Sunteți pe pagina 1din 7

Materials Chemistry and Physics 67 (2001) 8591

Liquid phase sintering of aluminium alloys


G.B. Schaffer , T.B. Sercombe1 , R.N. Lumley2
Department of Mining, Minerals and Materials Engineering, The University of Queensland, Brisbane, Qld 4072, Australia

Abstract The principle that alloys are designed to accommodate the manufacture of goods made from them as much as the properties required of them in service has not been widely applied to pressed and sintered P/M aluminium alloys. Most commercial alloys made from mixed elemental blends are identical to standard wrought alloys. Alternatively, alloys can be designed systematically using the phase diagram characteristics of ideal liquid phase sintering systems. This requires consideration of the solubilities of the alloying elements in aluminium, the melting points of the elements, the eutectics they form with aluminium and the nature of the liquid phase. The relative diffusivities are also important. Here we show that AlSn, which closely follows these ideal characteristics, has a much stronger sintering response than either AlCu or AlZn, both of which have at least one non-ideal characteristic. 2001 Elsevier Science B.V. All rights reserved.
Keywords: A-metals; B-sintering

1. Introduction Powder metallurgy (P/M) can be used to make high strength and high stiffness aluminium alloys [110]. Indeed, room temperature tensile strengths in excess of 800 MPa have been reported [11], which is approaching the theoretical limit for aluminium [12]. However, these alloys are not produced to near net shape and are therefore expensive to fabricate, which limits their use to niche applications in the aerospace industry. Conventional press-and-sinter P/M is an exemplary net shape process and therefore offers inexpensive manufacturing. The commercially available alloys are based on research that was done in the late 1960s to early 1970s [1316] and there has only been sporadic activity in the eld since then [1723]. This early work concentrated on wrought alloy compositions and the alloys were not designed to be sintered. Because sintering is the step in the P/M process that is most responsible for the development of strength and other properties, it is not surprising that current commercial alloys do not meet the requirements of many load bearing applications for which they may otherwise be suitable. Pressed-and-sintered alloys therefore require substantial improvement before widespread use is likely. This
Corresponding author. Tel.: +61-7-3365-4500; fax: +61-7-3365-3888. E-mail address: g.schaffer@minmet.uq.edu.au (G.B. Schaffer). 1 Present address: IRC in Materials, The University of Birmingham, Edgbaston B15 2TT, UK. 2 Present address: CSIRO Manufacturing Science and Technology, Private Bag 33, Clayton, South MDC, Vic. 3169, Australia.

paper reviews recent work at The University of Queensland in which the traditional compositional restraints have been relaxed. We begin, however, with a discussion of the ubiquitous oxide lm.

2. The surface oxide Aluminium is always covered by an oxide. The thickness of the oxide is dependant on the temperature at which it formed and the atmosphere in which it is stored, particularly the humidity. Fresh oxide on bulk aluminium at room temperature is widely reported as being 1020 thick [2427]. The thickness on atomised powder can vary from 50 to 150 [2831]. The oxide on aluminium is usually amorphous [28,31,32] and hydrated [27,31,33,34] with an adsorbed water layer [35,36]. The oxide crystallises to -Al2 O3 on prolonged annealing at temperatures above 350 C [32,37,38]. Similar transformations occur in bulk alumina [39]. The oxide prevents solid state sintering in low melting point metals [40], including aluminium [41], but not in all metals [4244]. This has been explained in terms of the relative diffusion rates through the oxide and the metal, for metals with stable oxides [4547]. The use of liquid phases is an alternative to solid state sintering. An essential requirement for effective liquid phase sintering is a wetting liquid [48]. The wettability of a solid by a liquid is determined by the work of adhesion, Wa , [49,50]: Wa = lv (1 + cos ) = sv + lv sl (1)

0254-0584/01/$ see front matter 2001 Elsevier Science B.V. All rights reserved. PII: S 0 2 5 4 - 0 5 8 4 ( 0 0 ) 0 0 4 2 4 - 7

86

G.B. Schaffer et al. / Materials Chemistry and Physics 67 (2001) 8591

where lv is the surface tension of the liquidvapour interface, sv the surface tension of the solidvapour interface, sl the solidliquid interfacial tension and the contact angle. A liquid is said to wet a solid when cos > 0. High melting point metal oxides are generally poorly wetted by liquid metals, except above the wetting threshold, a temperature beyond which Wa increases sharply [50]. Liquid aluminium is not therefore expected to wet alumina near the melting point of the metal. Indeed, the contact angle is variously given as 103 at 900 C [51], 160 at 800 C [52] or 162 at 950 C [53], although this is dependant on the partial pressure of oxygen and the presence of an oxide lm on the molten metal [54]. It has been suggested that the AlCuAl2 eutectic can wet Al2 O3 at 600 C [19]. However, neither Mg, Ce nor Ca additions to molten Al reduce the contact angle sufciently to produce wetting [51,52]. Since the work of adhesion of liquid metals on oxide surfaces increases with the free energy of formation of the metal oxide, it is unlikely that Cu will be efcacious. It is therefore apparent that the oxide on aluminium is a barrier to sintering and needs to be disrupted or otherwise removed. The oxidation of a metal, M, may be represented as M + O2 MO2 The free energy of formation, G = RT ln K1 (2) G, of the oxide is given by (3)

where R is the gas constant, T the temperature in kelvin and K1 the equilibrium constant given by K1 = (PO2 )1 (4) where PO2 is the partial pressure of oxygen when reaction (1) is at equilibrium. For aluminium at 600 C, a PO2 < 1050 atm is required to reduce the oxide [55]. Atmospheres containing hydrogen are often used in powder metallurgy. Hydrogen can reduce a metal oxide by the reaction: MO + H2 M + H2 O PH2 O K5 = PH2 (5)

which is a partial reduction reaction. This reaction is observed in studies of the oxidation behaviour of AlMg alloys [57,58] and at bonding interfaces in metal matrix composites [5963]. Detailed analytical transmission electron microscopy (Fig. 1) indicates that spinel crystallites are indeed present in a sintered AlMg alloy. The reaction may be facilitated during sintering by diffusion of the magnesium through the aluminium matrix and will be accompanied by a change in volume, creating shear stresses in the lm, ultimately leading to its break up. This will propitiate diffusion, wetting and therefore sintering. It has been shown that the sintering of aluminium is enhanced in the presence of magnesium [23,64,65]. More recently, X-ray photoelectron spectroscopy indicated that the surface oxide can be reduced in the presence of magnesium, which exposes fresh metal and facilitates the subsequent formation of AlN on exposed surfaces in a nitrogen atmosphere [66]. The effect that magnesium has on sintering can be shown by dilatometry (Fig. 2). An addition of >0.15% Mg causes shrinkage. The microstructures of the AlSn system show that liquid tin only wets aluminium in the presence of magnesium, when the dihedral angle is very sharp. Without magnesium, the dihedral angle is obtuse and the liquid is exuded during sintering (Fig. 3). By promoting sintering, magnesium also affects the mechanical properties (Fig. 4). The large increase in strength and ductility at 0.15% Mg is a direct consequence of improved interparticle bonding and densication following oxide rupture. The excess magnesium at concentrations >0.15% remains in solution in the aluminium, causing expansion by the Kirkendall effect and solid solution hardening. It is apparent that the oxide is not the barrier to the sintering of aluminium that it is traditionally considered to be. Other factors must therefore be the cause of the poor sintering response.

3. Alloy design Alloys are generally designed to accommodate the manufacture of goods made from them as much as the properties required of them in service. It is for this reason that sintered steels, for example, often contain copper or phosphorous in addition to carbon and nickel. Similarly, cast aluminium alloys are different to forging alloys which are different again to extrusion alloys. However, with one exception [67], this principle does not appear to have been applied to pressed and sintered P/M aluminium, although Savitskii only examined binary alloys, the oxide phase was not reduced and no allowance was made in the thermal cycle for the transient nature of the sintering liquids. The compositions of the current commercial alloys are compared to standard wrought material in Table 1. It is noteworthy that the compositions of the P/M alloys are essentially identical to those of the wrought material. It is therefore not surprising that their sintering response is poor.

The equilibrium constant for this reaction, K5 , is given by (6)

where PH2 and PH2 O are the partial pressure of hydrogen and water vapour, respectively. The ratio of partial pressures can be converted to the dew point, effectively the water vapour content. A dew point of 140 C at 600 C is required to reduce Al2 O3 [56]. Neither a dew point of 140 C nor a PO2 of 1050 atm is physically attainable and therefore aluminium cannot be sintered in conventional atmospheres. Magnesium is highly reactive and the free energy of formation of its oxide is more negative than that of the oxides of aluminium. Magnesium therefore has the potential to act as a solid reducing agent in this system. A possible reaction is 3Mg + 4Al2 O3 3MgAl2 O4 + 2Al (7)

G.B. Schaffer et al. / Materials Chemistry and Physics 67 (2001) 8591

87

Fig. 1. (a) Transmission electron micrograph of a sintered Al2.5%Mg alloy, showing a multitude of spinel crystallites. The inset shows the selected area diffraction pattern from this region; it can be indexed to spinel. (b) EDS spectra from (a) showing that the ne crystallites contain signicantly more magnesium and oxygen than does the aluminium matrix (c) [80].

Fig. 2. Dilatometry curves for AlxMg alloys, where x is 0, 0.15 and 1.5 wt.% Mg showing the effect of trace additions of magnesium to aluminium cause shrinkage during sintering [80].

Fig. 3. Exuded liquid on the surface of an Al8Sn alloy after sintering at 620 C.

88

G.B. Schaffer et al. / Materials Chemistry and Physics 67 (2001) 8591

Fig. 4. As sintered tensile properties for AlMg alloys sintered 30 min at 620 C [80].

Instead of producing sintered alloys which simply mimic existing wrought alloys, it is preferable to develop alloys that are specically designed to be sintered. Based on an understanding of fundamental liquid phase sintering phenomena, German and co-workers [68,69] recognised that it is possible to dene certain ideal phase diagram characteristics. The key features of an ideal liquid phase sintering system are as follows: The additive should have a lower melting point than the base. The alternative is a low melting point eutectic which is less advantageous because liquid formation does not occur spontaneously on heating. The solubility of the additive in the base should be low because this ensures that the additive remains segregated to particle boundaries and maximises the liquid volume. While the base should be soluble in the liquid, it is not necessary for the base to be soluble in the solid additive. Completely miscible liquids ensures that mass transport is not constrained. In addition, the base should also have a high diffusivity in the liquid. This ensures high rates of mass transport and therefore rapid sintering. 3.1. The AlCu system Copper is one of the primary alloying elements for aluminium, based largely on the substantial age hardening response of AlCu alloys. They are arguably the most widely studied P/M alloys [15,1922,70,71]; they are certainly the most widely used. The binary phase diagram is shown in
Table 1 The composition and properties of sintered aluminium alloys and the equivalent wrought alloys [79] Alloy Type Composition Cu 6061 601 2014 201 Wrought P/M Wrought P/M 0.3 0.25 4.4 4.4 Si 0.6 0.6 0.8 0.8 Mg 1 1 0.5 0.5 100 94 100 93 Density (%) T6 properties UTS (MPa) f (%) 310 232 483 323 12 2 13 0.5

Fig. 5. It has two of the ideal features: there is a single liquid phase in which aluminium is continuously soluble and the maximum solid solubility of copper in aluminium is 5.65% at 548.2 C. However, the melting point of copper is almost double that of aluminium. The liquid phase forms as a eutectic between (Al) and Al2 Cu, this is shown in Fig. 6. Because there is some solid solubility of copper in aluminium, the liquid is partially transient. The sequence of events during sintering of AlCu is: interdiffusion takes place on heating from room temperature and a series of AlCu intermetallics form; the rst liquid forms at 548 C on AlAl2 Cu ( ) boundaries; Cu is drawn from the liquid into solution in the aluminium and is replaced by dissolution of the intermetallics, which are replenished in turn by solid state diffusion from adjacent Cu particles; the intermetallics disappear when all the Cu is completely dissolved; all the liquid is absorbed into the Al particles if the Cu content is low, although most alloys retain some liquid throughout sintering.

Fig. 5. The AlCu phase diagram [81].

G.B. Schaffer et al. / Materials Chemistry and Physics 67 (2001) 8591

89

Fig. 7. The AlZn phase diagram [81].

Fig. 6. Optical micrograph of an Al5.5Cu alloy quenched from 575 C showing the eutectic liquid forming between the Al matrix and the Al2 Cu phase.

The major problem in this system is that the diffusivity of copper in aluminium is almost 5000 times faster than that of aluminium in copper. The diffusivity, D, of Cu in Al at 600 C is 5.01 109 cm2 s1 whereas the diffusivity of Al in Cu is 1.14 1012 cm2 s1 [72]. While the faster diffusivity of copper in aluminium enhances the rate of homogenisation, it causes expansion via the Kirkendall effect. Sintering of the AlCu system is therefore dependent on the process variables, particularly the copper particle size and the heating rate [73,74]. This is non-ideal. 3.2. The AlZn system The AlZn system (Fig. 7) shows some of the characteristics of an ideal liquid phase sintering system in that zinc has a lower melting point than aluminium, no intermediate phases form and there is complete miscibility in the liquid. However, the solid solubility ratio is non-ideal. The maximum solid solubility of zinc in aluminium is 83.1%, while the maximum solid solubility of aluminium in zinc is 1.2%. The liquid phase during sintering of Al with Zn is therefore highly transient and AlZn alloys are extremely process sensitive. Fast heating rates and coarse zinc particle sizes enhance sintering [73]. Where ne zinc particles are used, the zinc dissolves in the aluminium before substantial quantities of liquid phase can form. Where coarse zinc particles are used, the aluminium becomes locally saturated before

homogenisation is achieved. Hence the additive forms a liquid which aids sintering. High heating rates also favour liquid formation because the opportunity for diffusion to occur before melting is minimised and the reaction is delayed to higher temperatures where the equilibrium solubility is smaller and therefore local saturation can occur more easily. The 7000 alloys have the greatest response to age hardening of the conventional aluminium alloys and are therefore used as high strength forgings in the aerospace industry. Because zinc is a poor sintering aid, however, these alloys, which contain 38% Zn, do not have a good sintering response either. The high vapour pressure of zinc also gives rise to additional porosity in these alloys, particularly when elemental powders are used [75]. It is therefore necessary to use master alloy powders [76] or microalloying additions in order to achieve acceptable sintered properties [77]. 3.3. The AlSn system An examination of the binary aluminium phase diagrams indicates that AlSn is perhaps the only one which exhibits almost all of the features of an ideal system (Fig. 8). The melting point of tin (232 C) is considerably lower than that of aluminium (660 C) and there are no intermetallic phases. Tin is sparingly soluble in solid aluminium: the maximum solid solubility is <0.15%. Aluminium is completely soluble in liquid tin and no immiscible liquids form. In addition, the diffusivity of Al in liquid Sn is faster than the diffusivity of either Cu or Zn in liquid Sn and about ve times greater than the self diffusivity of liquid Sn [78]. In the presence of magnesium, tin is indeed a very effective sintering aid. This is illustrated in Fig. 9, which is a densication contour map for the AlSnMg system. The densication is a function of the green density, sintered density and theoretical density and is a measure of the sintering response: positive values indicate shrinkage, negative values indicate expansion; full density is achieved at a value of 1.

90

G.B. Schaffer et al. / Materials Chemistry and Physics 67 (2001) 8591

Fig. 8. The AlSn phase diagram [81]. Fig. 10. The macrostructure of a section of one tooth of a functionally graded, freeform fabricated gear. The centre is the Al8Sn4Mg alloy; the surface contains a 10 wt.% loading of alumina [82].

The closely spaced, parallel contour lines at low magnesium concentrations indicate that small quantities of magnesium are required to activate the system. The widely spaced, gently sloping contour lines at higher magnesium concentrations indicate that the system is relatively insensitive to Sn concentration and Mg levels greater than the critical concentration. At a tin concentration of 8%, the sintered density approaches 99% of theoretical. The AlSnMg system, having been designed to be sintered, is also effective for uncompacted powder, i.e. aluminium can be gravity sintered to near full density. This facilitates free form fabrication and rapid prototyping. By combining the exibility of free forming and the easy sintering of the AlSnMg system, functionally graded metal matrix composites can also be manufactured (Fig. 10). The mechanical properties of the AlSn system, however, are low because tin does not provide much strengthening.

Copper could be incorporated, but sintering of the quaternary system becomes complicated, partly because of the formation of immiscible liquid phases.

4. Conclusions Sintering of aluminium has always been considered to be problematical because of the oxide lm present on the surface of the powder particles. However, trace additions of magnesium react with the oxide to form spinel. This breaks up the oxide, which facilitates sintering. It is therefore apparent that in contradiction to the standard paradigm, the properties of pressed-and-sintered aluminium alloys are not limited by the oxide problem. Aluminium P/M alloys can be improved without recourse to hot working or master alloy powders if their design is based on an understanding of the underlying sintering processes and the characteristics of an ideal liquid phase sintering system. The AlSn system is close to ideal and AlSnMg alloys can be sintered to 99% of the theoretical density. Once an alloy has been optimised for sintering, a variety of new options can be realised. Free formed, functionally graded, aluminium matrix composites and high strength materials are two such examples.

Acknowledgements
Fig. 9. Map showing densication of sintered aluminium as a function of magnesium and tin concentration. Each contour represents a densication of 0.2.

This work has been supported in part by the Australian Research Council, ACL Bearing Company, Comalco Aluminium Ltd. and Ampal Inc.

G.B. Schaffer et al. / Materials Chemistry and Physics 67 (2001) 8591

91

References
[1] W.S. Cebulak, E.W. Johnson, H. Markus, Met. Eng. Quart. (1976) 37. [2] J.R. Pickens, J. Mater. Sci. 16 (1981) 1437. [3] P. Gilman, Met. Mater. (1990) 504. [4] E.J. Lavernia, J.D. Ayers, T.S. Srivatsan, Int. Mater. Rev. 37 (1992) 1. [5] H. Jones, in: H.M. Flower (Ed.), High Performance Materials in Aerospace, Chapman & Hall, London, 1995, pp. 340356. [6] I.A. Ibrahim, F.A. Mohamed, E.J. Lavernia, J. Mater. Sci. 26 (1991) 1137. [7] T.S. Srivatsan, I.A. Ibrahim, F.A. Mohamed, E.J. Lavernia, J. Mater. Sci. 26 (1991) 5965. [8] A.L. Geiger, J.A. Walker, JOM (1991) 8. [9] D.J. Lloyd, Int. Mater. Rev. 39 (1994) 1. [10] R. Bushby, Mater. World 5 (1997) 79. [11] J.Q. Guo, N.S. Kazama, Mater. Sci. Eng. A 232 (1997) 177. [12] E. Hornbogen, E.A. Starke, Acta Metall. Mater. 41 (1993) 1. [13] J.H. Dudas, W.A. Dean, Int. J. Powder Metall. 5 (1969) 21. [14] J.H. Dudas, C.B. Thompson, Mod. Dev. Powder Metall. 5 (1971) 19. [15] T. Watanabe, K. Yamada, Int. J. Powder Metall. 4 (1968) 37. [16] P.E. Matthews, Int. J. Powder Metall. 4 (1968) 39. [17] I. Amato, S. Corso, E. Sgambetterra, Powder Metall. (1976) 171. [18] R. Sundaresan, P. Ramakrishnan, Int. J. Powder Metall. Powder Technol. 14 (1978) 9. [19] W. Kehl, H.F. Fischmeister, Powder Metall. 23 (1980) 113. [20] W. Kehl, H.F. Fischmeister, Observations on dimensional changes during sintering of AlCu compacts, in: D. Kolar, S. Pejovnik, M.M. Ristic (Eds.), Sintering Theory and Practice, Proceedings of the Fifth International Round Table Conference on Sintering, Yugoslavia, Elsevier, Amsterdam, 1981. [21] W. Kehl, M. Bugajska, H.F. Fischmeister, Powder Metall. 26 (1983) 221. [22] F. Farzin-Nia, B.L. Davies, Powder Metall. 25 (1982) 209. [23] I.A. Shibli, D.E. Davies, Powder Metall. 30 (1987) 97. [24] H.J. Mathieu, M. Datta, D. Landolt, J. Vac. Sci. Technol. 3 (1985) 331. [25] I. Olefjord, A. Karlsson, Surface analysis of aluminium foil, in: T. Sheppard (Ed.), Aluminium Technology86, The Institute of Metals, London, 1986. [26] P. Marcus, C. Hinnen, I. Olefjord, Surf. Interf. Anal. 20 (1993) 923. [27] A. Nylund, I. Olefjord, Surf. Interf. Anal. 21 (1994) 290. [28] I.E. Anderson, J.C. Foley, J.F. Flumerfelt, Simplied aluminum powder metallurgy processing routes for automotive applications, in: W.F. Jandeska, R.A. Chernenkoff (Eds.), Powder Metallurgy Aluminum and Light Alloys for Automotive Applications, MPIF, Dearbon, MI, 1998. [29] A. Ozbilen, A. Unal, T. Sheppard, Inuence of oxygen on morphology and oxide content of gas atomised aluminium powders, in: W.M. Small (Ed.), Physical Chemistry of Powder Metals Production and Processing, The Minerals, Metals and Materials Society, 1989. [30] P. Nielsen, Y.L. Liu, N. Hansen, Manufacturing of aluminium composites with high purity matrix, in: 1993 Powder Metallurgy World Congress, Japan Society of Powder and Powder Metallurgy, Japan, 1993. [31] Y. Kim, W.M. Grifth, F.H. Froes, J. Met. 37 (1985) 27. [32] M.H. Jacobs, J. Microsc. 99 (1972) 165. [33] N.A. Thorne, P. Thuery, A. Frichet, P. Gimenez, A. Sartre, Surf. Interf. Anal. 16 (1990) 236. [34] I. Olefjord, A. Nylund, Surf. Interf. Anal. 21 (1994) 290. [35] J.C. Fuggle, L.M. Watson, D.J. Fabian, S. Affrossman, Surf. Sci. 49 (1975) 61. [36] J.F. Flumerfelt, I.E. Anderson, The surface chemistry of pure aluminum powders as measured with quadrupole mass spectrometry, in: 1998 International Conference on Powder Metallurgy and Particulate Materials, MPIF, Las Vegas, NV, 1998.

[37] K. Shinohara, T. Seo, H. Kyogoku, Zeitschrift fur Metallkunde 73 (1982) 774. [38] H.J. van Beek, E.J. Mittemeijer, Thin Solid Films 122 (1984) 131. [39] W.H. Gitzen, Alumina as a Ceramic Material, American Ceramic Society, Westerville, OH, Special Publication No. 4, 1970. [40] R.F. Smart, E.C. Ellwood, Nature 181 (1958) 833. [41] R.Q. Guo, P.K. Rohatgi, D. Nath, J. Mater. Sci. 32 (1997) 3971. [42] P. Ramakrishnan, G.S. Tendolkar, Powder Metall. 7 (1964) 34. [43] P. Ramakrishnan, Powder Metall. 9 (1966) 47. [44] T. Watanabe, Y. Horikosji, Int. J. Powder Metall. Powder Technol. 12 (1976) 209. [45] Z.A. Munir, J. Mater. Sci. 17 (1979) 2733. [46] Z.A. Munir, Powder Metall. 24 (1981) 177. [47] P.K. Higgins, Z.A. Munir, Powder Metall. Int. 14 (1982) 26. [48] W. Kingery, J. Appl. Phys. 30 (1959) 301. [49] F. Delannay, L. Froyen, A. Deruyttere, J. Mater. Sci. 22 (1987) 1. [50] J.V. Naidich, Prog. Surf. Memb. Sci. 14 (1981) 353. [51] S.W. Ip, M. Kucharski, J.M. Toguri, J. Mater. Sci. Lett. 12 (1993) 1699. [52] Y. Liu, Z. He, G. Dong, Q. Li, J. Mater. Sci. Lett. 11 (1992) 896. [53] D.T. Livey, P. Murray, Second Plansee Seminar, 1955. [54] J.-G. Li, Ceram. Int. 20 (1994) 391. [55] L.S. Darken, R.W. Gurry, Physical Chemistry of Metals, McGraw-Hill, New York, 1953, p. 349. [56] C. Lall, Int. J. Powder Metall. 27 (1991) 315. [57] G.M. Scamans, E.P. Butler, Metall. Trans. A 6A (1975) 2055. [58] D.H. Kim, E.P. Yoon, J.S. Kim, J. Mater. Sci. Lett. 15 (1996) 1429. [59] C.G. Levi, G.J. Abbaschian, R. Mehrabian, Metall. Trans. A 9A (1978) 697. [60] R. Molins, J.D. Bartout, Y. Bienvenu, Mater. Sci. Eng. A 135 (1991) 111. [61] J. Homeny, M.M. Buckley, Mater. Lett. 10 (1991) 421. [62] A. Munitz, M. Metzger, R. Mehrabian, Metall. Trans. A 10A (1979) 1491. [63] A.D. McLeod, C.M. Gabryel, Metall. Trans. A 23A (1992) 1279. [64] S. Miura, Y. Machida, Y. Hirose, R. Yoshimura, The effect of varied alloying methods on the properties of sintered aluminium alloys, in: 1993 Powder Metallurgy World Conference, Japan 1993. [65] K. Nishiyama, E. Suzuki, Processing and wear properties of P/M AlPb alloys, in: 1993 Powder Metallurgy World Congress, 1993. [66] A. Kimura, et al., Appl. Phys. Lett. 70 (1997) 3615. [67] A.P. Savitskii, Liquid Phase Sintering of the Systems with Interacting Components, Russian Academy of Sciences, Tomsk, 1993. [68] R.M. German, B.H. Rabin, Powder Metall. 28 (1985) 7. [69] R.M. German, Liquid Phase Sintering, Plenum Press, New York, 1985. [70] F.J. Esper, G. Leuze, Powder Metall. Int. 3 (1971) 123. [71] T. Watanabe, K. Yamada, Effects of Copper Adding Methods on the Strength of Sintered AluminiumCopper Alloys, Vol. 19, Castings Research Laboratory, Waseda University, 1968. [72] H. Mehrer, Landolt Bornstein Numerical Data and Functional Relationships in Science and Technology, Vol. III/26, Springer, Berlin, 1991. [73] R. Lumley, G.B. Schaffer, Scripta Mater. 35 (1996) 589. [74] R. Lumley, G.B. Schaffer, Scripta Mater. 39 (1998) 1089. [75] R.N. Lumley, G.B. Schaffer, Metall. Mater. Trans. A 30A (1999) 1682. [76] M. Muhlburger, P. Paschen, Zeitschrift fur Metallkunde 84 (1993) 346. [77] G.B. Schaffer, S.H. Huo, Powder Metall. 42 (1999) 219. [78] C.H. Ma, R.A. Swalin, Acta Metall. 8 (1960) 388. [79] Aluminum and Aluminum Alloys, ASM Specialty Handbook, ASM International, Materials Park, OH, 1993. [80] R. Lumley, T.B. Sercombe, G.B. Schaffer, Metall. Mater. Trans. A 30A (1999) 457. [81] Alloy Phase Diagrams, ASM Handbook, Vol. 3, ASM International, Materials Park, OH, 1992. [82] T.B. Sercombe, G.B. Schaffer, P. Calvert, J. Mater. Sci. 34 (1999) 4245.

S-ar putea să vă placă și