Sunteți pe pagina 1din 5

Scripta Materialia 54 (2006) 789–793

www.actamat-journals.com

Rheo-diecasting of Al–Si–Pb immiscible alloys


X. Fang *, Z. Fan
Brunel Centre for Advanced Solidification Technology (BCAST), Brunel University, Kingston Lane, Uxbridge UB8 3PH, UK

Received 15 July 2005; received in revised form 14 November 2005; accepted 15 November 2005
Available online 5 December 2005

Abstract

Immiscible Al–Si–Pb alloys are successfully produced using the newly developed rheo-diecasting process (RDC). The microstructure
of RDC Al–Si–Pb alloys is characterised by spherical and fine Pb particles dispersed uniformly in the Al–Si alloy matrix. The discussion
focuses on the mechanism of the RDC process for immiscible alloys.
 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Rheo-diecasting; Immiscible system; Semisolid processing; Aluminium alloys

1. Introduction for bearing applications. Al-based alloys have the advan-


tage of high strength to weight ratio, and therefore they
It has been commonly accepted that it is desirable that a are widely used in the automotive industry.
bearing material should have the following characteristics Immiscible alloys, such as Al–Pb, Al–In and Al–Bi
[1]: alloys, cannot be cast using conventional casting methods
due to the large density difference between the two phases
• Embeddability, to allow hard particles to embed in the in their liquid state [2,3]. Among these three systems, the
bearing material. most difficult system to cast conventionally and obtain a
• Compatibility, to avoid damaging or roughening the uniform distribution of the soft phase through the hard
counterface. phase, is the Al–Pb system. This is because the density
• Conformability, to accommodate slight misalignment in of Pb (11.3 g cm3) is much higher than that of In
the bearing system. (7.3 g cm3) or Bi (9.8 g cm3). As a result, Al–Pb alloys
• High strength, to support applied loads. are chosen as model materials for processing immiscible
• Fatigue resistance and corrosion resistance for a longer systems; if homogenous dispersion of Pb in the Al matrix
service life. in the solid state is achieved then it is reasonable to assume
it is possible to obtain uniform dispersion of In or Bi in an
To compromise and satisfy these conflicting demands, most Al matrix. Several methods have been proposed to produce
bearing materials contain two phases. One is a soft and immiscible Al–Pb alloys, such as stir casting [4], rheocast-
ductile phase, which provides good embeddability and con- ing [5], rapid solidification [6] and spray forming [7], but
formability; the other is a hard and strong phase, which has the microstructures produced by these methods are still
strength and fatigue resistance. Traditionally, soft metal not homogeneous. The rheomixing process using a twin-
elements such as Pb, Sn, In and Bi are incorporated in screw mechanism is a new approach, developed by Fan
harder metal elements such as Cu and Al to produce alloys et al. [8]. Immiscible Zn–Pb alloys have been successfully
mixed and extruded using this technique [9]. In this paper,
we investigate the microstructure and mechanical proper-
*
Corresponding author. Tel.: +44 1895 266418; fax: +44 1895 269758. ties of Al–Si–Pb alloys produced by a rheo-diecasting
E-mail address: mtsrxxf@brunel.ac.uk (X. Fang). (RDC) process developed recently at Brunel University.

1359-6462/$ - see front matter  2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.scriptamat.2005.11.021
790 X. Fang, Z. Fan / Scripta Materialia 54 (2006) 789–793

2. Experimental procedures average size of the Pb particles in the RDC Al–Si–3.8Pb


alloy (Fig. 1(a)) was 2.6 lm and the shape factor of the
The RDC equipment used in this work consists of a Pb particles was 0.89. Increasing the Pb concentration to
twin-screw slurry maker and a standard high-pressure 7.2 wt.% (Fig. 1(b)) increased the average size of Pb parti-
die-casting (HPDC) machine. The specially designed pro- cles to 11 lm and decreased their average shape factor to
files of the screws are closely intermeshing, self-wiping 0.86. Further increasing of the Pb concentration to
and co-rotating. The fluid flow inside the slurry maker is 17.5 wt.% (Fig. 1(c)), increased the average size of the Pb
characterised by high shear rate and high intensity of particles to 14 lm and decreased the shape factor to 0.82.
turbulence. The details of the slurry maker have been There is no significant segregation through the cross
described in Ref. [10]. section of the tensile test samples. The microstructure of
Al–Si–Pb alloy was prepared by melting a standard Al–Si matrix is characterised by spherical and uniformly
commercial A357 alloy with appropriate addition of pure distributed a-Al primary particles, very fine a-Al particles
Pb (the purity of Pb is 99.97 wt.%). The A357 alloy used produced during secondary solidification and very fine
has a composition Al–7Si–0.6Mg (in wt.%). The alloy eutectic. The average size of the a-Al primary phase with
was melted at 750 C, premixed at 620 C using a propeller different Pb concentrations varies from 49 to 55 lm, and
and processed at temperatures between 595 C and 605 C. the volume fraction of primary phase depends on the
The multi-phase mixture was sheared for 120 s, with a fixed mixing temperature. The alloys Al–7Si–3.8Pb and Al–
screw rotation speed of 500 rpm before it was transferred 7Si–7.2Pb were mixed at 605 C; therefore, the volume
to a DCC280 HPDC machine (LK Machinery, Hong fractions of a-Al primary phase were 15.3% and 14.5%,
Kong) with a 280-ton clamping force, to produce standard respectively. When the mixing temperature decreased to
tensile test samples. The die used for casting the test sam- 595 C, the volume fraction of a-Al primary phase in the
ples has six cavities, of which four are tensile test samples RDC Al–7Si–17.2Pb alloy increased to 37.6%. The shape
and two are Charpy test samples. The dimensions of the factors of a-Al primary particles for all of the RDC Al–
tensile test samples are 6 mm in gauge diameter, 60 mm Si–Pb samples had no significant difference. The higher
in gauge length and 150 mm in total length. The tempera- magnification micrograph (Fig. 2) shows that the second-
ture of the die was kept at 220 C during RDC. The ary solidification a-Al phase in all of the samples is fine
RDC process was optimised against the processing param- (<10 lm); this is attributed to the high cooling rate in the
eters and die casting conditions. The tensile test samples die.
produced for mechanical property testing were processed Two liquid phases, Al-rich phase and Pb-rich phase, co-
under these optimised conditions. exist at temperatures ranging from the immiscible gap to
Samples for metallographic examination were cut from the monotectic temperature (TM) of Al–Si–Pb alloys. These
the cross section of the tensile test samples, ground and two liquid phases segregate quickly during solidification
polished down to 1 lm. The microstructures were exam- due to two factors, one is Stokes motion (US) caused by
ined using conventional optical microscopy. The average the density difference between two liquids, and the other
equivalent diameter is used to describe the size of Pb parti- is Marangoni motion (UM) resulting from a temperature
cles and a-Al primary particles. The volume fraction, shape gradient, i.e. the liquid droplet moves to a higher tempera-
factor and particle size for both Pb particles and a-Al ture region in order to reduce interfacial energy. Thus, it is
primary particles were determined using a ZEISS KS300 impossible to cast immiscible alloys using conventional
image analysis system. The mechanical properties were techniques. We developed a two-step strategy for mixing
carried out using a universal materials testing machine immiscible alloys. The two-stage process, for achieving a
(Instron 5801) at a crosshead speed of 2 mm/min. uniform distribution of Pb phase in an Al–Si alloy matrix,
The equivalent diameter d and shape factor f are is schematically illustrated in Fig. 3. The first stage of the
described by the following formulas: process creates a coarse Pb liquid dispersion by premixing
pffiffiffiffiffiffiffiffiffiffiffi (Fig. 3(a)). The twin-screw slurry maker does not have
d ¼ 4A=p; ð1Þ
distributive mixing action, so the premixing is required to
f ¼ 4pA=P 2 ; ð2Þ achieve coarse uniform distribution of Pb droplets in the
A357 alloy melt matrix. In order to achieve uniform com-
where A and P are the area and perimeter of a particle,
position, the premixing started from 750 C and cooled
respectively.
down to 620 C, which is just above the liquidus of the
melt. The slurry maker was running whilst the pre-mixture
3. Results and discussion of Al–Si–Pb melt was introduced. Feeding the pre-mixture
into the slurry maker in this way guarantees the melt is
Fig. 1 shows the microstructures of the RDC Al–Si–Pb immediately under a shear action and this avoids the
samples with different additions of Pb. The Pb particles possibility of segregation.
(dark phase) in all of the samples are dispersed uniformly The second stage of the process generates a fine liquid
in the Al matrix. The resultant microstructures of Pb par- dispersion through high shear mixing and stabilises the
ticles and a-Al primary phase are listed in Table 1. The fine liquid dispersion by creating a semisolid slurry
X. Fang, Z. Fan / Scripta Materialia 54 (2006) 789–793 791

Fig. 1. Optical micrographs of Al–Si–Pb alloys produced by the RDC process: (a) 3.8 wt.% Pb sheared at 605 C with a shear rate of 534 s1 for 120 s;
(b) 7.2 wt.% Pb sheared at 605 C with a shear rate of 534 s1 for 120 s; (c) 17.2 wt.% Pb sheared at 595 C with a shear rate of 534 s1 for 120 s.

Table 1
The resultant microstructure of the RDC Al–Si–Pb alloys LAl

Sample Pb particles a-Al particles SAl


d (lm) f VF (%) d (lm) f LPb
Al–7Si–3.8Pb 2.6 0.89 15.3 55 0.72
Al–7Si–7.2Pb 11 0.86 14.5 49 0.70
Al–7Si–17.2Pb 14 0.82 37.6 52 0.68
VF—volume fraction; d—particle size and f—shape factor.
(a) (b)

Fig. 3. Schematic illustration of the two-stage process for achieving a


uniform distribution of soft phase in an Al–Si alloy matrix: (a) premixing;
(b) rheomixing.

simultaneously (Fig. 3(b)). For binary systems of immiscible


alloys, solid particles are only obtained when the mixture of
two liquids is cooled down to a temperature between the
monotectic temperature and the eutectic temperature. How-
ever, for the ternary system Al–Si–Pb, primary a-Al particles
can be achieved when the melt temperature drops below its
liquidus. At a mixing temperature below the liquidus, the
high heat transfer capability of the slurry maker results in
the quick generation of a semisolid slurry. The viscosity of
the semisolid slurry increases exponentially with the volume
fraction of solid particles [11]. The effects of the viscosity of
Fig. 2. Optical micrograph with higher magnification to show the the semisolid slurry on Us and UM can be described by Eqs.
microstructure of the matrix of the RDC Al–Si–Pb alloy. (3) [12] and (4) [13], respectively, as follows:
792 X. Fang, Z. Fan / Scripta Materialia 54 (2006) 789–793

2gDqðg þ g0 Þ 2 350 20
Us ¼ r; ð3Þ
3gð2g þ 3g0 Þ UTS
  dr  300
2dT  j
dx dT
TE
UM ¼ r; ð4Þ 15
ð2g þ 3g0 Þð2j þ j0 Þ 250

UTS (MPa)

Elongation (%)
where Dq is the density difference between the two liquids, 200
g is gravitational acceleration and r is the size of the liquid 10
droplet. dT/dx is the temperature gradient and dr/dT is the 150
variation of the interfacial energy between the Al and Pb
100
liquids with change in temperature. 5
In Eqs. (3) and (4), the viscosities of the liquid matrix (g) 50
and liquid droplets (g 0 ) affect both Us and UM, while the
conductivities of the liquid matrix (j) and the liquid droplet 0 0
0 5 10 15 20
(j 0 ) only affect UM. Eqs. (3) and (4) indicate that both Us
Lead concentration (wt.%)
and UM are inversely proportional to the viscosity of the
semisolid slurry. When primary particles form, the viscosity Fig. 4. Mechanical properties of the RDC Al–Si–Pb alloys as a function
of the melt in the slurry maker greatly increases. Therefore, of Pb concentration.
both Us and UM are slowed down due to the increase of the
meltÕs viscosity. The lower the volume fraction of the Pb
phase, the larger the distance between Pb droplets, if the using the mixing temperature to vary the volume fraction
sizes of Pb droplets are similar. Pb droplets need to travel of the a-Al primary particles. The mixture of Pb and Al–
further to merge into a larger droplet. Thus, a lower volume Si semisolid slurry, produced by the slurry maker, is very
fraction of a-Al primary is required for a lower Pb content stable. The subsequent die-casting can greatly speed up
to avoid the collision of Pb droplets, compared to the Al– the solidification of the semisolid slurry, depending on
Si–Pb alloy with higher levels of Pb. A volume fraction of the temperature of the die. Additionally, the turbulence
15% a-Al primary phase is enough for Al–Si–Pb alloys with and shear generated when the semisolid slurry passes
a Pb concentration lower than 7.2 wt.%, to obtain uniform through the gate of the die can further prevent segregation.
dispersion of Pb particles. However, when the Pb concen- Therefore, immiscible Al–Si–Pb alloys can be produced
tration increases to 17.2 wt.%, the volume fraction of a-Al using the RDC process and the resulting microstructure
primary phase has to be increased to 37.6% in order to has fine Pb particles uniformly dispersed in an Al–Si
avoid Pb phase segregation. matrix.
Apart from the viscosity for stabilisation of the melt, the The morphology of the a-Al primary particles is spher-
size of the Pb droplets also slows down the Us and UM ical and fine (Table 1), as a result the semisolid slurry of the
motions. Although the size of the Pb droplets is quite Al–Si–Pb alloy has a much better fluidity than those with a
coarse after premixing, they will be reduced further to a coarse dendritic a-Al phase. The benefit of this superior
finer size under the high shear rate and intensive turbulence fluidity of the semisolid slurry is to reduce segregation of
offered by the twin-screw slurry maker. According to the the Pb phase and a-Al primary particles during diecasting.
theoretical analysis [14], the droplet deformation is a result Fig. 4 shows that when the Pb concentration is less than
of balancing the interfacial tension (tending to keep the 10 wt.%, both the ultimate tensile strength and uniform
droplet spherical), with the viscous force (tending to elongation of Al–Si–Pb alloys have no significant differ-
elongate the droplet). When the interfacial tension can no ence. When the Pb concentration increases to 17.2 wt.%,
longer balance the viscous force, the deformation becomes the ultimate tensile strength and uniform elongation
unstable and the droplet will burst. Therefore, under the decrease by about 20% and 40%, respectively.
high shear rate and intensive turbulence, the coarse Pb The strength of the Pb phase is much lower than that of
droplets are elongated, broken up into smaller droplets Al–Si matrix; meanwhile the ductility is much higher. The
and dispersed uniformly in the semisolid slurry matrix. dispersed Pb particles act as a type of inclusion in the Al–Si
According to Eqs. (3) and (4), Us and UM are proportional matrix; therefore, the smaller the Pb particles, the less the
to r2 and r of the liquid Pb droplets, respectively. Conse- effect they have on the tensile strength and elongation of
quently, smaller sizes of the Pb droplets greatly slow down the Al–Si–Pb alloy. That is why the ultimate tensile
both the Marangoni motion and Stokes motion. No shear strength and elongation of the RDC Al–Si–Pb alloy
is exerted on the melt between releasing the melt from the decreases slowly with the Pb concentration.
slurry maker and transferring to the high-pressure die-cast-
ing machine. If the size of the Pb droplets is either not fine
enough or the viscosity of semisolid slurry is not high 4. Concluding remarks
enough, segregation will occur. However, using this two-
stage process, the size of the Pb droplets is very small In the present work, rheo-diecasting technology has
and the viscosity of semisolid slurry can be controlled by been used to process immiscible Al–Si–Pb systems. The
X. Fang, Z. Fan / Scripta Materialia 54 (2006) 789–793 793

RDC process produces a uniform dispersion of Pb particles References


in an Al-alloy matrix. The size of the a-Al primary phase is
approximately 50 lm and the average size of the Pb parti- [1] Kingsbury GR. In: Friction and wear of sliding bearing materials.
cles increases with the increase in Pb concentration. The Pb 10th ed. In: Olson D, editor. ASM handbook, vol. 18. Materials Park
(OH): ASM International; 1992. p. 743.
particle size increased from 2.6 lm in Al–7Si–3.8Pb to [2] Singh RN, Sommer F. Rep Prog Phys 1997;60:57.
14 lm in Al–7Si–17.2Pb alloy. The Pb particles were dis- [3] Ratke L, Diefenbach S. Mater Sci Eng 1995;R15:263.
persed uniformly throughout the cross section of the tensile [4] Pathak JP, Tiwari SN, Malhotra SL. Met Technol 1979;6:442.
test piece. The stabilisation of the liquid Pb droplets in the [5] Agarwala V, Satyanarayana KG, Agarwala RC, Garg R. Mater Sci
multi-phase mixture is achieved by two sequential mecha- Eng 2002;A327:186.
[6] Sheng HW, Zhou F, Hu ZQ, Lu K. J Mater Res 1998;13:308.
nisms: (a) increased viscosity by the presence of primary [7] Zhang Y, Hao Y, Liu Z, Chen G. Mater Sci Eng 1988;98:119.
a-Al particles; (b) fine and uniformly dispersed Pb particles [8] Fan Z, Bevis MJ, Ji S. PCT Patent, 1999, WO 01/23124 A1.
in the Al–Si semisolid matrix. It is also found that both the [9] Fang X, Fan Z. Mater Sci Technol 2005;21:366.
ultimate tensile strength and elongation of Al–Si–Pb alloys [10] Fang X, Ji S, Fan Z. In: Proceedings of the 8th international
conference on semisolid metal processing. Limassol Cyprus, 2004. p.
decrease with increasing Pb content.
02–3.
[11] Joly PR, Mehrabian R. J Mater Sci 1976;11:1393.
Acknowledgement [12] Levich VG. Physicochemical hydrodynamics. Englewood Cliffs
(NJ): Prentice-Hall; 1962.
The financial support from the EPSRC is gratefully [13] Young N, Goldstein J, Block M. J Fluid Mech 1959;6:350.
acknowledged. [14] Tang H, Wrobel LC, Fan Z. Mater Sci Eng 2004;365A:325.

S-ar putea să vă placă și