Elsevier

Wear

Volume 271, Issues 11–12, 2 September 2011, Pages 2845-2851
Wear

Wear behaviour of interpenetrating alumina–copper composites

https://doi.org/10.1016/j.wear.2011.05.042Get rights and content

Abstract

The wear behaviour of a variety of alumina–copper interpenetrating composites was tested as a function of copper ligament diameter and volume fraction of copper. The wear mechanisms of pure copper and pure alumina were adhesive and abrasive wear, respectively. In the composites with 1 μm, 5 μm and 15 μm copper ligament diameters, the wear mechanism was a mixture of adhesive and oxidative; in composites with a 30 μm copper ligament diameter a mixture of abrasive and oxidative. Increasing the amount of copper decreased hardness and thus increased wear, except where cyclic tribolayer behaviour occurred or where the alumina grains were weakly bonded. Increasing the copper ligament diameter decreased wear, although this trend was only clear under a load of 20 N. The composites with the highest wear resistance had the highest copper ligament diameter of 30 μm. This was probably due to the higher heat conductivity and fracture toughness caused by the coarse, fibrous copper network replicating the wool felt used to produce it. This was possibly also because these composites had the smallest alumina grain size.

Highlights

► Increasing copper fraction increased the wear rate, except where a tribolayer formed or where the alumina grains were weakly bonded. ► Increasing the copper ligament diameter decreased the wear rate. ► The grain size of alumina affected the wear behaviour. ► Composites with the coarsest copper network had the highest wear resistance, probably due to the higher heat conductivity and fracture toughness.

Introduction

Metal ceramic composites are an exciting field of research, enabling a combination of properties not possible in metals or ceramics alone. Alumina–copper composites enable the high heat and electrical conductivity and high toughness of copper to be combined with the high stiffness and hardness of alumina [1].

There are three broad types of ceramic–metal composites. Metal Matrix Composites (MMC) have a continuous metal network with Vmetal > Vceramic. Ceramic Matrix Composites (CMC) have a continuous ceramic network and Vceramic > Vmetal. In Interpenetrating Network Composites (INC) both components are continuous. The only limitation on composition is that there is enough of one phase for it to percolate. Depending on the form of the phase this is around 20–30%, less for fibres and more for particles. The advantages of a continuous ceramic network include a higher Young's modulus, hardness and load bearing capacity than MMCs.

Interpenetrating metal ceramic composites are generally produced via squeeze casting [2] or gas pressure infiltration [3] due to the poor wettability of most metals on oxide ceramics. The wetting angle of copper on alumina is extremely low with a contact angle varying between 124° [4] and 170° [5]. By adding CuO to copper, wettability increases so much that porous alumina can be infiltrated without external pressure [6]. The drawback is that an aluminate forms at the interface of copper and alumina [7] which can decrease the interface strength [8]. Upon cooling after infiltration, large residual stresses build up in the composite (tension in copper and compression in the alumina) due to the difference in thermal expansion coefficient [9], [10]. Scientifically this is an interesting aspect of ceramic–metal composites, since it can affect the mechanical properties and thermal expansion behaviour.

The combination of properties possible in alumina–copper composites makes them particularly interesting for wear applications, for example in automobile, aerospace and even bicycle parts.

Until now relatively few wear studies have been published on alumina–copper composites, with only one found by the authors in the literature. In that study, copper was reinforced with particulate alumina [11]. The authors did not find any articles on the wear of alumina–copper interpenetrating composites, or indeed any copper-based interpenetrating composites. A similar system which has already received quite some attention is alumina–aluminium composites. Aluminium–alumina MMCs have been explored for over 3 decades, interpenetrating composites since the mid-1990s. One drawback of this combination is the relatively low melting point of aluminium (660 °C) limiting the possible temperatures of application. Since copper has a higher melting point (1083 °C), alumina–copper composites have the potential for applications in a wider temperature range. Copper also has a higher thermal conductivity than aluminium (401 W/m K compared with 236 W/m K at 0 °C [12]), allowing heat generated at the wear surface to be more rapidly dispersed. This reduces thermal mismatch stresses and the likelihood of thermal shock.

Wear is not a materials property, but rather a system response [13]. There are many factors influencing wear behaviour, which make it difficult to compare results from different laboratories or different testing methods. Even results from a single laboratory and using the pin-on-disc dry sliding wear test have been shown to have a variation of 28–56% [14]. Despite this, it is interesting to compare trends found in previous studies. Relevant results from the literature have been summarized below.

In aluminium MMCs, increasing alumina content has been found to increase wear resistance [15], [16], [17]. With fibrous alumina reinforcement this effect was only seen up to a content of 20%, above which there was no further improvement [16]. There were no studies found on the effect of the amount of alumina reinforcement in copper MMCs, but in Cu–TiB2–TiN MMCs, the same trend as above was found [18]. Surprisingly, El-Hadek and Kaytbay found the wear resistance of pure copper to be higher than that of copper–alumina MMCs with 15% alumina and grain sizes of 6 and 20 μm [11]. When reinforced with 100 nm alumina grains the wear resistance of the composite was higher. The larger grains could be more easily removed from the copper matrix, thus increasing weight loss, whereas the 100 nm grains were more strongly bonded to copper. In alumina–aluminium interpenetrating composites, increasing alumina content was also found to increase wear resistance [19], [20]. It was also found that interpenetrating composites have a significantly higher wear resistance than MMCs [21], [22]. It would be expected that the same trend be found in alumina/copper interpenetrating composites.

Increasing metal cell size was recently found to increase wear resistance in alumina–aluminium interpenetrating composites [9], [19]. It was postulated that a larger distance between thicker alumina struts provided a more effective shielding of the composite. There are conflicting results in the literature on the effect of ceramic grain size on the wear resistance of MMCs. Some studies found increasing particle size increases wear resistance in aluminium MMCs [15], [16], stating that increased particle size causes more of the load to be carried by the hard particles thus lowering the wear of the softer matrix. Others found that wear resistance increases with decreasing particle size, both in Al–MMCs [23] and Cu–MMCs [9], [11], [24]. These studies postulated that larger particles are more likely to contain microcracks, therefore fracture more readily and decrease wear resistance. The effect of ceramic grain size on the wear behaviour of interpenetrating composites has not been found by the authors in the literature. This is not surprising, since the (fully sintered) ceramic phase is continuous throughout the composite and one would not expect the grain size to have a significant effect. In the wear of pure alumina, refining grain size was found to slightly increase wear resistance in the mild wear regime, and significantly delay the transition from a mild to a severe wear regime [25]. It is generally agreed that increasing the applied load increases the wear rate [17], [18], [19], [21], [23]. In the mild wear regime this only seems to have a small effect. However, the transition from mild to severe wear, at which point a sudden increase in wear rate of around two orders of magnitude, occurs earlier (in terms of sliding distance) with increasing load [17].

In this study, alumina–copper composites have been produced by gas–pressure infiltration of liquid copper into a porous ceramic perform [26]. The porous ceramics were produced using different sacrificial preforms in order to achieve a wide range of microstructures and metal contents, as in [27]. Using this method, a metal content from 20 to 55% and a metal ligament/cell diameter of 0.5–30 μm has been achieved. This allows for a variation in properties which can be tailored to suit the application. The effect of metal content, ceramic grain size and metal cell or ligament diameter on the wear behaviour of the composites has been explored. Selected samples were also tested at a different load.

Section snippets

Production of composites

Porous Al2O3 samples of varying pore size and porosity were made by three different methods:

  • a)

    Partial sintering of a 5 μm Al2O3 (HVA FG, Almatis) slurry with 45 vol% solid in distilled water with 0.2 wt% Dolopix CE64, Zschimmer und Schwarz (a deflocculant) and 0.2 wt% Glydol Glydol N 109, Zschimmer und Schwarz (a wetting agent) to get a pore size of 0.5–1 μm. Final porosity was controlled by the sintering temperature: 1450 °C for 40%, 1550 °C for 35%, 1620 °C for 30%, 1660 °C for 25% and 1695 °C for 20%

Microstructure

SEM micrographs of the microstructures of the four different types of samples are presented in Fig. 1. The same magnification is used for all samples for the purpose of comparison. The size and shape of the starch and wool sacrificial preforms were faithfully reproduced, however, the proportion of copper varied up to 4% from the target value.

The residual porosity of starch based samples C20 and R20 was quite high, around 20%. This would suggest that there were regions of closed porosity which

Conclusion

In alumina–copper interpenetrating composites, increasing the amount of copper decreased hardness and increased wear, except where cyclic tribolayer behaviour occurred or where the alumina grains were weakly bonded. Under a load of 20 N increasing the copper ligament diameter decreased the wear rate. Under a load of 10 N this trend was not clear, although the specific wear rates were generally lower than under a 20 N load. The wear mechanisms of pure copper and pure alumina were adhesive and

Acknowledgements

This work was supported by the EU Network of Excellence project Knowledge-based Multicomponent Materials for Durable and Safe Performance (KMM-NoE) under the contract number NMP3-CT-2004-502243. Rod Martin (MERL, Hertfordshire) kindly enabled the wear testing to be conducted and assisted in the interpretation of the results obtained. Jan Dusza (IMR-SAS, Košice) facilitated hardness measurements and helped to interpret initial SEM images of the wear tracks. Mark Hoffman (UNSW, Sydney) encouraged

References (36)

  • H. Prielipp et al.

    Strength and fracture-toughness of aluminum alumina composites with interpenetrating networks

    Materials Science and Engineering: A

    (1995)
  • B.D. Flinn et al.

    Evolution of defect size and strength of porous alumina during sintering

    Journal of the European Ceramic Society

    (2000)
  • F.F. Lange et al.

    Method for processing metal-reinforced ceramic composites

    Journal of the American Ceramic Society

    (1990)
  • N.A. Travitzki et al.

    Microstructure and properties of metal infiltrated RBSN composites

    Journal of the European Ceramic Society

    (1992)
  • T.E. O’Brien et al.

    Effect of oxygen on the reaction between copper and sapphire

    Journal of the American Ceramic Society

    (1974)
  • B.C. Allen et al.

    Surface tension and contact angles in some liquid metal–solid ceramic systems at elevated temperatures

    Transactions of the Metallurgical Society AIME

    (1959)
  • M. Diemer et al.

    Influence of oxygen partial pressure and oxygen content on the wettability in the copper–oxygen–alumina system

    Journal of the American Ceramic Society

    (1999)
  • K.P. Trumble

    Prediction of a critical temperature for aluminate formation in alumina/copper–oxygen eutectic bonding

    Journal of the American Ceramic Society

    (1999)
  • Cited by (34)

    • Fabrication, mechanical properties, and wear behaviors of co-continuous TiC-steel composites

      2022, Materials Characterization
      Citation Excerpt :

      They found that the wear mechanisms of the samples were abrasive wear, oxidation wear and two-body wear under high load, while three-body wear worked under lower load, and the 3D continuous network of SiC ceramics as the reinforcement of Al-based composite can avoid the third body wear effectively. Jami Winzer et al. [18] prepared a variety of alumina‑copper interpenetrating composites with different copper ligament diameters and volume fractions of copper, and reported that the dominant mechanism in the interpenetrating composite with narrow copper ligament diameters was a mixture of adhesive and oxidative wear, while abrasive and oxidative wear pertained the composite with larger copper ligament diameters. Fu et al. [19] reported that the SiC foam structures, which with various inter-skeleton distances and skeleton distributions, had an obvious influence on the wear behavior of the co-continuous composites by affecting the dispersion and transmission of the load and friction heat during the wear process.

    • Synthesis and mechanical properties of TiC-Fe interpenetrating phase composites fabricated by infiltration process

      2018, Ceramics International
      Citation Excerpt :

      The ceramic-metal interpenetrating phase composites (IPCs) have many potential applications where wear and cut resistance and thermal conductivity are required [1,2]. The ceramic-metal IPCs, such as SiC-Al, Al2O3-Al, Al2O3-Cu and Ti2AlC-Mg, have good performance because of their unique three-dimensional (3D) network structure [3–8]. Among these composites, TiC-Fe composite exhibits excellent mechanical and wear resistant properties due to synergy of high strength and toughness of Fe and high hardness and favorable wear resistance of TiC [9,10].

    • Effect of microstructure on mechanical properties and residual stresses in interpenetrating aluminum-alumina composites fabricated by squeeze casting

      2018, Materials Science and Engineering: A
      Citation Excerpt :

      The measurements of alumina preforms porosity are summarized in Table 2 for different contents of rice starch (RS) used as the pore forming agent (PFA). The rice starch was selected based on the study of PFA's in alumina preforms [35]. The total porosity of our preforms as measured by mercury porosimetry varied from 34.75% to 54.39% for 0%RS and 40%RS, respectively.

    • Wear behavior of copper matrix composites reinforced by γ-Cu<inf>5</inf>Zn<inf>8</inf> nanoparticles

      2017, Powder Technology
      Citation Excerpt :

      Hence, the sufficient time for achieving an appropriate morphology of alloy powders is 15 h milling times. Fig. 5d shows the FESEM micrograph after 20 h milling, which indicates that the powder particles are strongly fined and that it is not an appropriate morphology, because the specific surface area of spherical particle is low, while flake has higher surface area [12–23], moreover, it should be noted that the flake morphology of copper powders provided new surfaces for embedding nano powders and, therefore, nano particles easily embedded on the flake copper powders. Furthermore, cold welding between fine and coarse copper powders was very low [24].

    View all citing articles on Scopus
    View full text