Sn) alloy


D. Kent a,b, , G.B. Schaffer a , J. Drennan b
a

Division of Materials, School of Engineering, The University of Queensland, Qld 4072, Australia
b Centre for Microscopy & Microanaylsis, The University of Queensland, Qld 4072, Australia
Received in revised form 12 May 2005; accepted 19 May 2005

Abstract
The age hardening response of a sintered Al3.8 wt% Cu1.0 wt% Mg0.70 wt% Si alloy with and without 0.1 wt% Sn was investigated. The
sequence of precipitation was characterised using transmission electron microscopy. The ageing response of the sintered AlCuMgSi(Sn)
alloy is similar to that of cognate wrought 2xxx series alloys. Peak hardness was associated with a fine, uniform dispersion of lath shaped
precipitates, believed to be either the  or Q phase, oriented along 0 1 0 directions and  plates lying on {0 0 1} planes. Natural ageing
also resulted in comparable behaviour to that observed in wrought alloys. Porosity in the powder metallurgy alloys did not significantly affect
the kinetics of precipitation during artificial ageing. Trace levels of tin, used to aid sintering, slightly reduced the hardening response of the
alloy. However, this was compensated for by significant improvements in density and hardness.
2005 Elsevier B.V. All rights reserved.
Keywords: Age hardening; Aluminium; Sintering; Powder metallurgy

1. Introduction
In recent years, the transport industry has been revolutionised through the use of light weight metals and plastics
[1,2]. There is particular interest in aluminium powder metallurgy (PM), as it is a means by which complex, net-shape,
light weight components can be produced cost effectively
[3]. Traditionally, aluminium powder products were thought
difficult to sinter and their properties deemed poor. Recent
research has found that it is possible to improve sintering of
pressed aluminium powders through the use of trace element
additions [47]. These alloys show a marked improvement in
strength and density to that of traditional press and sinter aluminium alloys. Specifically, the sintering response and subsequent material properties obtained using a typical 2xxx series
alloy (Al4.4Cu0.8Si0.5Mg) has been improved through
the addition of trace amounts of tin [5].
Current commercial aluminium powder metallurgy alloys
are based on the wrought 6xxx series (AlMgSi) and 2xxx
series (AlCuMg) of alloys [3], both of which exhibit age

Corresponding author.
E-mail address: d.kent@uq.edu.au (D. Kent).

0921-5093/$ see front matter 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2005.05.104

hardening. However, very little research has been conducted


into age hardening of sintered aluminium alloys. This work
was, therefore, undertaken to ascertain whether data on precipitation hardening of similar wrought aluminium alloys is
applicable to the precipitation hardening of the PM alloy.
The effects of increased levels of porosity and trace additions of tin used to aid sintering were also examined.
Precipitation in AlCuMgSi alloys is complex and a
large variety of precipitate phases may form. Not all reactions are understood and not all of the precipitates have been
characterised. The precipitates range from binary and ternary
phases similar to those found in lower order alloys, to complex quaternary phases. Equilibrium phases which may form
include (Al, Cu), (Mg, Si), S (Al, Cu, Mg), Q (Al, Cu,
Mg, Si) and Si, depending on the composition of the alloy [8].
At intermediate ageing temperatures (130200 C) strengthening in AlCuMgSi alloys with Cu 4 wt% and a high
Cu:Mg ratio has been shown to be in part due to the 
(Al2 Cu) phase [914]. The phase forms as large numbers
of fine platelets lying on {0 0 1} planes, uniformly distributed throughout the microstructure. In conjunction with
 , the peak-aged microstructure of these alloys contains fine
rod or lath-like precipitates oriented along 0 1 0 direc-

66

D. Kent et al. / Materials Science and Engineering A 405 (2005) 6573

tions. Due to its fine nature, the identity and composition


of this phase is more controversial. It has been proposed
to be either the S [14] or Q ( ) phase [10,13,15]. However, Eskin et al. [9,11,16] have shown that interpretation
of precipitation in AlCuMgSi alloys must be done with
respect to the composition of the supersaturated solid solution
(SSSS). X-ray microprobe analysis was used to measure the
composition of the SSSS in AlCuMgSi alloys containing 4 wt% copper and correlated to the phases formed during
artificial age hardening [9,11]. It was shown that the precipitate microstructure, which formed in alloys with an excess
of magnesium (Mg:Si > 1.73 wt%), with respect to the stoichiometry of the Mg2 Si phase, could be predicted fairly well
from the equilibrium phase diagram. However, alloys with an
excess of silicon (Mg:Si < 1.73 wt%) contain the  phase at
peak hardness, instead of a metastable form of the Q phase, as
would be predicted by the equilibrium phase diagram. Hence,
peak hardness was attributed to the presence of  platelets
lying on {0 0 1} planes and  laths oriented along 0 1 0
directions. Those alloys with an excess of silicon were found
to exhibit the highest levels of strengthening and absolute
hardness [11].

2. Experimental methods
The aluminium alloy used in this study was supplied by
Ampal Inc. as a pre-mix of elemental powders with a nominal
composition of Al3.8 wt% Cu1.0 wt% Mg0.70 wt% Si,
plus 1.5 wt% Acrawax as a die lubricant. To enable testing and
comparison of the age hardening response with and without
tin, air atomised tin powder was added to the alloy powders
prior to mixing. The tin concentration was 0.1 wt%. The alloy
containing tin is known as Ampalloy 2712.
The alloy powders were mixed for 30 min in a Turbula
mixer. The powders were compacted at 200 MPa using a
floating die and a hand operated Carver hydraulic press. The
powder compacts were generally prepared in the form of disc
shaped samples (10 mm 20 mm), though some smaller
cylindrical samples (10 mm 10 mm) were used for investigating hardness after short artificial ageing periods.
Sintering was performed in a sealed Modutemp horizontal
tube furnace, controlled by a Eurotherm BTC 9090 controller.
The samples were sintered under a high purity nitrogen atmosphere with a dewpoint of <60 C and <5 ppm O2 . The
sintering was done in two stages. The samples were heated
at 20 C min1 to 300 C and held at this temperature for 1 h
to allow de-waxing. They were then heated at 20 C min1
to 590 C, sintered at this temperature for 1 h and furnace
cooled to below 200 C, after which they were removed from
the furnace to air cool.
Solution treatment was performed in a sealed vertical tube
furnace, controlled by a Eurotherm BTC 9090 controller. The
samples were solution treated by heating at 20 C min1 to
500 C and holding for 1 h under a dry nitrogen atmosphere.
They were quenched directly into water at room temperature.

The furnace was fitted with an electro-magnetic specimen


holder, which when released, dropped the samples through
an aluminium foil seal directly into the water below, enabling
rapid quenching rates. Samples were either aged immediately
or stored in liquid nitrogen for future use.
Artificial ageing of Ampalloy 2712 was conducted at
20 C intervals from 140 to 200 C for periods up to 840 h
(35 days). Natural ageing was conducted on the alloy for
times in excess of 10,000 h (>400 days). Investigations were
also conducted into a tin free version of the alloy at 180 C,
140 C and room temperature. Two methods were used to age
the specimens. For longer ageing times (>30 min) large disc
shaped samples (10 mm 20 mm) were heated in a recirculating air furnace. For short ageing times (<30 min) smaller
cylindrical samples (10 mm 10 mm) were immersed in a
tempering oil bath, aged and water quenched.
Hardness testing was conducted on a Buehler Rockwell
hardness tester using the Rockwell B (HRB) scale with a
100 kg load and a 1/16 ball indenter. Calibration was performed on standardised testing blocks prior to taking each set
of measurements. The testing surfaces of the samples were
ground to grade 500 SiC grit to remove the outer oxide layers and to provide a consistent surface finish. Each reported
value is the average of at least eight separate measurements,
taken from both sides of the cylinders.
Sintered densities were determined by the Archimedes
method as per Metal Powder Industries Standard 42 [17].
This involves measurement of the closed porosity of the sintered body. However, ethanol was used instead of water to
prevent reaction with the alloy elements. Measurement of
sample weights was performed to an accuracy of 0.0001 g.
Theoretical densities, t , were estimated by:
t =

t


i=1

ni
n

(2.1)

where t is the number of elements in the alloy, i the density of


element i, ni calculated from Eq. (2.2) is the wt% of element i
(Wi ) divided by its molar mass (Mi ) and n is the total number
of moles given by Eq. (2.3).
ni =
n=

Wi
Mi

t


ni

(2.2)

(2.3)

i=1

Thin foils for transmission electron microscopy (TEM)


were prepared from sintered samples. Slices were cut from
the samples using a Struers Accutom-50 fitted with a diamond
impregnated cutting wheel (430-CA). All cutting was performed under a constant jet of coolant and the cutting speed
used was 0.06 mms1 . The slices were ground to 100 m
thick with 1200 grit SiC paper. Discs 3 mm in diameter were
hand punched from the slices. Solution treatment and ageing of samples for TEM was performed on the discs prior to
electro-polishing. The treated and aged discs were finished

D. Kent et al. / Materials Science and Engineering A 405 (2005) 6573

Fig. 1. Ageing response of Ampalloy 2712 at various ageing temperatures.

with a 5000 grit SiC paper. The discs were electro-polished


using a Tenupol-3 twin jet electro-polisher. This was conducted at 10 V with a 25% nitric acid/75% methanol electrolyte at a temperature of 30 C and a flow-rate of 6.5.
Samples were washed in a series of methanol baths immediately after polishing. Transmission electron microscopy was
performed using a JEOL 2010 operating at 200 kV fitted with
a Gatan double tilt specimen holder. Bright field TEM images
were obtained in conjunction with corresponding selected
area electron diffraction (SAED) patterns.
3. Results and discussion
3.1. Age hardening response of Ampalloy 2712
3.1.1. Articial age hardening
The artificial age hardening response of Ampalloy 2712,
which contains tin, is shown in Fig. 1. It is characteristic of wrought alloys with similar compositions.1 Greatest
strengthening was associated with ageing at 140 C as it
results in a fine dispersion of precipitates most closely resembling that of the critical dispersion necessary for maximum
strengthening [18]. The time taken to achieve peak hardness
increased at low ageing temperatures because the rate of precipitation is largely controlled by diffusion of solute elements,
which is highly dependent on temperature [19].
A typical microstructure obtained from Ampalloy 2712
after ageing for 3 h at 160 C is shown in Fig. 2. At this stage,
many small difficult to resolve precipitates, around 1.52 nm
in diameter are distributed throughout the matrix. They were
observed after the initial rapid hardening rate slowed shortly
before a second marked increase in the hardening rate. Their
formation was associated with much of the overall hardening.
The corresponding SAED pattern exhibits faint continuous
streaking along 1 0 0 through matrix spots and the {1 1 0}
1 The behaviour of wrought alloy 2014 was used for comparison with
Ampalloy 2712 as this is the wrought alloy that it closely resembles. However, Ampalloy 2712 contains 0.1% tin, slightly less copper and slightly
more magnesium than 2014.

67

Fig. 2. Bright field TEM image of Ampalloy 2712 aged 3 h at 160 C viewed
along 0 0 1 showing coherent precipitates. Inset: corresponding SAED
pattern exhibiting streaking through {1 1 0} positions.

positions. Faint diffraction spots can also be seen at the


{1 1 0} positions. The small size of the precipitates and the
streaking observed in the SAED pattern is consistent with that
of coherent rod shaped precipitates oriented along 0 1 0 ,
possibly GPB zones and/or an early form of the needle/lath
type phase observed later in the artificial ageing sequence of
the alloy. Faint diffraction spots observed in the SAED pattern
at forbidden {1 1 0} positions most likely result from oxide
reflections [20] as no signs of precipitation were observed to
be associated with them. Evidence to support the formation
of GPB zones during the early stages of ageing was noted
by Bonfield and Datta [12] during TEM investigations of a
wrought 2014 alloy. Dutta et al. [13] also reported effects consistent with formation of coherent zones in the early stages
of ageing wrought 2014 alloys, which they attributed to fine
scale GPB zones.
A typical microstructure obtained after ageing for 6 h at
160 C is shown in Fig. 3. Coherent needle shaped precipitates oriented along 0 1 0 have formed throughout the
matrix. Streaking in the SAED pattern is more defined than it
was previously and now exhibits intensity maxima. These
effects are attributed to the evolution of the coherent rod
shaped precipitates observed earlier in the ageing sequence.
Intensity maxima observed about {1 1 0} positions in the
SAED pattern are believed not to relate to the start of 
precipitation as there is no sign of characteristic {1 1 0}
diffraction spots. Based on interpretations by Dutta et al. [13]
and Gao et al. [10] these needle shaped precipitates may be
the Q phase. However, on the basis of studies done on the
decomposition of the SSSS by Eskin et al. [9,11,16], the needle shaped precipitates are more likely to be the  phase.
A typical peak-aged microstructure obtained after ageing for 18 h at 160 C is shown in Fig. 4(a) with the inset
showing the corresponding SAED pattern. At this stage, two
primary precipitate morphologies can be seen. These are a
plate shaped phase lying on {0 0 1} planes and a fine rod
or lath like phase oriented along 0 1 0 . The corresponding
SAED pattern exhibits superlattice spots at {1 1 0} together
with continuous streaking along 1 0 0 and intensity max-

68

D. Kent et al. / Materials Science and Engineering A 405 (2005) 6573

Fig. 3. Bright field TEM image of Ampalloy 2712 aged 6 h at 160 C viewed
near 0 0 1 exhibiting three orthogonal variations of a needle shaped precipitate phase oriented along 0 1 0 . Inset: corresponding SAED pattern
with streaks through {1 1 0} positions that exhibit maxima about these
positions.

ima surrounding the {1 1 0} positions. The plate shaped


phase observed in peak-aged microstructures was found to be
the  phase. The SAED pattern, inset Fig. 4(a), exhibits characteristic {1 1 0} diffraction spots with surrounding maxima
corresponding to the (1 0 1) and (0 1 1) variants and the
effects of double diffraction. This is consistent with findings
by other researchers from investigations into wrought 2014
alloys [914,16].

The identity of the lath shaped precipitates oriented along


0 1 0 is less certain. The scale and distribution of these
precipitates is consistent with their having developed directly
from the needle shaped phase observed in the early stages of
ageing. A lath shaped phase of similar morphology is generally acknowledged to form as a precursor to the Q phase
during ageing of wrought 2014 type alloys [9,10,13]. Some
researchers believe this phase to be an early form of the complex Q phase and so designate it as Q [10,13] while others
assert it is actually the  phase [9,11,16]. Dutta et al. [13] and
Gao et al. [10] propose that the precipitates are the Q phase,
principally on the basis that the equilibrium Q phase forms as
laths oriented along 0 1 0 in over-aged microstructures of
these alloys. Eskin et al. [9,11,16] contend that the  phase is
responsible for hardening and may be a precursor to the equilibrium Q phase in these alloys. Eskin [9] demonstrated that
the phases formed during ageing depend on the composition
of the SSSS and was able to correlate this with precipitate
microstructures for a wide range of compositions. Fig. 4(b)
shows a SAED pattern obtained from Ampalloy 2712 in
the peak-aged condition after prolonged exposure. The extra
diffraction spots indicated show good correspondence to the
{1 1 3} positions identified in an indexed diffraction scheme
for the  phase published by Eskin [9]. It is likely that the lath
shaped phase and corresponding diffraction effects observed
for peak-aged Ampalloy 2712 relate to the precipitation of
the  phase, although similar phenomena have also been
attributed to the Q phase. Even with a detailed chemical
analysis, identification of the phases is complicated by the
fact that a variety of compositions have been found for the
Q phase [21] and the  phase has been shown to be able to
incorporate copper [22].
The peak hardness of the alloy is slightly lower than that
for similar wrought 2014 alloys after a typical T6 ageing
treatment of 18 h at 160 C [23]. Nevertheless, the peak hardness in the sintered samples compared very favourably to
the wrought alloys, considering their increased porosity and

Fig. 4. Ampalloy 2712 aged 16 h at 160 C: (a) bright field TEM image viewed near 0 0 1 showing two primary precipitate morphologies, a plate shaped
phase on {0 0 1} and a lath shaped phase oriented along 0 1 0 . Inset: corresponding SAED pattern exhibiting diffraction spots at {1 1 0} positions with
maxima surrounding these positions. (b) SAED pattern obtained after prolonged exposure. The diffraction spot labelled is thought to relate to (1 1 3) .

D. Kent et al. / Materials Science and Engineering A 405 (2005) 6573

Fig. 5. Ampalloy 2712 aged 96 h at 160 C: (a) bright field TEM image
viewed near 0 0 1 revealing precipitate microstructure. Inset: corresponding SAED pattern exhibiting diffraction spots at {1 1 0} positions and
streaking along 1 0 0 directions. (b) Higher magnification image of (a)
showing plate shaped precipitates, marked A, and lath shaped precipitates,
marked B.

slightly reduced copper concentration. Researchers investigating early sintered PM alloys also noted that 2xxx series
type alloy compositions responded well to age hardening
[24,25]. However, those alloys differed significantly from the
alloy used in the present study in that they had a finer grain
size and were produced using greater compaction pressures
of the order of 8001000 MPa.
Microstructures typical of Ampalloy 2712 in the slightly
over-aged condition, obtained after ageing for 96 h at 160 C,
are shown in Fig. 5. Fig. 5(a) shows a plate shaped phase lying
on {0 0 1} planes and a lath shaped phase oriented along
0 1 0 directions, with the corresponding SAED pattern
inset. This pattern exhibits well-developed superlattice spots
at {1 1 0} together with continuous streaking and intensity maxima parallel to 1 0 0 about the {1 1 0} positions.
Fig. 5(b) shows the two primary precipitate morphologies at
a higher magnification, marked A and B. A is a plate shaped

69

precipitate viewed side on while B is a lath like precipitate


viewed end on. Note, the extensive strain fields around some
of the other A type precipitates and the presence of c-shaped
strain contrast on one side of the B type precipitates. These
strain fields observed in conjunction with the lath shaped precipitates are indicative of a loss of coherency on one side of
the lath. This supports findings by Eskin [9] who states that
upon softening, the coherent  phase is replaced by forms of
the semi-coherent  phase. One of these, the  C phase, has
an identical structure to the Q phase and varies only in composition [22,26]. Differentiation of lath shaped precipitates
observed here is, therefore, not possible without a detailed
chemical analysis.
Gao et al. [10] have noted that the number of needle or lath
shaped precipitates oriented along 0 1 0 which form during
ageing of wrought 2014 type alloys increases with increasing
levels of silicon. They also found that the number of  plates
increased with increasing levels of silicon and were generally
in contact with the lath shaped phase. It was proposed that the
lath shaped precipitates might provide heterogeneous nucleation sites for a dense and uniform dispersion of  . This is
consistent with the current results that show that the  phase
formed after the lath shaped precipitates and frequently in
contact with them.
The microstructure of the slightly over-aged alloy also
contained occasional large precipitates with Moire fringes
roughly aligned along [0 1 0] as exhibited in Fig. 6. Moire
fringes running through the particle aligned near [0 1 0] are
characteristic of plate shaped Si precipitates in aluminium
and results from the cubecube relationship between the
respective matrices [27]. The measured fringe spacing of
1.68 nm agrees well with the theoretical spacing of 1.62 nm
calculated from the respective {0 0 1} planar spacings. The
slight difference between the measured and theoretical Moire
fringe spacings is believed to relate to the slight misalignment
of the respective matrices observable in Fig. 6(a). Faulting in
the fringes in Fig. 6(b) is indicative of a dislocation type defect
[28]. Gao et al. [10] and Eskin et al. [16] have also observed

Fig. 6. Ampalloy 2712 aged 96 h at 160 C: (a) bright field TEM image viewed near 0 0 1 showing a large precipitate with Moire fringes roughly oriented
along [0 1 0] . (b) Higher magnification image of (a) revealing Moire fringes measured at 1.68 nm spacings.

70

D. Kent et al. / Materials Science and Engineering A 405 (2005) 6573

Fig. 8. Bright field TEM image of Ampalloy 2712 aged 12 days at room
temperature viewed near 0 0 1 exhibiting mottled microstructure textured
along 0 1 0 . Inset: corresponding SAED pattern with streaking between
matrix spots oriented along 1 0 0 directions.

Fig. 7. Bright field TEM image of Ampalloy 2712 in the very over-aged
condition (aged 150 h at 200 C) showing large coarse plate and lath shaped
precipitates, viewed near 0 0 1 . Inset: corresponding SAED pattern showing diffraction spots at and surrounding the {1 1 0} positions.

plate shaped Si precipitates in an AlCuMgSi alloy with


high levels of Si in the peak-aged condition.
A microstructure typical of a highly over-aged state,
shown in Fig. 7, consists of large plates of  lying on
{0 0 1} together with some coarse laths/rods oriented along
0 1 0 . The corresponding SAED pattern reveals spots at
{1 0 0} indicative of the  phase and the surrounding maxima along 1 0 0 have thickened due to coarsening of the
plates. Dutta et al. [13] also note that coarse  plates are
present in highly over-aged microstructures in these alloys.
The coarse rods/laths oriented along 0 1 0 are believed to
be the equilibrium Q phase, which has been identified in overaged wrought alloys [10].
In summary, the artificial ageing response of Ampalloy
2712 was characteristic of similar wrought alloys. High
levels and rapid rates of hardening occurred for both
natural and artificial ageing. Large numbers of coherent
GPB type zones form during the early stages of ageing.
These evolve into coherent needle shaped precipitates
oriented along 0 1 0 . Due to their fine scale and the
similarities in the structure and composition of the Q and
 phases, it is difficult to positively identify the needle
shaped phase. The needle-shaped precipitates are believed
to evolve into the lath shaped precipitates observed later in
the precipitation sequence. Peak hardness is associated with
a fine uniform dispersion of lath shaped precipitates oriented
along 0 1 0 and  plates lying on {0 0 1} planes. The
over-aged microstructure contains coarse  plates lying
on {0 0 1} . The lath shaped precipitates oriented along

0 1 0 , observed earlier in the precipitation sequence lose


coherency during over-ageing and are believed to eventually
evolve into the equilibrium Q phase. Si precipitates were
also found to be present in over-aged microstructures.
3.1.2. Natural age hardening
Ampalloy 2712 exhibited high rates of hardening during
natural ageing. The peak hardness in the T4 condition is
80% of that in the T6 condition. The hardness increased
from around 20 HRB to over 60 HRB in the first 24 h. The
hardness then remained at 65 HRB (see Fig. 1). These rates
and levels of hardening are similar to those reported to occur
in naturally aged wrought 2014 alloys [29]. High levels of
hardening were also reported during natural ageing of early
sintered alloys, similar in composition to 2xxx series wrought
alloys [24,25].
Figs. 8 and 9 were obtained from a sample of Ampalloy
2712 aged 12 days at room temperature. The corresponding
SAED patterns exhibit continuous streaking along 1 0 0
directions through the aluminium matrix spots, which are
characteristic of GP zones. Fig. 8 exhibits a mottled appearance due to the effects of large numbers of these coherent
precipitates. GP zones have previously been noted to form
in conjunction with GPB zones during natural ageing of
an Al2.5Cu1.2Mg0.24Si alloy [30]. The formation of
GP zones during natural ageing contrasts with findings for
wrought 2014 alloys. Bonfield and Datta [29] examined natural ageing in samples aged from 1 to 1000 h (40 days). It was
observed that the microstructure typically exhibited a mottled appearance but did not reveal any precipitation effects in
SAED patterns, which was attributed solely to the presence
of fine scale GPB type zones.
Numerous helical dislocations were also observed
throughout the microstructure, oriented along 0 1 0 as seen
in Fig. 9. Bonfield and Datta [29] have observed helical dislocations during natural ageing, however, in that case they
were oriented along 1 1 0 , while these are aligned along

D. Kent et al. / Materials Science and Engineering A 405 (2005) 6573

71

Fig. 10. Typical optical microstructure of Ampalloy 2712 at 96% theoretical density. The black intergranular regions are pores.
Fig. 9. Ampalloy 2712 aged 12 days at room temperature. Bright field
TEM image viewed near 0 0 1 showing helical dislocations oriented
along 0 1 0 directions. Inset: corresponding SAED pattern with streaking
between matrix spots oriented along 1 0 0 directions.

0 1 0 . Helical dislocations oriented along 0 1 0 have


previously been linked to the addition of silicon to AlCuMg
alloys [30]. Wilson [30] proposed that the helical dislocations
formed as a result of the pinning of screw dislocations by silicon atoms along their length. Hutchinson and Ringer [31] also
observed helical dislocations formed in an Al2.5Cu1.5Mg
alloy with additions of silicon. They noted that the dislocation helices form mainly along 1 1 0 with small amounts
of silicon (0.1 wt%), but occur increasingly along 0 1 0
directions with the larger amounts of silicon. This is consistent with the current alloy, which has relatively high silicon
content.
3.2. Effects of porosity
Compaction pressure was varied in order to alter the levels
of porosity in sintered samples of Ampalloy 2712. A typical sintered microstructure for Ampalloy 2712 compacted at
200 MPa is shown in Fig. 10 (4% porosity). Samples of the
alloy with porosity levels ranging from 18.5 to 3.5% were
solution treated and aged. Artificial ageing was conducted at
140 C and 180 C until peak hardness was achieved. Naturally aged specimens were aged for more than a week prior
to testing. Fig. 11 shows that the maximum hardness varies
linearly with porosity over the range tested.
Porosity had little effect upon the age hardening response
of Ampalloy 2712, other than its direct effect on hardness,
which was mainly related to the reduction in effective load
bearing area. That the slopes of the curves in Fig. 11 are
similar for all ageing temperatures suggests that porosity is
the primary factor influencing the hardness values.
The times taken to reach peak hardness were similar to
those observed in near dense wrought 2014 alloys [12,13].
Hence, porosity did not have an obvious effect on the rate

of the ageing response. This contrasts with work by Davies


and Farzin-Nia [32] who found that the rates of artificial age
hardening observed in sintered AlCu and AlCuSi alloys
generally exceeded those of their wrought counterparts. This
was attributed to an increase in the number of active sites
for heterogeneous nucleation of the principle strengthening
phases,  and  .
The  phase is also partly responsible for hardening
in peak-aged wrought 2014 alloys [914,16] and has been
shown to form heterogeneously on lath shaped precipitates
which form during the early stages of hardening [10,14]. The
 phase was also observed to form after and frequently in contact with a similar lath shaped precipitate phase in Ampalloy
2712. Hence, it is proposed that increased porosity has little
effect on the rate of precipitation in Ampalloy 2712, as it is
already enhanced though heterogeneous nucleation by other
precipitate phases.
3.3. Effects of tin
Samples were prepared from identical powders but without trace additions of tin. The sintered density of standard

Fig. 11. Linear relationship between maximum hardness and porosity for
porosity levels between 3.5 and 20% at various ageing temperatures for
Ampalloy 2712.

72

D. Kent et al. / Materials Science and Engineering A 405 (2005) 6573

been shown to suppress its action in these alloys [35]. Thus,


tin in Ampalloy 2712 may act to reduce the effective magnesium concentration available to form precipitate phases
responsible for hardening.
The precipitate microstructure of the tin free alloy, shown
in Fig. 13, is similar to that of the standard Ampalloy 2712
alloy containing tin. Plates of  on {0 0 1} planes, which
exhibit characteristic {1 1 0} diffraction spots in the SAED
pattern were observed in conjunction with an unidentified
rod shaped phase oriented along 0 1 0 . Additionally, the
precipitation sequence for Ampalloy 2712 was shown to be
similar to wrought alloys, which do not generally contain tin.
Fig. 12. Hardness curves for Ampalloy 2712 and for a tin free version of the
same alloy are essentially similar at all ageing temperatures.

4. Conclusions

Ampalloy 2712 samples was typically 96% while for the


tin free alloy it was typically 94%. Ageing was conducted
at 180 C, 140 C and at room temperature. Fig. 12 shows
hardness curves for the tin containing and tin free versions of
the alloy. The hardness curves are almost identical, which is
surprising given the 2% difference in the relative densities of
the alloys. The relationship between porosity and hardness
for Ampalloy 2712, exhibited in Fig. 11, shows that a 2%
difference in the level of porosity corresponds to a difference
of almost 8% in hardness. This is consistent with the results
of Sercombe [33]. This suggests that tin may slightly reduce
the age hardening response of Ampalloy 2712.
In contrast, trace additions of tin are known to dramatically enhance the artificial ageing response of wrought AlCu
alloys [3436], although small additions of magnesium have

The powder metallurgy alloy, Ampalloy 2712, responds


well to both natural and artificial ageing treatments, exhibiting rapid rates and high levels of hardening. Research into
precipitation hardening in traditional wrought alloys was
found, in general, to also apply to the sintered alloy. The ageing response of the sintered alloy, the precipitation sequence
and phases formed were akin to those described in the literature. Peak hardness was associated with a fine uniform
dispersion of lath shaped precipitates oriented along 0 1 0
directions and  plates lying on {0 0 1} planes. Natural age
hardening was related to the formation of fine scale coherent precipitates, believed to be GP zones. Porosity has little
effect on the ageing response of the alloy other than its direct
effect on hardness, related to the effective load bearing area.
Tin additions may slightly reduce the hardening of the alloy,
although benefits associated with enhanced sintering generally outweigh this effect.

Acknowledgement
This work was funded by Ampal Inc., a member of the
US Bronze Group of Companies, whose support is gratefully
acknowledged.
References

Fig. 13. Bright field TEM image of tin free alloy aged 294 h at 140 C viewed
near 0 0 1 . It shows plate shaped precipitates on {0 0 1} planes together
with needle/rod shaped precipitates oriented along 0 1 0 directions and
the corresponding SAED pattern (inset).

[1] I.J. Polmear, Light AlloysMetallurgy of the Light Metals, third


ed., Arnold, London, 1995.
[2] D. Slavnich, Auto. Eng. 27 (2002) 5260.
[3] G.B. Schaffer, Mater. Forum. 24 (2000) 109124.
[4] T.B. Sercombe, G.B. Schaffer, Acta Mater. 47 (1999) 689697.
[5] T.B. Sercombe, G.B. Schaffer, Mater. Sci. Eng. A 268 (1999) 3239.
[6] G.B. Schaffer, S.H. Huo, Powder Metall. 42 (1999) 219226.
[7] G.B. Schaffer, S.H. Huo, J. Drennan, G.J. Auchterlonie, Acta Mater.
49 (2001) 26712678.
[8] J. Crowther, J. Inst. Met. 76 (19491950) 201236.
[9] D.G. Eskin, J. Mater. Sci. 38 (2003) 279290.
[10] X. Gao, J.F. Nie, B.C. Muddle, Mater. Sci. Forum 217222 (1996)
12511256.
[11] D.G. Eskin, Z. Metallkd. 83 (1992) 762765.
[12] W. Bonfield, P.K. Datta, J. Mater. Sci. 11 (1976) 16611666.

D. Kent et al. / Materials Science and Engineering A 405 (2005) 6573


[13] I. Dutta, C.P. Harper, G. Dutta, Metall. Trans. 25A (1994)
15911602.
[14] G. Rionto, A. Zanada, Mater. Lett. 37 (1998) 241245.
[15] B. Dubost, J. Bouvaist, M. Reboul, Aluminium Alloys: Their Physical and Mechanical Properties, vol. II, Engineering Materials Advisory Service Ltd., Chalottesville, 1986, pp. 11091123.
[16] D.G. Eskin, V.S. Zolotorevskii, V.V. Istomin-Kastrovskii, A.A.
Aksenov, Russ. Metall. 2 (1989) 111115.
[17] Method for determination of density of compacted or sintered powder metallurgy products, Metal Powders Industry Federation, Princeton New Jersey, MPIF Standard 42 (1997).
[18] E. Orowan, in: J.W. Martin (Ed.), Precipitation Hardening, Butterworth Heinemann, Oxford, 1948, pp. 124125.
[19] J.W. Martin, Precipitation Hardening, second ed., Butterworth Heinemann, Oxford, 1998.
[20] J.K. Park, A.J. Ardel, Metall. Trans. 14A (1983) 19571965.
[21] L.F. Mondolfo, Aluminium AlloysStructure and Properties, Butterworths, London, 1976.
[22] K. Murayama, K. Hono, W.F. Miao, D.E. Laughlin, Metall. Trans.
32A (2001) 239246.
[23] J.R. Davis, Aluminium and Aluminium Alloys, third ed., ASM International, Ohio, 1993.

73

[24] G.D. Cremer, J.J. Cordiano, Trans. Am. Inst. Min. Metall. Eng. 1574
(1943) 152165.
[25] C.G. Goetzel, Treatise on Powder Metallurgy, Interscience Publishers
Inc., New York, 1949.
[26] C. Cayron, P.A. Buffat, Acta Mater. 48 (2000) 26392653.
[27] D.R.G. Mitchell, Nucl. Instrum. Methods Phys. B 140 (1998)
107118.
[28] D.B. Williams, C.B. Carter, Transmission Electron Microscopy,
Plenum Press, New York, 1996.
[29] W. Bonfield, P.K. Datta, J. Mater. Sci. 12 (1977) 10501052.
[30] R.N. Wilson, J. Inst. Met. 97 (1969) 8086.
[31] C.R. Hutchinson, S.P. Ringer, Metall. Trans. 31A (2000) 2721
2733.
[32] B.L. Davies, F. Farzin-Nia, Int. J. Powder Metall. 19 (1983) 197
209.
[33] T.B. Sercombe, Non-conventional sintered aluminium alloy powders,
Thesis, The Department of Mining, Minerals and Materials Engineering, The University of Queensland, Brisbane, 1998.
[34] H.K. Hardy, J. Inst. Met. 80 (19511952) 483492.
[35] H.K. Hardy, J. Inst. Met. 78 (19501951) 169194.
[36] S.P. Ringer, K. Hono, T. Sakurai, Metall. Trans. 26A (1995)
22072217.

Das könnte Ihnen auch gefallen