You are on page 1of 6

Foaming characteristics of Al–Si–Mg (LM25)

alloy prepared by liquid metal processing


R. Nadella*, S. N. Sahu and A. A. Gokhale
LM25 (Al–Si–Mg) foams were made by melt processing route using Al foam scrap turnings as
thickening agent in place of aluminium powder or Ca metal. The effects of foaming temperature
and excess Mg content on melt expansion and cell structure were studied. The selection of
foaming temperature significantly influenced foam decay. While severe collapse occurred at
670uC, foam decay is hardly observed at 640uC. Moreover, presence of excess Mg (3 wt-%) in
this alloy enhanced collapse at 670uC, presumably due to its surface tension lowering effect.
Despite foam collapse, a well defined and uniform cell size could be obtained at 670uC as
compared to the foams obtained at 640uC. Uniaxial compression tests showed serrated plateau
stress–strain behaviour with extensive localised but non-catastrophic damage leading to
progressive breaking away of foam pieces. Overall, these experiments showed the potential to
manufacture good quality LM25 foam using inexpensive Al foam turnings.
Keywords: Al–Si alloy, Al foam, Foam collapse, Compression

Introduction elements on the quality of Al foams has not been


investigated.
Aluminium foams are ultra light weight materials, which The aim of the present work is twofold. First, to check
have a high potential for structural and functional the feasibility of preparing Al alloy foam with Al–Ca
applications involving sound and high energy absorp- foam scrap turnings as the source of oxides. In this way,
tion. Among various manufacturing techniques, liquid in addition to cost reduction, major presence of Ca,
metal processing, involving the addition of foaming which is not an alloying element in aluminium alloys,
agent to the thickened aluminium melt, has gained an can be avoided. The other aim is to compare the foam
attention due to its simplicity and low cost. Although quality with and without Mg, which oxidises readily and
large amount of literature is available on Al foams, very is a surface tension reducing element in Al. The work
limited work1–3 exists on foams of one of the most also tries to examine the interplay between the Mg
prominent cast alloys, LM25 (A356). Yang and Nakae content and the foaming temperature.
have carried out foaming experiments by varying TiH2
content (0?5–2?5%), and foaming temperature (620– Materials and experimental method
660uC) on A356 alloy with aluminium powder (5 wt-%
and 60 mm size) as the thickening agent.1 They suggested The materials used for the foam synthesis are LM25
an optimum combination of temperatures (640–650uC) alloy ingots (Al–6?9Si–0?28Mg–0?12Fe), Al foam turn-
and TiH2 levels (1–1?5%) for good foamability. A ings (Al–2Ca–0?72Ti) and TiH2 powder (all elemental
uniform cell structure with high void content (86%) compositions are described in wt-%). The foaming
could be obtained with 1% TiH2 foamed at 640uC. They experiments were carried out in a cylindrical clay
graphite crucible (inner diameter about 90 and
have also reported that rise in viscosity (by varying the
410 mm height) with 1?25 kg of alloy melt, following
stirring time) resulted in generation of more particles
the general principles of Alporas technique described
which aided in obtaining higher porosity levels in
elsewhere.4 Instead of Ca, pure aluminium foam turn-
foams.2 Yang et al.3 showed that in an Al–6?7Si–2?8Cu
ings of size range from 75 to 105 mm (4 wt-%) were
alloy with 2 wt-%Ca and 1 wt-%TiH2 foams with
added to the melt for thickening at 680uC and stirring
different pore structures could be produced by varying
was carried out at 1200 rev min21 for 10–13 min. Upon
the stirring time at a foaming temperature of 667uC
cooling to foaming temperature, 1?25 wt-%TiH2 was
(940 K). In spite of its potential cost savings, no
added to liquid metal and the temperature was main-
attempts to utilise foam scrap for Al foam production tained in order to facilitate the blowing agent (TiH2)
were reported. Also, the role of surface tension reducing decomposition and foaming process. Once the foaming
activity ceased, which was confirmed visually, the
crucible was removed from the furnace and allowed to
Light Alloy Casting Group, Defence Metallurgical Research Laboratory, cool in air for solidification of the foam. The holding
Kanchanbagh, Hyderabad 500 058, India time at the foaming temperature varied between 3 and
*Corresponding author, email nvravi_in@yahoo.co.uk 7 min. Keeping all other experimental conditions

ß 2010 Institute of Materials, Minerals and Mining


Published by Maney on behalf of the Institute
Received 27 February 2009; accepted 17 March 2009
908 DOI 10.1179/174328409X430500 Materials Science and Technology 2010 VOL 26 NO 8
Nadella et al. Foaming characteristics of Al–Si–Mg (LM25) alloy

a simple scale, as shown in Fig. 1. Such a large drop in


height cannot be attributed to thermal contraction or
solidification shrinkage.
The macrostructures of the foams in longitudinal
section (Fig. 2) showed well defined cell structure, except
for F-3 foam. This could have been expected since the
foam turnings contain foam stabilising oxides (due to
original foam processing) and a large amount of surface
oxides. Thus foam turnings seemed to have served the
role of Al powder used by other researchers1,2 as a
thickening agent. The vigorous stirring before foaming
must have helped in the creation of new oxides and
distribution of oxides present in foam turnings. In the
foams made at 670uC, folded areas were seen in the top
and side portions. This suggests that the drop in height
during cooling phase was accompanied by severe cell
collapse, and this terminology will be adopted to
describe the height drop in what follows.
In all cases, solid (unfoamed) portion was present in
1 Liquid foam decay of sample F-3 during cooling from the bottom region of the foam ingot. Similar observa-
670uC tions have been reported in the past and were attributed
to liquid drainage.2 Interestingly, chemical analysis
unchanged, two foaming temperatures, namely 640 and showed only 0?33%Ti in the solid bottom portion and
670uC, and three excess Mg levels, namely 0, 0?25/0?5 1–1?5% in the foamed portion. The lower Ti content in
and 3% were used (Table 1). the solid bottom was probably due to the lack of
The foaming ratio is computed as the ratio of total thorough mixing and consequently little penetration of
height of solid foam h to the height of the original melt TiH2 up to the bottom of the alloy melt, explaining lack
ho. For evaluation, the solid cylindrical foams were of foaming. However, considering that the bottom
sectioned longitudinally with electrodischarge sawing. portion of the melt did not foam at all, the foaming
Cell size measurements were carried out on scanned ratios were recomputed by deducting the height of the
sections of electrodischarge sawing cut foams by linear solid bottom from both numerator and denominator of
intercept method. In cases where the cell structure is not the ratios, as presented in Fig. 3.
easily discernible, an approximate estimate was made. In general, foaming ratios were higher at the foaming
Chemical composition was analysed through wet analy- temperature of 640uC than at 670uC. In the case of foam
sis. Metallography was carried out on selected foam F-3, the liquid foam expanded to a maximum height of
samples by optical and scanning electron microscopy 272 mm, as reported above (Fig. 1). This corresponds to
(SEM). a foaming ratio of 13?6 in the liquid state (after
Compression tests were carried out at an initial strain compensating for the unfoamed bottom). This value is
rate of 0?05 s21 on electrodischarge machined wire cut slightly higher than all the foaming ratios at 640uC, as
samples of 25625630 mm size, with longer dimension shown in Fig. 3. Thus, it can be inferred that the lower
corresponding to the foaming direction as well as foaming ratios at 670uC were not due to any inherent
loading direction. lack of foamability but due to foam collapse in the liquid
state.
Results and discussion The cell structure of the collapsed regions shown by
magnified image in Fig. 2 indicates that they are
Foam collapse and quality compressed under atmospheric pressure. This would be
Soon after attaining maximum expansion, liquid foams possible only if the cell internal pressure fell below the
made at 670uC dropped to various heights during atmospheric pressure, while the foams were in liquid
cooling, depending on the Mg content. Such a behaviour state. The former would decrease if hydrogen loss from
was not observed in foams made at 640uC. The height the cells by diffusion through the liquid cell walls to the
reduction was most severe in foam F-3, from initial atmosphere was significant. Considering that the hydro-
272 mm to final 62 mm in the solidified foam. The gen diffusion coefficient in liquid aluminium is very
variation of liquid foam height with time after attaining high,5 this mechanism is a strong possibility. The closer
maximum expansion was monitored for foam F-3 using the cells are to the external surface, the greater would be

Table 1 Details of foaming experiments on LM25 alloy

Foam no. Excess Mg, wt-% Cell size, mm Foaming temperature, uC

F-1 … 2.72 670


F-2 0.25 2.82
F-3 3 …
F-4 … 5.20 640
F-5 0.5 6.82*
F-6 3 3.64–6.51
*Approximate estimate from the macrostructure.

Materials Science and Technology 2010 VOL 26 NO 8 909


Nadella et al. Foaming characteristics of Al–Si–Mg (LM25) alloy

2 Macrostructures of LM25 foams produced at two foaming temperatures: excess Mg addition made to parent alloy is
indicated

the hydrogen loss. Indeed the collapsed regions are expected. This may further aid cell coalescence in the
found to be near the external surfaces. Coalescence upper regions and foam collapse. Whatever is the actual
between neighbouring cells during compression under mechanism of cell collapse, it is clear that in such cases,
atmospheric pressure is also likely. the oxide particles present in the cell walls, though
Effect of foaming temperature on collapse can be suitable to give a thermodynamically stable foam (as
explained as follows. Higher the foaming temperature, indicated by large foaming ratios in the liquid state),
greater is the diffusion coefficient of hydrogen in liquid were not adequate to give stability against atmospheric
aluminium and the longer it takes to reach the liquidus pressure and/or liquid drainage.
temperature, leading to greater loss of hydrogen and, Figures 2 and 3 show that even the standard LM25
therefore, greater amount of collapse as observed. alloy foamed well, i.e. excess Mg is not essential for
Similarly, if the foaming temperature is high, more formation of foam. However, in the absence of collapse,
liquid drainage and, therefore, cell wall thinning i.e. at 640uC foaming temperature, presence of excess
(especially from upper regions of the foam) can be Mg helped in increasing the foaming ratio, though no
systematic trend could be observed.
Effect of Mg on foaming behaviour of LM25 alloy
can be analysed as follows. The calculated liquidus
temperatures Al–Si alloys, with an algorithm developed
by Hernandez et al.,6 based on their chemical composi-
tions, showed that addition of Mg up to 3% to LM25
alloy does not lower its liquidus temperature by .0?2uC.
The corresponding superheats and therefore hydrogen
loss from liquid foam would be insensitive to the Mg
content (barring the small influence Mg may have on
hydrogen diffusion in liquid Al).
However, presence of Mg in LM25 modifies its
surface tension, which is an important physical property
(other than viscosity) influencing foaming behaviour in
gas blown liquid foams. Measured data by Goicochea
et al.7 at 973 K (700uC) indicates that addition of 8%Si
3 Foaming ratios of LM25 foams as varied with Mg con- to pure Al decreases the surface tension from 869 to
tent and foaming temperature 855 mN m21. With further addition of 0?5% Mg, this

910 Materials Science and Technology 2010 VOL 26 NO 8


Nadella et al. Foaming characteristics of Al–Si–Mg (LM25) alloy

5 a typical microstructures of LM25 foams exhibiting


change of eutectic Si morphology between the solid
a 670uC (F1); b 640uC (F4)
bottom and the originated foam. Magnified pictures
4 Effect of foaming temperature on cell morphology of
(inset) show the needle-like structure (left) in the solid
LM25 foam
bottom, and the fibrous Si network (right) in the foam.
b High magnification SEM image shows a part of that
value comes down to 850 mN m21. But adding 3%Mg to network within the cell wall
Al–8%Si brings down the surface tension to
y800 mN m21. Although the above composition differs
slightly from that of the alloy used in our investigations, better defined cells as illustrated by the magnified images
the trend seems clear. Also, as shown by Anson et al., of LM25 foam (Fig. 4). This is possibly due to higher
hydrogen atmosphere has no dramatic influence on the nucleation rate expected in view of faster dissociation of
surface tension of A356 alloy.8 Alloys with lower surface TiH2 creating greater supersaturation of hydrogen in
tension require smaller activation energy to form bubbles, liquid alloy. Further experimentation may be necessary
and, therefore, are expected to foam more easily. Thus, the to optimise the foaming temperature for minimising
lower surface tension of excess Mg containing alloys foam collapse and simultaneously obtaining fine and
(especially with 3%) must have resulted in easy bubble well defined cell structure.
nucleation. Although the effect of melt composition on
coalescence is not well documented, it is known that alloys Structure
with low surface tension are more vulnerable to cell Cell wall microstructures showed a regular network of
coalescence as demonstrated for Al foams.9 eutectic Si and the foam microstructures did not exhibit
With increasing Mg content, a higher volume fraction any noticeable variation with change in their Mg content
of oxides is expected to form during melt stirring cycle. and the foaming temperature. Close examination
At present it is not known whether these oxides are revealed a distinct difference in Si morphology with a
effective in stabilising aluminium foams, in the way well refined Si within the cell wall in contrast to the
alumina or Ca–Al oxides are. The large amount of needle-like Si in the solid bottom. This is illustrated in
collapse observed in the LM25z3%Mg alloy foamed at Fig. 5, which was taken in the transition point between
670uC (F-3) would indicate that the surface tension solid and foam. In the areas where the foaming has just
effect is more dominant than particle stabilisation. originated, one can see the morphological change in
For LM25z3%Mg alloy foamed at 640uC combina- eutectic Si that formed a network within the cell walls. A
tion of easy bubble nucleation and vulnerability to high magnification SEM image (Fig. 5b) clearly shows
bubble coalescence (both due to lower surface tension) the refined Si, which is in the range of 2–5 mm. In Al–Si
might be responsible for the cell size variation observed alloys, eutectic Si forms during solidification much below
(Fig. 2 and Table 1). For the base alloy, it was observed the coherency temperature and hence has no effect on the
that the higher foaming temperature gave finer and foam stability. The presence of Si network indicates that

Materials Science and Technology 2010 VOL 26 NO 8 911


Nadella et al. Foaming characteristics of Al–Si–Mg (LM25) alloy

that is the last solidified portion within the cell wall.


Calcium is known to induce chemical modification of
eutectic Si in Al–Si alloys and the work of Knuutinen et al.
showed even at the trace levels of 36 ppm, considerable
structural refinement of eutectic Si takes place.10 In the
present work, 0?08 wt-%Ca in the alloy helped to achieve
the Si modification. However, it may not be entirely due to
Ca, because Ca is supposed to be distributed throughout
even in the unfoamed solid portion.

Compression tests
Initial compression tests were carried out on one of the
good quality foams (F-1) for which the density is in the
range of 0?35–0?38 g cc21. Typical compression stress–
strain curve deduced from the load displacement data is
given in Fig. 6a. Consistent with any Al foam, the plot
exhibits three distinct regions, namely an initial elastic
portion, yield point with a long plateau region followed
by a region of rapid increase in stress corresponding to
foam densification.
Compressive strength is taken as the highest stress
(upper yield stress) immediately after the elastic defor-
mation. The drop in stress that followed is called lower
yield stress. Plateau stress is defined as the average stress
over a strain range of 0?1–0?6. Plateau strain ep is
computed where the plateau stress crossed the tangent
associated with densification region (Fig. 6a). Within the
limited data, it can be stated that the compression
strength and plateau stress seem to vary with the relative
density (Table 2). Within 2 mm deformation (0?07
strain), the load dropped by .50% indicating severe
damage during the initial straining of the sample.
During the course of deformation, the plateau region
exhibited serrations (inset in Fig. 6a) and visual
observations showed the breakage of cell walls into
small pieces, which indicates that these foams are not
ductile. A partially compressed (0?1 strain) specimen
exemplifies this point (Fig. 6b). It can also be seen that
the deformation is entirely confined to small portion of
the sample (in this case, the bottom most region,
Fig. 6b). This behaviour is much similar to Al/SiCP
composite foams, as observed by Ruan et al.,11 where
progressive damage and densification of small areas
characterised the overall compression. The brittle nature a typical compression stress–strain curve showing pla-
of LM25 foams may be due to the presence of Si teau region followed by densification (inset: enlarged
particles and in particular to its network distribution portion of plateau region exhibiting serrations); b par-
within the cell wall (Fig. 5a). tially compressed sample (y0?1 strain) with broken foam
pieces
6 Compression tests on LM25 alloy foam
Conclusions
LM25 alloy foams were successfully prepared with Al tendency at 670uC. Overall, higher foaming ratios were
foam turnings additions and TiH2. Foaming temperature obtained with 640uC as the foaming temperature. A
selection significantly influenced the foam decay with change in Si morphology from unmodified needles to well
higher foaming temperature of 670uC leading to severe modified fibrous network is noticed in the transition
foam collapse. On the other hand, foam decay is hardly region where the foaming has just initiated. Overall, these
noticed at the foaming temperature of 640uC. Presence of experiments exhibited the potential to obtain good quality
extra Mg (3 wt-%) in this alloy aggravated the collapsing LM25 foam using cheaper Al foam turnings.
Table 2 Details of compression tests performed at strain rate of 561022 s21

Relative density (r/rs)* Compression strength, MPa Lower yield stress, MPa Plateau stress, MPa Plateau strain

0.149 4.9 2.3 4.1 0.82


0.142 4.5 1.1 3.7 0.78
0.130 4.3 1.0 …{ …{
*rs is the solid density of LM25 which is 2685 kg m23.
{Test discontinued at 0?1 strain.

912 Materials Science and Technology 2010 VOL 26 NO 8


Nadella et al. Foaming characteristics of Al–Si–Mg (LM25) alloy

Uniaxial compression tests in these foams showed 3. D. H. Yang, B. Y. Hur, D. P. He and S. R. Yang: Mater. Sci. Eng.
similar behaviour as that of Al–Ca foam, but with A, 2007, A445–446, 415–426.
4. T. Miyoshi, M. Itoh, S. Akiyama and A. Kitahara: Adv. Eng.
localised damage induced during the initial phase of Mater., 2000, 2, 179–183.
deformation. A fine Si network within the cell walls may 5. D. E. J. Talbot: ‘The effects of hydrogen in aluminium and its
be responsible for the observed behaviour. alloys’, 141; 2004, London, Maney Publishing.
6. F. C. R. Hernandez, M. B. Djurdjevic, W. T. Kierkus and
Acknowledgements J. H. Sokolowski: Mater. Sci. Eng. A, 2005, A396, 271–
276.
The authors sincerely thank the Defence Research and 7. J. Goicoechea, C. Garcia-Cordovilla, E. Louis and A. Pamies:
J. Mater. Sci., 1992, 27, 5247–5252.
Development Organisation for their financial assistance 8. J. P. Anson, R. A. L. Drew and J. E. Gruzleski: Metall. Mater.
and the Director, Defence Metallurgical Research Trans. B, 1999, 30B, 1027–1032
Laboratory, for his encouragement. 9. T. Miyoshi, T. Kasai, T. Mukai and K. Higashi: Proc. Conf. on
‘Cellular metals and metal foaming technology’, Bremen,
References Germany, June 2001, Verlag MIT, 167.
10. A. Knuutinen, K. Nogita, S. D. McDonald and A. K. Dahle:
1. C. C. Yang and H. Nakae: J. Alloys Compd, 2000, 313, 188–191. J. Light Met., 2001, 1, 229–240.
2. C. C. Yang and H. Nakae: J. Mater. Proc. Technol., 2003, 141, 11. D. Ruan, G. Lu, F. L. Chen and E. Siores: Compos. Struct., 2002,
202–206. 57, 331–336.

Materials Science and Technology 2010 VOL 26 NO 8 913

You might also like