Вы находитесь на странице: 1из 26

3.

01 An Introductory Overview of MMC Systems, Types, and Developments


T. W. CLYNE University of Cambridge, UK
3.01.1 INTRODUCTION 3.01.1.1 The Historical Context of MMCs 3.01.1.2 Property Prediction for MMCs and the Concept of Load Transfer 3.01.2 A SURVEY OF MMC TYPES AND DEVELOPMENTS 3.01.2.1 Broad Microstructural Features and Property/Ease of Processing Combinations 3.01.2.2 Discontinuously-reinforced MMCs 3.01.2.2.1 Particulate MMCs 3.01.2.2.2 Short fiber and whisker-reinforced MMCs 3.01.2.3 Continuously-reinforced MMCs 3.01.2.3.1 Multifilament MMCs 3.01.2.3.2 Monofilament MMCs 3.01.2.4 Layered MMC Systems 3.01.2.5 Special Category MMCs 3.01.2.5.1 Metallic foams 3.01.2.5.2 Cermets 3.01.2.5.3 Reactively processed MMCs 3.01.2.5.4 Intermetallic matrix composites (IMCs) 3.01.2.5.5 Graded MMC coatings and structures 3.01.3 SOME CURRENT PROCESSING AND PERFORMANCE ISSUES FOR MMCS 3.01.3.1 3.01.3.2 3.01.3.3 3.01.3.4 Process Economics Microstructural Quality Control Damage, Ductility, and Fracture Environmental Stability and Recycling 1 1 3 4 4 6 6 8 9 9 10 12 13 13 14 16 19 19 20 20 21 21 22 22 24

3.01.4 MMC CONTACTS, NETWORKS, AND SOURCES OF INFORMATION 3.01.5 REFERENCES

3.01.1

INTRODUCTION

3.01.1.1

The Historical Context of MMCs

This first chapter of the volume will largely serve to set the scene for an appraisal of the current status of MMCs, with relatively little in the way of detailed information but extensive cross-referencing to other chapters of the volume and pointers given to other sources. 1

Reinforced materials based on metals have long been of technological significance. For example, layered structures of metallic and nonmetallic constituents, made by repeated lamination and hammering, were produced in several ancient civilizations (Muhly, 1988). Of

An Introductory Overview of MMC Systems, Types, and Developments short fibers. A combination of good properties, low cost, and high workability has made them attractive for many applications and their commercial exploitation has already become quite significant. These materials fall somewhere between the dispersion-strengthened and fiberstrengthened extremes. They differ from dispersion hardened systems in having large (*1 100 mm diameter) reinforcing particles, which contribute negligible Orowan strengthening, and in containing a relatively high (*550%) volume fraction of reinforcement, such that load transfer from the matrix is significant. However, unlike long fiber reinforced systems, the matrix does bear a substantial load and its strength is important. These distinctions are highlighted by the schematic plots (Clyne and Withers, 1993) shown in Figure 1, which illustrate how strengthening is strongly dependent on reinforcement size for dispersion and precipitation hardened metals, but is sensitive to reinforcement content and aspect ratio for MMCs. This is not, however, to say that the presence of the reinforcement has no effect on dislocation structures within MMCs. In fact, there are various effects, the details of which are given in Chapter 3.02, this volume. A further complication worthy of note is that the reinforcement may significantly influence the manner in which fine precipitates nucleate and grow in the matrix: this is explored in depth within Chapter 3.03, this volume. In general, there has been progressive development on a number of fronts over the past 15 years, although commercial and scientific interest has tended to fluctuate somewhat in response to changing economic and technical climates. A broad summary of the current situation would be that most aspects of the behavior of MMCs are now fairly well understood, so that the limitations and attractions of their processing and performance characteristics are reasonably clear. A major objective of this volume is to give detailed and up-to-date information about these characteristics. Industrial usage levels still remain relatively low at present, but this is changing, at least for certain types of discontinuously-reinforced MMCs, as the slow process of conversion to use of a new class of material is tackled for individual components. Whether usage of MMCs will in due course rise to levels comparable with those of the main alternatives currently considered by engineers for structural and functional purposes, or whether they will remain at the level of a resource considered only for relatively specialized purposes, is not clear at this stage. While processing limitations and economic factors mean that very large tonnage exploitation looks unlikely, it does appear probable that

course, it is only in relatively recent times that the significance of various microstructural features in metals has been fully appreciated, allowing the systematic development of monolithic and composite metal-based materials. Dispersion hardened metals and precipitation hardening systems were developed several decades ago. In both cases, the basis of the strengthening mechanism is to impede dislocation motion with small particles. Dislocations are forced to bow around these obstacles, via a mechanism first identified by Orowan. This is achieved by the incorporation of either fine oxide particles or precipitates nucleated and grown within a metallic matrix. These must be closely spaced (5*1 mm apart) if significant strengthening is to be achieved. Since it is generally possible to achieve finer distributions (by nucleation and growth) in precipitation hardened systems, these normally exhibit higher strengths at room temperature. However, dispersion strengthened systems show advantages at elevated temperature, because of the high thermal stability of the oxide particles. These materials would not, however, normally be classified as true composites. While there is no universally accepted definition of a composite, it is commonly assumed that it is only when significant load transfer occurs between matrix and reinforcement under the influence of an applied stress that the term can properly be applied. When a composite is subjected to an external load, the matrix is relieved of a substantial proportion of that load by the presence of the reinforcement. On this basis, conventional dispersion and precipitation hardened systems are not composites, since they typically contain only around 1% or less of second phase and at such levels the reinforcing constituent cannot significantly reduce the stress borne by the matrix. The development of true MMC materials, such as aluminum or copper reinforced with 3070% of continuous tungsten or boron fibers, was initiated in the 1960s. As is the case with most polymeric composites (PMCs), an applied load is largely borne by the fibers in such a material and the matrix microstructure and strength are relatively unimportant. However, it should be recognized that, in contrast to the corresponding PMCs, commercial usage of long fiber reinforced MMCs has not yet reached levels which could be considered industrially significant, at least in terms of tonnage consumption, and it is by no means certain that such levels will be reached in the future. Discontinuously-reinforced metal composites were developed during the 1980s, with attention focused on Al-based matrices reinforced with SiC particles, or Al2O3 particles or

Introduction

Figure 1 Schematic illustration (Clyne and Withers, 1993) of the magnitude and range of operation of composite strengthening mechanisms, as a function of reinforcement aspect ratio (s), size (d), and volume fraction (f). The strength (ratio of the yield stress of the composite to that of the matrix) is plotted in (a), while in (b) both this and the corresponding stiffness ratio are shown. In case (a), matrix strengthening dominates, with the reinforcement constituting too low a volume fraction to carry a significant proportion of the load. In (b), both strengthening and stiffening effects are primarily a consequence of load transfer to the reinforcement, which is too coarse to strengthen the matrix by affecting dislocation motion.

MMCs will come to be considered as a viable option for a wide range of applications. The situation concerning industrial usage of discontinuously-reinforced MMCs is summarized in Chapter 3.26, this volume, while the more specialized applications being developed for long fiber systems are outlined in Chapter 3.27, this volume. Chapters 3.30 and 3.31, this volume, cover recent activities in the area of MMC recycling and the development of metallic foams, which may be considered as a special type of MMC.

3.01.1.2

Property Prediction for MMCs and the Concept of Load Transfer

When designing an MMC, an objective might be to combine the high ductility and formability of the matrix with the stiffness and load-bearing capacity of the reinforcement,

or perhaps to unite the high thermal conductivity of the matrix with the low thermal expansion of the reinforcement. In attempting to identify attractive matrix/reinforcement combinations, it is often illuminating to derive a merit index for the performance required, in the form of a specified combination of properties. Appropriate models can then be used to place upper and lower bounds on the composite properties involved in the merit index, for a given volume fraction of reinforcement. The use of maps, with material properties as axes, can then be very useful in highlighting how the combinations offered by different classes of material compare with each other. The framework for such comparisons and predictions, covering a range of areas within materials science, has been clearly set out by Ashby. In Chapter 3.29, this volume, he outlines the application of these approaches to MMCs. In order to implement these, a systematic and reliable database of properties is needed.

An Introductory Overview of MMC Systems, Types, and Developments and the remainder by the matrix. Provided the response of the composite remains elastic, this proportion will be independent of the applied load and it represents an important characteristic of the material. It depends on the volume fraction, shape, and orientation of the reinforcement and on the elastic properties of both constituents. The reinforcement may be regarded as acting efficiently if it carries a relatively high proportion of the externally applied load. For example, stiff fibers aligned in the loading direction carry a relatively high load, whereas particles and transversely oriented fibers do not. The concept of elastic load transfer has long been familiar to those working with (polymer-based) fiber composites. It is readily translated to the elastic behavior of MMCs, although calculation of the load partitioning is often more complex as a result of the greater interest in discontinuous (short fiber and particulate) reinforcement, as opposed to continuous fibers. These cases can, however, be treated using numerical techniques or analytical methods such as the Eshelby approach (which is described in MMC textbooks such as the two cited above and is also covered in detail in Chapter 1.14, Volume 1, of the present series). Efficient load transfer often results in higher strength, as well as greater stiffness, because the reinforcement is usually stronger, as well as stiffer, than the matrix. While stiffening effects are straightforward, strengthening is slightly more complex, since there may be contributions both from load transfer and from in situ matrix strengthening (see Figure 1). While the latter can often be predicted using well-established laws and correlations drawn from dislocation theory and metallurgical experience, strengthening by load transfer is less simple, particularly when the matrix starts to undergo plastic deformation. This is expected to cause rapid transfer of load to the reinforcement, but in practice this often stimulates other phenomena, such as internal damage development (leading to fracture) and/or stress relaxation effects such as creep. These two topic areas are respectively covered within the present volume in Chapters 3.043.08 and 3.133.15. 3.01.2 3.01.2.1 A SURVEY OF MMC TYPES AND DEVELOPMENTS Broad Microstructural Features and Property/Ease of Processing Combinations

Shortcomings of the currently available MMC database and predictive capacity, in terms of both scope and reliability, are responsible for much of the caution sometimes expressed by engineers concerning wider MMC usage. For many composite properties, upper and lower bounds can be identified, based on corresponding properties of the constituents. In practice, however, there is often interest in establishing composite properties to a greater precision than is possible by the use of bounds, which in many cases are widely separated. This can be relatively complex, particularly for MMCs, in which certain matrix properties may be significantly affected by the presence of the reinforcement. Nevertheless, reliable approaches have been developed for prediction of many of the basic properties of MMCs. These include elastic stiffness, thermal expansion, the onset of plasticity, and work hardening characteristics. These areas are systematically covered in textbooks such as those of Clyne and Withers (1993) and Arsenault and Taya (Taya and Arsenault, 1989). This volume does not include comprehensive treatment of these areas, although some of them are covered in chapters contained in Volume 1 of the present series. Chapters in this volume are largely directed towards more complex performance characteristics, such as the fracture, creep, wear, and corrosion behavior of MMCs. These are in practice often pivotal to the question of whether available MMCs are suitable for particular applications, or could be made so by appropriate microstructural control. Since the present volume does not include treatment of factors affecting the stiffness, yielding, and work hardening of MMCs, it is appropriate here to give a brief outline of the concept of load sharing between the matrix and the reinforcing phase, which is central to an understanding of these characteristics. Under an applied mechanical load, the stress within an MMC may vary from point to point, but the proportion of the external load borne by each of the individual constituents (matrix and reinforcement) can be evaluated by finding the volume-averaged stress within each of them. The external load must equal the sum of the volume-averaged loads borne by the constituents, so that
(1 7 f)sm + fsr = sA (1)

which relates the volume-averaged matrix and reinforcement stresses sm and sr to the applied stress sA, with a volume fraction f of reinforcement. Thus, for a simple two-constituent MMC under a given applied load, a certain proportion of that load will be carried by the reinforcement

Generic classification as an MMC encompasses a wide range of scales and microstructures. Common to them all is a metallic matrix,

A Survey of MMC Types and Developments

Figure 2 A schematic depiction (Clyne and Withers, 1993) of the main MMC systems, classified according to the type of reinforcement.

which is normally contiguous. The reinforcing constituent is in most cases a ceramic, although there are exceptions to this and MMCs can be taken to encompass materials reinforced with relatively soft and/or compliant phases, such as graphitic flakes, lead particles, or even gases. It is also possible to use refractory metals, intermetallics, or semiconductors, rather than true ceramics. This flags up a number of uncertainties about whether various materials should be considered as MMCs. Examples include AlSi eutectic, cast irons, and the so-called dual phase steels (which contain about 20% of large, hard martensitic particles in a ferritic matrix). There is no correct answer to this question, but it can certainly be argued that such materials are much closer to conventional MMCs in terms of behavior and performance characteristics than the precipitation hardened and dispersion strengthened systems referred to above in Section 3.01.1.1. MMC types are commonly subdivided, as depicted in Figure 2, according to whether the reinforcement is in the form of (i) long aligned fibers (allowing high reinforcement contents), (ii) short fibers (with or without a degree of alignment), or (iii) particles, which are at least approximately equiaxed. A similar classification is commonly applied to PMCs. A major difference between the two classes of composite is that, while the progression (i) to (iii) is in general one of decreasing industrial significance for PMCs, the emerging picture for MMCs is that the current level of commercial interest rises for the same sequence. It might be considered that, for both PMCs and MMCs, the broad trends on moving from (i) to (iii) would be for reductions in performance, but compensatory improvements in formability and ease of processing. However, while this is probably a fair generalization for PMCs, certain points can be identified for MMCs which alter the picture somewhat. First, metals are highly formable, making secondary processing very attractive

provided the reinforcement content is such as to allow this, whereas for PMCs it is only with thermoplastic matrices that postconsolidation forming is possible. Second, long fibers, particularly when aligned and present at high volume fractions, severely constrain plastic deformation of the matrix in MMCs, inhibiting the highly efficient energy absorption mechanism of dislocation motion and thus substantially reducing the toughness of the metal. In PMCs, the same inhibition of matrix plasticity occurs, but the matrix toughness is in any event very low and energy absorption via interfacial debonding and fiber pull-out becomes highly significant. Such fiber pull-out is often rather more difficult to promote in MMCs, for complex reasons concerned with interfacial properties and fracture behavior, but in any event the inescapable fact is that, while polymeric matrices are substantially toughened by the introduction of long fibers, most metallic matrices are substantially embrittled. This effect is largely responsible for the trends in popularity between different MMC types noted above. It does not, of course, eliminate the possibility of applications being identified in which this shortcoming of long fiber MMCs is outweighed by attractive aspects of their performance. Of course, in examining the possibility of using novel materials such as MMCs for particular applications, the economic factors involved are often of paramount importance. These may be difficult to quantify in some cases, since there are often indications that production and processing costs would fall significantly as usage became well established, but it may be difficult to obtain guarantees or accurate projections. Nevertheless, approaches are available to cost-benefit analysis for a switch to use of a different material. In Chapter 3.29, this volume, a framework is presented within which performance and cost indicators can be manipulated in order to assess the overall benefit, with various examples of how MMCs fare in

An Introductory Overview of MMC Systems, Types, and Developments are in principle also worth considering. In fact, silicon carbide and alumina score rather heavily over most of the competing alternatives in terms of availability and cost, and they are also quite attractive in terms of relevant strength properties. There are additional issues which may need consideration, such as thermal expansion, thermal/electrical conductivity, chemical compatibility with the matrix during processing, and the ease with which a strong interfacial bond can be formed. These aspects are respectively considered in Chapters 3.13, 3.16, 3.20, 3.21, 3.22, and 3.05, this volume. While all ceramics can in general be taken as electrical insulators in comparison with metals, their thermal conductivity can be comparable to, or even greater than, those typical of metallic systems. This arises because heat can be conducted by phonons, as well as electrons, and phonon transport is favored by a light, stiff crystal lattice. The thermal conductivities of SiC and diamond are particularly high. However, it should be noted that phonons are readily scattered by defects and experimental values are often relatively low as a consequence of the presence of grain boundaries and other microstructural features. Chemical reaction during processing can occur with some particulate reinforcements. Silicon carbide can be particularly problematic when incorporated into Al- and Ti-based MMCs. There is extensive documentation (Skibo et al., 1988; Lloyd et al., 1989; Lloyd 1994) concerning the reaction between SiC and aluminum melts and there is also evidence (Warwick and Smith, 1991; Gordon et al., 1994; Turner et al., 1996) that SiC reacts with Ti during solid-state consolidation to a greater extent than occurs with certain alternative reinforcements such as TiB2. Reactions also occur between SiC reinforcement and ferrous matrices (Pelleg, 1999). Alumina is usually less reactive than SiC in aluminum, but it does react quite strongly with titanium at elevated temperature. Magnesium is rather different from Al and Ti in that it does not readily for a stable carbide, but it does have a high affinity for oxygen. The greater stability of Al2O3 compared with SiC in aluminum is therefore reversed for magnesium matrices. In general, while coatings or other surface treatments may be worth considering for fibers (particularly monofilaments), economic and practical considerations mean that particulate reinforcement is normally introduced into MMCs in the virgin state. This may, however, be such that a surface oxide layer is present and deliberate thickening of this layer, for example by heat treatment in air, has in some instances been

comparison with competing materials for different classes of application. It is emphasized that the property profiles obtainable with MMCs often fall outside the envelopes available from conventional (monolithic) materials and that this might have particular relevance and attractions for certain applications. Costs, however, are sometimes inherently high and this must be balanced against performance benefits considered in the context of the application. Realistic quantitative assessments of this type will be essential if MMC usage is to become fully established in the applications for which this would be genuinely beneficial.

3.01.2.2 3.01.2.2.1

Discontinuously-reinforced MMCs Particulate MMCs

This type of MMC has been the subject of intensive study for a range of industrial applications. While much of this interest centers around aluminum alloy matrices, there have also been activities concerning titanium-, steel, and magnesium-based systems. The particulate material employed is most commonly SiC or Al2O3, but many others have been investigated. These include TiB2, B4C, SiO2, TiC, WC, BN, ZrO2, etc. In general, the attributes ideally exhibited by particulates for MMC reinforcement are: (i) high stiffness (ii) low density (iii) high hardness (iv) adequate toughness (v) good availability/low cost for suitablysized powder (*120 mm diameter). Detailed information about particulate reinforcement, and also about whiskers (see Section 3.01.2.2.2), is given in Chapter 1.06, Volume 1. In order to get an overview of how the candidate materials compare with each other, a map (Ashby, 1993) such as that shown in Figure 3 can be constructed. This depicts the combinations of stiffness and density offered by various potential reinforcements, together with those of some metallic matrices. The dotted lines represent gradients corresponding to constant values of the ratio between the stiffness (raised to different powers) and the density. For a given ratio (relative importance of stiffness and density), a reinforcement lying far towards the upper left from the line on which the matrix lies will be efficient, whereas one lying close to that line will provide little benefit. It can be seen that, on this basis (i.e., neglecting attributes (iii)(v) in the list above), SiC and Al2O3 are confirmed as being attractive, although others, such as B4 C (and possibly diamond!),

A Survey of MMC Types and Developments

Figure 3 Representation (Ashby, 1993), on a plot of Young's modulus against density, of the property combinations exhibited by common metallic matrices and by various candidate reinforcement materials. These include both fibrous and particulate constituents. The three dotted guidelines shown have gradients corresponding to specific ratios of the modulus, raised to different powers, to the density. Each ratio may be regarded as a merit index (M), the value of which increases most sharply in the directions shown by the three arrows.

found to have a beneficial effect on interfacial bonding or other characteristics (Ribes and Suery, 1989; Bardal and Hier, 1991; Gergely and Clyne, 1998). Particulate MMCs are usually manufactured on a commercial basis either by a melt incorporation and casting technique or by powder blending and consolidation (see Chapters 3.21 and 3.25, respectively). More specialist production routes, which are not as widely used at industrial production levels, involve reactive processing (see Section 3.01.2.5.3) or spray codeposition (see Chapter 3.23, this volume). Quality control objectives include the elimination of excessive interfacial reaction during processing, particularly for melt routes (see Chapter 3.21, this volume), and also the avoidance of microstructural defects such as poor interfacial bonding, internal voids, and cluster-

ing of the reinforcement, (see Chapters 3.04, 3.05, and 3.06, this volume, respectively). Typically, reinforcement particles are about 10 20 mm in diameter and constitute about 10 30% by volume of the material, although MMCs in which the values concerned lie outside of these ranges have been studied and are available commercially (particularly finer particles and higher particle contents). The range of applications for which particulate MMC components are being developed is wide, with a particular concentration in the automobile industry (see Chapter 3.26, this volume). They are certainly being very heavily exploited in formula 1 racing, for which components such as brake calipers are almost universally made of MMC material, but applications are now emerging for production cars. An example is shown in Figure 4, which is a

An Introductory Overview of MMC Systems, Types, and Developments

Figure 4 (a) Chassis and frame structure of a Lotus Elise sports car, showing various innovative features and the logos of some collaborating firms. Item 20 is a particulate MMC brake disk produced by Lanxide Corporation, which is also shown, (b), in close-up. The cooling fins help to dissipate the heat and keep the temperature below that at which the matrix would start to soften appreciably.

photo of a brake disk on a production car. In this application, it is the superior wear performance of MMCs (see Chapter 3.19, this volume) which is of primary interest, although the stiffness enhancement is also beneficial. This allows replacement of conventional cast iron disks, at a total weight saving of several kilograms, provided that the temperature of the aluminum matrix can be kept reasonably low. As detailed in Chapter 3.19, this volume, transition to a severe wear mode occurs above a critical temperature in these materials. The high thermal conductivity of Al (which is retained when SiC reinforcement is introducedsee Chapter 3.16, this volume) favors the avoidance of hot spots on the disk and the active cooling stimulated by the incorporation of cooling fins in the design ensures that the disk as a whole is kept fairly cool. 3.01.2.2.2 Short fiber and whisker-reinforced MMCs

Short fiber reinforced MMCs first attracted widespread attention in the mid-1980s with the

development of aluminum diesel engine pistons selectively reinforced with short alumina (Saffil) fibers (Donomoto et al., 1983). Other rather similar (aluminosilicate) fibers have also been extensively employed for applications of this type. The components concerned are normally produced by a melt infiltration route (see Chapter 3.20, this volume) and some details of the piston case history are given in Section 3.20.1.2 of Chapter 3.20, this volume. These fibers have a fine-grained polycrystalline microstructure (Birchall, 1983) and they are manufactured via a melt spinning route (see Chapter 29 in Volume 1 of the present series). Typical fiber diameters are a few microns and they are initially produced in lengths of several hundred microns. However, they are often broken up during MMC manufacture, so that fiber aspect ratios in the composite material might range from 100 down to about 3 or so, depending on the processing route and conditions. Short fiber MMCs can offer some attractive combinations of properties and processability. They can be manufactured by a melt route such as squeeze infiltration or by powder blending and consolidation (see Chapters 3.20, and 3.25,

A Survey of MMC Types and Developments this volume, respectively). Interfacial characteristics are partly dependent on the degree of reaction during processing, which is in turn affected by the surface chemistry of the fibers: for example, Saffil fibers have a thin silica-rich surface layer which tends (Cappelman et al., 1985; Kaufmann and Mortensen, 1992) to react with an aluminum melt during processing, particularly if Mg is present. Some secondary processing, such as forging and extrusion, can be carried out under appropriate conditions and in certain cases superplastic matrix behavior can be stimulated, particularly when thermal cycling is imposed (see Chapter 3.15, this volume). However, their formability is in general markedly inferior to that of particulate MMCs. On the other hand, certain property advantages over particulate MMCs are common, particularly in terms of resistance to creep and, to a lesser extent, wear (see Chapters 3.14, and 3.19, this volume, respectively). The mechanical properties of fibers such as Saffil are relatively good, but they can in general be appreciably enhanced if their finegrained structure can be replaced by that of a single crystal. The so-called whiskers which result (Levitt, 1970) when fine, slender fibers are produced as monocrystals created considerable interest as early as the 1960s. Whiskers are usually submicron, or perhaps around a micron, in diameter, with aspect ratios of up to several hundred. Their tensile strengths are often very high and, depending on the material and crystallographic growth direction, they may also exhibit higher elastic moduli along the fiber axis than the corresponding polycrystal (Parvizi-Majidi, 1993). The main barrier to widespread commercial use of whiskers was initially that it was difficult to produce them under economically attractive conditions. This situation evolved somewhat in the late 1980s with the development of methods for the production of SiC whiskers from various cheap starting materials. For example, their extraction from rice husks has been described (Nutt, 1988). This led to a surge in interest and various studies (Nardone and Strife, 1987; Hong et al., lez-Doncel, 1989; 1988; Daehn and Gonza Mason et al., 1989; Christman et al., 1989; Badini et al., 1991; Barlow and Hansen, 1991; Dutta et al., 1993) were undertaken on SiC whisker-reinforced MMCs. There was also work (Matsubara et al., 1987; Kang et al., 1991) on MMCs reinforced with whiskers of other ceramics such as Si3N4 and Al2O3. However, while excellent properties have been reported for whisker-reinforced MMCs, and there is still some interest in them, work in this area tailed off in the early 1990s and there has been little commercial exploitation. This is

largely a consequence of handling difficulties. These commonly arise with very fine fibers, which tend to form tenacious ball-like structures and are difficult to orient in a controlled way. However, the most significant problem of this type concerns perceived health hazards. Whiskers and whisker fragments in the slightly submicron size range can become airborne very readily and are likely to reach the lungs because they tend to evade the natural protective mechanisms operative in the nasal passages and throat. Asbestos fibers behave similarly. When evidence was produced (Birchall et al., 1988) indicating that SiC whiskers are potential carcinogens in this situation, it became clear that rigorous safety measures would be necessary during handling prior to incorporation in the matrix and this represented a strong disincentive to further research and development activities.

3.01.2.3 3.01.2.3.1

Continuously-reinforced MMCs Multifilament MMCs

A number of long fiber MMCs systems have been investigated and some of these have been used in certain industrial and military applications (see Chapter 3.27, this volume). However, as a consequence of processing difficulties and of the constraints on ductility and toughness outlined in Section 3.01.2.1, this usage has in general remained rather specialized and limited. The term multifilament refers to relatively small diameter (*530 mm diameter) fibers which, as a consequence of their low beam stiffness (which scales as the fourth power of the diameter), are flexible enough to be handled as tows or bundles which can be woven, braided, filament wound, etc. The materials concerned include carbon fibers (see Chapters 1.02 and 1.03, Volume 1), SiC fibers (Chapter 1.05, Volume 1), and various oxide fibers (Chapter 1.06, Volume 1). There are various other types of multifilament in common use (e.g., see Chapters 1.07, 1.09, 1.10, 1.11, and 1.12, Volume 1), but these are either unable to survive the elevated temperatures involved in MMC production (e.g., polymeric and organic fibers) or are of limited interest for MMCs because of relatively poor mechanical characteristics such as stiffness or creep resistance (e.g., most glass fibers). Multifilament MMCs can be produced by melt infiltration, although problems arise with unidirectionally aligned fibers in that the applied melt pressure transverse to the fiber axis tends to bring them into close contact and reduce the channels between them to such small dimensions that the melt is unable to penetrate

10

An Introductory Overview of MMC Systems, Types, and Developments

Table 1 Data for the maximum curvature sustainable before fracture for various fibers used in MMCs. Diameter, d (mm) SiC monofilament NicalonTM HM Carbon HS Carbon SaffilTM SiC whisker
Source: Hull and Clyne, 1996.

Young's modulus E (GPa) 400 190 390 250 300 450

Fracture strength s* (GPa) 2.4 2.0 2.2 2.7 2.5 5.0

Maximum curvature kmax (mm1) 0.08 1.4 1.4 2.7 5.5 22.2

Minimum bend radius (mm) 12.5 0.71 0.71 0.37 0.18 0.045

150 15 8 8 3 1

(see Chapter 3.20, this volume). This problem can be reduced by the introduction of particulate or transverse fibers, although it is also possible in some cases to arrange for the infiltration to take place in the axial direction (Liu et al., 1998). Other processing techniques, including powder metallurgy approaches, have been employed but are often rather unsatisfactory. Carbon fibers are not very popular as MMC reinforcement, primarily as a consequence of interfacial reactions which occur during processing and the galvanic corrosion effects which can take place in service. Chemical reaction problems are severe for aluminum, titanium, and ferrous alloys, but much less pronounced with magnesium, which does not form a stable carbide. Magnesium-based carbon multifilament MMCs have therefore received some attention (Hall et al., 1989; Hall, 1991; Kagawa and Nakata, 1992). There has also been interest in carbon fiber-reinforced zinc alloys, another system in which interfacial reaction is limited. Attempts have been made to protect carbon fibers with a surface coating, such as titanium nitride (Popovska et al., 1997), but in general this is difficult and expensive for multifilaments. There is a particular problem with aluminum alloys, in that the interfacial reaction product, Al3 C4, is hygroscopic, so that graphite fiber-reinforced aluminum tends to undergo rapid corrosion in aqueous environments (see Chapter 3.18, this volume). Although there have been claims (Wendt et al., 1994) that alloys and processing conditions can be identified which lead to carbon fiber reinforced aluminum which is quite corrosion-resistant, the problem has severely limited the development of this type of MMC. While silicon carbide has been a successful reinforcement in particulate- and, to some extent, whisker-reinforced MMCs, there is a shortage of multifilament SiC fibers which are suitable for incorporation into metallic matrices. Multifilaments are available commercially, under tradenames such as Nicalon, which

are derived from polycarbosilane (PCS) precursors in a similar way to the production of carbon fibers from polyacrylonitrile (PAN). However, although these are nominally SiC, they actually contain considerable amounts of free carbon and silica (see Chapter 1.05, Volume 1), leading to excessive reaction with most metallic matrices during processing. There are, however, a number of oxide multifilaments which are fairly resistant to attack by molten metals. Notable among these are polycrystalline alumina fibers, usually composed primarily of the stable a phase (Lavaste et al., 1995), which have been studied for use in aluminum (Folgar et al., 1984; Shetty and Chou, 1985; Isaacs and Mortensen, 1991) and, to a lesser extent, titanium (Warren et al., 1995) matrices. Commercial interest in all such systems has remained low, although there has been some activity concerning reinforcement of intermetallic matrices with oxide multifilaments (see Chapter 3.28, this volume). 3.01.2.3.2 Monofilament MMCs

Monofilaments are large diameter (*100 150 mm diameter) fibers, most of which are produced by chemical vapor deposition (CVD) of either SiC or boron onto a core of carbon fiber or tungsten wire. The details of production methods and microstructural features are given in Chapter 1.04, Volume 1. A consequence of the large diameter is that monofilaments are much less flexible than multifilaments, so that they normally need to be handled as single fibers rather than bundles and precautions are necessary to avoid causing damage by the imposition of sharp curvature during processing operations. This aspect of the handling characteristics of various fibers is illustrated by the data shown in Table 1, which shows calculated maximum curvatures sustainable before fracture occurs. This is related (Hull and Clyne, 1996) to the fracture strength, sf, Young's modulus, E, and diameter, d, of the fiber

A Survey of MMC Types and Developments

11

Figure 5 Photo of a bladed ring (bling), which would replace a substantial disk-plus-attached-blades assembly within a gas turbine engine. Titanium alloy reinforced with SiC monofilaments aligned in the hoop direction of the ring would confer the necessary stiffness, strength, and creep resistance characteristics on the component, which is about 300 mm in diameter (photo courtesy of Rolls Royce plc). kmax 2sf Ed 2

It can be seen that, while whiskers can be bent until their radii of curvature are as small as 50 mm and multifilaments to less than 1 mm, a monofilament is likely to fracture when its radius of curvature approaches 10 mm. This can certainly be a source of inconvenience when manipulating monofilaments, particularly since the lack of flexibility can lead to surface damage which in turn makes fracture more likely if the fiber is bent so as to induce tensile stresses in the damaged region. There are, however, advantages in having such a large diameter. One of these is that interfacial reaction consumes a much smaller proportion of the fiber than would be the case for multifilaments, simply because the interfacial area is much smaller. There is also much greater scope for tailoring the surface chemistry and introducing surface coatings. This can often be done as part of the fiber production process. A lot of effort has gone into the development of coatings on SiC monofilaments for incorporation into titanium and titanium aluminide matrices (see Chapter 3.24, this volume). Virtually all fiber materials react with Ti at elevated temperature, so that use of coated monofilaments is one of the few approaches offering scope for control of this problem. Thick graphitic coatings (which are progressively consumed but prevent defects from forming on the fiber itself) have been popular and there has also been work (Nathan and Ahearn, 1990; Kieschke and Clyne, 1991; Warwick and Smith, 1991; Shatwell, 1995) on various duplex layers, such as TiB2/C and

Y/Y2O3, which are designed to slow interfacial reaction rates down to very low levels. Monofilament-reinforced MMCs are mainly produced by the foilfiberfoil (diffusion bonding) route or by the evaporation of relatively thick layers of matrix material onto the surface of the fiber, followed by hot pressing. Full details of both methods are given in Chapter 3.24, this volume. Work in this area is very much oriented towards titanium and titanium-based matrices. Fortunately, titanium diffusion bonds to itself very readily, mainly because it dissolves its own surface oxide layer at elevated temperature in controlled atmosphere. The evaporation method is slower and more expensive than the foilfiberfoil route, but it does produce material in which the fiber distribution is more uniform. This is an important advantage in view of extensive evidence for various types of MMC that the toughness and ductility are impaired by clustering of the reinforcement, although it should be recognized that matrix plasticity is heavily constrained by the presence of the fibers in this type of MMC, so that toughness values tend to be relatively low even if there is no clustering. There has been a lot of interest in the selective use of monofilament-reinforced titanium for critical components in aeroengines. The presence of the SiC monofilaments confers a dramatic improvement in the creep resistance of titanium (see Chapter 3.14, this volume) and there is also a substantial enhancement to the stiffness. It may also be noted that the resistance to compressive failure and buckling collapse is considerably improved (see Chapter 3.09, this volume). These are all extremely valuable property improvements for titanium in demanding applications within a gas turbine engine. However, such enhancement occurs exclusively or predominantly in the direction of fiber alignment and careful account must be taken of the nature of the imposed stress field, and the effect of internal stresses from differential thermal contraction, when designing the fiber orientation within the component. Nevertheless, it has been concluded that the enhancements in critical properties are such that there could be complete redesign of certain gas turbine components, with dramatic weight savings and consequent benefits. An example is provided by the bladed ring (bling) shown in Figure 5, which would replace a much heavier conventional disk and blade assembly, leading to further weight savings in the bearing configurations needed to support the shaft. Such a component requires stiffness and mechanical stability characteristics which could only be provided by the presence of an efficient hoop-wound reinforcement. It is thought that

12

An Introductory Overview of MMC Systems, Types, and Developments Furthermore, these techniques are rather slow if three-dimensional components rather than thin films are required. Faster production is possible using various diffusion bonding methods and, for cases in which it is acceptable for the ceramic layers to become discontinuous, it may be possible to use roll bonding of thin metal strips with surface oxide layers or to roll layered material down to reduce the layer thickness. Fine layered metal/ceramic structures often have interesting physical and functional properties and these are covered in detail in Chapter 3.12, this volume. They may also exhibit certain attractive mechanical properties, such as a high yield stress. However, coarser scale materials are quite likely to be of more interest for mechanical and structural purposes. This is partly a consequence of factors related to the processing militating against the economic production of very fine scale structures, but in fact certain key mechanical characteristics are often superior with coarser structures. For example, consider the issue of energy absorption during fracture of laminated ceramic/metal materials. For these materials, the toughness is dominated (Shaw et al., 1993; Pateras et al., 1997; Howard et al., 1998; Hwu and Derby, 1999) by the energy absorbed as the metallic layers undergo plastic deformation during necking and rupture. Since the volume of material in which plastic deformation occurs, per unit area of fracture surface, is larger when the layer thickness (hm) is greater, coarser structure have higher fracture energies. Modeling of the energy absorption process leads to an equation of the form
Gtot 1 fm Gc fm Y
 max

monofilament-reinforced titanium alloys could provide the necessary performance, but substantial technical and economic obstacles will need to be surmounted if this is to become viable. 3.01.2.4 Layered MMC Systems

A class of MMCs which has attracted attention recently is that in which the two constituents are in the form of alternate layers of some sort. Such arrangements might range in scale from layer thicknesses of a few nanometers up to macroscopic laminates made by bonding together plates which are several centimeters thick. Many of the mechanical properties of such systems, particularly those relating to elastic behavior, can readily be predicted using the simple concept of either an equal strain (loading parallel to the plane of the layer) or an equal stress (transverse loading) being imposed on both constituents. There is also likely to be relatively high constraint on the plastic deformation of the metallic layers, leading in many cases to high work-hardening rates. In addition to simple bonded layers of monolithic ceramic and metal layers, there has also been interest in other arrangements, such as alternate layers of metal and polymer-based long fiber composite, which is the basis of the so-called ARALL material (Vogelesang and Gunnink, 1983; Bucci et al., 1987), and also the related GLARE material (Schijve, 1993; Vasek et al, 1997). These are covered in detail in Chapter 2.09. One of the attractions of layered systems is that, in comparison with the corresponding fibrous or particulate composites, it is often relatively easy to manufacture components with this geometry. This is also true of ceramic/ceramic layered systems, which can be manufactured by tape casting and sintering techniques (Clegg et al., 1990). There is considerable interest in these ceramic composites, since repeated crack deflection at interfaces can lead to dramatic enhancement of the toughness (Phillipps et al., 1993) as a consequence of the energy absorbed by interfacial debonding. The principles involved are covered in Chapter 4.17, Volume 4. A range of fabrication methods can be used to produce metalceramic laminates. These include vapor deposition or sputtering to produce very thin layered structures, such as are used in a variety of device and electronic applications (see Chapters 3.11 and 3.12, this volume). Particular attention must be paid to the danger of excessive interdiffusion and chemical reaction during deposition for such structures.

hm w 2

where w
0

n d Y

in which Gc is the fracture energy of the ceramic, fm is the metal volume fraction, sY is the effective yield stress of the metal, sn is the nominal stress on the ligament during traction, wmax is the normalized crack opening displacement at which the metal ligament fractures, and the integral w is the area under a plot of sn/sY vs. w. The broad validity of this treatment is confirmed by the predictions and experimental data shown in Figure 6, which refer to AlAl2O3 laminates tested at room temperature (Pateras et al., 1995). Details of the thermomechanical behavior and damage development characteristics of metalceramic laminates are given in Chapters 3.10 and 3.11, this volume.

A Survey of MMC Types and Developments

13

Figure 6 Measured and predicted fracture energy values for failure of metal-ceramic laminates, plotted against the thickness of the metal layers. The experimental data (Pateras et al., 1995) refer to specimens produced with a fixed ceramic layer thickness, so that the metal volume fraction, fm, is different in each case.

3.01.2.5 3.01.2.5.1

Special Category MMCs Metallic foams

It is possible to consider a metallic foam as a composite with metallic and gaseous constituents, i.e., an MMC with a gas as the reinforcing constituent. Indeed, this is a sensible approach to the prediction of many of the properties of metallic foams. If the gas is taken to have a Young's modulus of zero and a very low bulk modulus, then predictions of the elastic properties of the foam, using methods such as those outlined in Chapters 1.13 and 1.14, Volume 1, should be quite reliable. For other properties, such as the thermal conductivity (see Chapter 3.16, this volume), the fact that the reinforcement is a gas becomes relevant, since convection may have a significant effect, particularly in open cell foams (Lu et al., 1998). However, it should be noted that treating the metallic framework as a continuum, within which gaseous inclusions are dispersed, may be rather inappropriate for many purposes, particularly if the gas volume fraction is high (>*80%). It has thus been common to employ a framework for mechanical property prediction which is based on an assembly of structural elements, usually the edges and faces of the cells, taken to have a certain geometrical shape. This may range from a simple cube to more complex polyhedra such as tetrakaidecahedra or pentagonal dodecahedra. The resulting relationships between porosity content and mechanical properties have been studied in considerable detail. These are covered in the text of Gibson and Ashby (1997) and summarized in Chapter 3.31, this volume.

Metallic foams have attracted considerable industrial and scientific interest recently. This has arisen partly because advances in processing technology (Davies and Zhen, 1983; Banhart and Baumeister, 1998b; Gergely and Clyne, 1999), which are summarized in Chapter 3.30, this volume, have led to the prospect of stock and shaped metallic foam components becoming available at prices comparable with corresponding bulk material (on a weight basis) and partly because it has been appreciated that there is potential for the attainment of attractive property combinations, particularly in terms of specific stiffness (Simone and Gibson, 1998a, 1998b) and specific energy absorption (Banhart and Baumeister, 1998a; Fusheng et al., 1998). This has led to the active exploration (Seeliger, 1997; Wood, 1997) of the use of metallic foams for a range of industrial applications, including several concerning automobile components. As it happens, many metallic foams are MMCs in two senses: first because all metallic foams can be treated as constituting a special class of MMC and second because the cell walls are often made up of MMC material. This is the case because, depending on the processing methodology employed, it is often necessary to stabilize the foam against cell coarsening and collapse while the metal is in a liquid or semiliquid state. This is commonly achieved by introducing a dispersion of ceramic into the melt, either as oxide films or as ceramic particles, which in effect raise the viscosity of the melt. Details are given in Chapter 3.30, this volume. It can be seen from Figure 7(a) and (b) that introduction of such ceramic particles can help to retain a fine and uniform cell structure. This is known to result in improved

14

An Introductory Overview of MMC Systems, Types, and Developments as ceramics which have been toughened by the presence of a small proportion of metal, in practice the metallic phase often forms some sort of partially interlinked network, so that they are in effect particulate-reinforced MMCs with an exceptionally high loading of ceramic particles. On the topic of nomenclature, it should be noted that the terms hardmetal and cemented carbide are also commonly used in describing products within this class of material. The differences between these are sometimes unclear and largely arise from historical origins (Mari, 2001). The term hardmetal was originally reserved for the carbides, nitrides, borides, and silicides of the metals of the fourth to sixth group of the Periodic table of elements. Prominent among these are ceramics such as tungsten carbide (WC). These exhibit relatively high thermal and electrical conductivity, but they have mechanical properties, such as high hardness, which are typical of ceramics. The term has come to signify material produced by bonding together such ceramic particles with a metal binder, which is usually an alloy of Co, Ni,or Fe. Sometimes, hard metals based on WC (commonly WC/Co) are called cemented carbides, while those based on TiC (commonly TiCxN1x/MoNi) are termed cermets, to indicate their more ceramic character (e.g., lower electrical conductivity). However, it should be noted that this terminology is not very logical or consistently used. For example, the thermal conductivity of ceramics can be higher than those of metals, as a consequence of efficient heat transport via phonons, so that metallic character is not necessarily implied by a high value. The electrical conductivity of ceramics, on the other hand, is normally negligible compared with metals, so that it is the connectivity of the metallic constituent which will be the most significant factor in determining the apparent electrical resistivity (see Chapter 3.16, this volume. It is possible (Goetzel, 1984) to classify cermets into four main groups, as outlined in the paragraphs below. (i) Carbide-based cermets

Figure 7 Structures of metallic foams produced by a melt route (Gergely and Clyne, 1999). Two cell structures are shown, corresponding to the same production conditions except that the SiC particulate content was (a) 10 wt.% or (b) 20 wt.%. Also shown, (c), is the microstructure of a cell wall for a 20 wt.% SiCp case, illustrating how the presence of the ceramic particles is likely to promote rupture of the wall.

mechanical properties. However, the presence of (relatively coarse) ceramic particles within the cell walls (see Figure 7(c)) probably promotes cell wall rupture and hence inhibits the attainment of high pore content in combination with fine cell diameter. It may also be noted that there are probably deleterious effects on certain mechanical characteristics, particularly concerning plasticity and energy absorption, since some embrittlement of the cell walls would be expected to result when ceramic particles are introduced into them. This is an area requiring further work. 3.01.2.5.2 Cermets

While the term cermet is simply an amalgamation of parts of the words ceramic and metal, so that various materials incorporating both of these constituents could in principle be encompassed, in practice it is normally used (Mari, 2001) to designate an assembly of ceramic particles bonded together by a small amount of a metallic phase. Some authorities (Tinklepaugh and Crandall, 1960) have specified that the ceramic should constitute at least 70% by volume and there should be little solubility between metallic and ceramic phases at the preparation temperature. There is thus a strong case for regarding cermets as a special class of MMCs. While they could be considered

Most cermets in industrial use are of this type. They include cemented carbides based on WC, which are widely used for cutting tools and wear-resistant parts, TiC cermets for high-temperature components in propulsion systems, and chromium carbide (Cr3C2) cermets for applications requiring good corrosion resistance. The metallic binder may be nickel-, molybdenum-, cobalt- or aluminum-based. The system in most common use is WCCo. It has

A Survey of MMC Types and Developments been estimated (Mari, 2001) that 65% of the total world production of tungsten is devoted to the fabrication of cermets for cutting tools. (ii) Oxide-based cermets

15

These include materials comprising particles of Al2O3 in a Cr or CrMo matrix, which is often used when resistance to contact with liquid metals is required, and SiO2 in brass, bronze, or lead, for components to be exposed to high frictional forces. There are also more specialized systems, such as uranium dioxide (UO2) or thorium dioxide (ThO2) in aluminum, stainless steel, or tungsten, which are used as fissile constituents in nuclear reactor fuel elements. (iii) Boride-based cermets

In this case the ceramic is a boride of one of the transition metals. These cermets are stable against attack by reactive metals in the molten or vapor state. For example, a combination of ZrB2 and SiC, in a boron matrix, is resistant against erosion from the propulsion gases of chemical rockets (Goetzel, 1984). (iv) Carbon-containing cermets

These cover materials which contain free graphite, usually present to enhance electrical contact or to provide surface lubrication. The metal binder is usually tin, lead, or zinc. However, the class also encompasses material comprising diamond particles in metallic matrices, which are used in special tools (Goetzel, 1984). There has been interest recently in MMCs containing diamond reinforcement, in the form of either particles (Johnson and Sonuparlak, 1993) or fibers (May et al., 1994), so this is at least an example where similar systems have been examined as cermets and as MMCs. One of the keys to the successful exploitation of cermets in industrial applications concerns their ease of processing. They are commonly produced by blending of ceramic and metallic powders, followed by liquid phase sintering. Similar techniques are employed to produce certain MMCs (see Chapter 3.25, this volume). Typically, the ceramic particles used in cermet production are 110 mm in diameter. Blending involves a milling operation which tends to coat the ceramic particles with metal. This is usually followed by cold isostatic pressing or injection molding to give the required shape and then holding at a suitable temperature under vacuum, inert gas, or hydrogen. During liquid

phase sintering, particle rearrangement occurs (German, 1996), driven by capillarity forces. This may or may not lead to effective densification, depending on the degree of wetting in the system. In some cases, notably for the oxidebased cermets, it is often necessary to impose unixial or hydrostatic pressure in order to eliminate porosity. This is reminiscent of the melt infiltration process used to form MMCs (see Chapter 3.20, this volume), but the very high ceramic content and fine scale of the structure means that in the absence of wetting very high pressure might be necessary to ensure that liquid flows into all the cavities. One approach to the problem of inadequate wetting is deliberately to promote selected chemical reactions during sintering. For example, cermets such as TiAlTiB2 can be produced via the XD process (Christodoulou et al., 1988; Suzuki et al., 1993) by reacting powder blends such as Al, Ti, and B. More details of reactive processing techniques are given in Section 3.01.2.5.3. However, in general the occurrence of pronounced chemical reaction leads to difficulties in the form of uncontrolled volume changes, heat release, and undesirable phase formation. Fortunately, the carbides and nitrides of the hardmetals group are in general thermodynamically more stable than the carbides or nitrides of most binder metals, so that there are few problems of significant chemical attack during sintering. Some metal ion exchange may occur, and in certain cases there is a degree of dissolution of the ceramic in the metal and reprecipitation on solidification. Such processes probably have the effect of raising the bond strength (see Chapter 3.05, this volume), but for most systems there is little danger of thick, brittle interfacial reaction zones forming during sintering. Of course, this is largely because cermet systems have been chosen partly with this chemical compatibility in mind, since there are not the same constraints of economics and matrix mechanical properties which operate when choosing the constituents for MMCs: the aluminium-, titanium-, and magnesium-based matrices commonly used for MMCs are attractive in these terms, but are very reactive in the liquid state, which often leads to problems during this type of processing. One point which has become well established (Subrahmanyam et al., 1986; Ramnath and Jayaraman, 1989; Nerz et al., 1991; Khan and Clyne, 1996; Li et al., 1996; Jacobs et al., 1997; Khan et al., 1998) is that even cermet systems which are quite stable during sintering, such as WCCo, can undergo various chemical reactions if exposed, even for a short time, to very high temperatures (well above the fusion point

16

An Introductory Overview of MMC Systems, Types, and Developments with cermet technology. For example, the promotion of higher interfacial bond strengths is probably desirable for many MMCs and study of the mechanisms by which this is achieved in certain cermets may be instructive. Similarly, it is possible that the understanding of areas such as processing, residual stresses, constraint effects, and fracture mechanisms, acquired during study of MMCs, may lead to further improvements in the economics of production or the mechanical performance characteristics of cermets. Irrespective of the details of the dependence on particle size and binder content, it can be seen from Figure 8 that very attractive combinations of room temperature hardness and toughness are obtainable with sintered cermets. Of course, for applications such as cutting tools, the behavior at high temperature is also very important. The binder will tend to soften as the temperature is raised, but this will not necessarily impair the hardness very much, particularly at relatively low binder levels, since it is heavily constrained by the presence of the surrounding particles. Cermets thus exhibit excellent hot hardness (markedly superior to high-speed tool steels) in the range 700 1000 8C. However, at temperatures above this, grain boundary sliding of the carbides starts to occur, leading to extensive plastic deformation. This is pronounced in WCCo cermets. The TiC cermets, commonly having TiCNMoNi formulations, are more resistant to this effect and are thus commonly used in high-speed cutting operations which may raise the temperature of the tool above 1000 8C. For such applications, their exceptional hot hardness offsets the rather lower hardness and toughness these systems exhibit at room temperature when compared with WCCo. 3.01.2.5.3 Reactively processed MMCs

of the matrix). This happens, for example, during thermal spraying of cermet coatings, which is in extensive industrial use. Such reactions are particularly likely in an oxidizing environment, so that the chemistry of the flame and surrounding atmosphere may be important. It has also been shown (Khan et al., 1998) that excessive reaction leads to degradation of coating properties, particularly the resistance to abrasion and erosion. It may be noted that, even in the absence of noticeable chemical reaction during sintering, the characteristics of the binder within cermets may be rather different from that in the monolithic state. The binder is commonly in an effectively elastically strain hardened state (Sigl and Fischmeister, 1988), probably as a consequence of the high triaxial constraint and the plastic deformation imposed during cooling. It may also exhibit metastable phases (Exner, 1979). Nevertheless, in general the presence of the binder does confer a substantial increase in toughness, as well as facilitating the consolidation process. It has been shown (Read, 1955) that a thin, continuous film of binder separating the ceramic particles is beneficial in terms of toughness enhancement, probably because it inhibits crack propagation from occurring exclusively through the ceramic constituent. The mechanical properties of a typical cermet such as sintered WCCo can cover quite a wide range, depending primarily on the Co content, as well as the details of the manufacturing method. This is illustrated by the data in Figure 8. Both hardness and abrasion resistance fall off with increasing Co content and also with increasing particle size. Density varies inversely with Co content and is independent of particle size. Fracture toughness (Lueth, 1974; Pickens and Gurland, 1978; Bouaouadja et al., 1994; Ravichandran, 1994), on the other hand, rises with increasing Co content and with WC particle size. These trends are rather similar to those exhibited by conventional particulate MMCs (see Chapters 3.07 and 3.19, this volume), except for the effect of particle size on toughness. This is probably because cracks tend to bifurcate as a result of the presence of larger ceramic particles in cermets, while such particles tend to act as defects in MMCs. There may, however, be a dependence on the likelihood of large particles becoming cracked during processing, and possibly also on the interfacial bond strength and on the degree of clustering in MMCs (see Chapters 3.04, 3.05, and 3.06, this volume). It seems likely that research into the optimization of particulate MMCs could benefit from improved familiarity with the experience and expertise associated

An area which has attracted considerable interest is that of MMCs produced by the passage of a molten metal infiltration front through a packed ceramic bed of some sort, with or without associated chemical reaction. Such materials often bear a marked resemblance to the cermets described in the preceding section, but a distinction can be drawn in terms of certain differences in processing conditions and in the alloy and ceramic systems which are commonly involved. Much of the work done in this area has its origins in attempts to facilitate the melt infiltration process, particularly with aluminum-based melts. When liquid metal is injected into a relatively fine array of particles or fibers, the

A Survey of MMC Types and Developments

17

Figure 8 Dependence on Co content of (a) hardness, (b) abrasion resistance, (c) density, and (d) fracture toughness of sintered WCCo cermets. These data were compiled by Khan (1997) from the work of Santhanam et al. (1990) and Lueth (1974).

applied pressure needed to generate the necessary meniscus curvature at the infiltration front can be very large (Clyne and Mason, 1987; Nourbakhsh et al., 1989; Mortensen and Wong, 1990; Young, 1991), typically at least several MPa: this issue is covered in Chapter 3.20, this volume. Such high pressures are difficult to apply safely using pneumatic systems and often require the introduction of hydraulic ram systems, adding considerably to processing cost and complexity. Considerable effort has been invested in the search for methods by which this requirement can be reduced or eliminated. Ideally, wetting and/or chemical reaction occurring at the infiltration front would be such as to promote spontaneous infiltration without the need for external application of pressure. While it has proved difficult to promote rapid infiltration under these conditions, MMC products can be made in this way with good near net shape characteristics, particularly when the ceramic content is high. The basic problem when attempting to infiltrate liquid aluminum into Al2O3 or SiC preforms is one of poor wetting. The introduction of wetting agents such as K2ZrF6 has been found (Schamm et al., 1991) to be quite effec-

tive, but such additions are cumbersome and inconvenient. After various trials, however, it became clear that wetting could be strongly promoted by chemical reactions arising from the introduction of magnesium into the melt and nitrogen into the surrounding atmosphere (preferably in the absence of oxygen). Such conditions are experimentally fairly easy to arrange. Compounds such as aluminum nitride, AlN, and MgAl2O4 spinel are then formed (Aghajanian et al., 1989a, 1989b, 1991; Scholz and Greil, 1991; Scholz et al., 1993). While the details of the thermodynamic and kinetic characteristics involved are complex, the important point is that the reactions involved take place at locations and at rates which are such as to promote spontaneous penetration of an Albased melt into an array of ceramic particles or fibers. Commonly, the melt employed is a binary or multicomponent AlMg alloy and the process is carried out under a nitrogen atmosphere. A schematic depiction of the process is shown in Figure 9. Terms such as the PRIMEX process have been applied to this and related procedures, notably by the Lanxide Corporation, which holds a number of patents in the area.

18

An Introductory Overview of MMC Systems, Types, and Developments

Figure 9 Schematic depiction (Singh, 1999) of the PRIMEX process developed by Lanxide Corporation for production of MMCs by preform infiltration.

Figure 10 Optical micrograph (Singh, 1999) showing the microstructure of an Al70 vol.% SiC MMC produced using the PRIMEX process.

Figure 11 Components of a chuck for securing Si wafers during processing of electronic devices, made from PRIMEX processed Al70 vol.% SiC MMC material (Singh, 1999).

Unfortunately, the reaction kinetics and local melt flow characteristics are such that even when infiltration does occur spontaneously, it tends to be relatively slow. Typically it might take many minutes, or even hours, for a pre-

form to become fully infiltrated. This problem, together with the fact that the presence of relatively high reaction product contents may impair mechanical properties, has meant that reactive processing is not extensively used for the production of conventional low-cost MMCs for applications such as pistons or brake disks. However, the process has clear attractions for more specialized applications, particularly those requiring high ceramic contents, which are rather difficult to produce by standard powder processing methods. For example, the Al70 vol.% SiC MMC shown in Figure 10 was produced using the PRIMEX process, in the form of the components shown in Figure 11. These are part of a specialized chuck used in the electronics industry for wafer handling, which has demanding stiffness, conductivity, and thermal expansivity requirements (Singh, 1999). Note the bimodal size distribution of the SiC particles shown in Figure 10, which facilitates the generation of high ceramic contents in the initial powder compact. Techniques have also been developed in which different types of chemical reaction are promoted during MMC production. These can be distinguished from PRIMEX-type processes on the grounds that the reaction products are often important constituents in the final microstructure, rather than being incidental byproducts of a reaction primarily designed to promote wetting. This is similar in principle to the XD processes mentioned in the previous section as being used to generate cermet constituents. For example, directional oxidation or nitridation has been used to infiltrate a carbide

A Survey of MMC Types and Developments preform with a reactive metal in oxidizing or nitriding atmospheres, leading to systems such as Al2O3SiC or AlNSiC, usually with appreciable levels of residual aluminum also being present in the final product. In some systems it has been found (Xiao and Derby, 1993) that interfacial dopants can be beneficial. Such techniques have been termed (Newkirk et al., 1987; Urquhart, 1991) DIMOX methods by the Lanxide Corporation. The final metal volume fraction in such materials is often quite low and there are certainly grounds in many cases for classifying them as cermets. 3.01.2.5.4 Intermetallic matrix composites (IMCs)

19

There are a number of intermetallic compounds which have quite attractive combinations of properties, particularly in terms of their high temperature resistance to creep, oxidation (Grabke et al., 1992), and wear. The major drawback concerning their use for structural purposes is commonly their rather poor fracture toughness at relatively low temperatures. They therefore share many characteristics with ceramics, although the partially metallic nature of the interatomic bonding leads to relatively high values for properties such as electrical conductivity and there is often at least some scope for promoting dislocation mobility and designing the microstructure so as to toughen the material somewhat. The production of composites with intermetallic matrices (Kumar and Bao, 1994) thus usually has the primary aim of raising the fracture energy of the material relative to that of the unreinforced matrix. There are two broad approaches to the toughening of intermetallics by the introduction of a reinforcing constituent, which is usually in fibrous form. The first is to add a relatively tough and ductile fiber (Deve and Maloney, 1991; Rao et al., 1992), commonly a high melting point metal. There may be problems of interfacial reaction during processing and coated fibers have been explored (Deve et al., 1990) in attempts to control this. Basically, the toughening in these systems comes simply from energy absorbed during plastic deformation of the reinforcement: fibers are preferred since, depending on their geometrical arrangement, it is difficult for crack propagation to occur without fiber deformation. The second approach is to introduce a reinforcement which does not itself have a high toughness, but raises the fracture energy by promoting interfacial bonding and consequent crack deflection. This mechanism can operate even when the reinforcement is in particulate

form, although with both fibers and particles the potential for toughening tends to be lower than with ductile reinforcement. The detailed mechanics of toughening by crack deflection with a brittle matrix, and optimization of processing (Stoloff and Alman, 1990) so as to promote microstructural features which enhance the toughness via this mechanism, are covered in several chapters within Volume 4 of the present series. There is thus a degree of uncertainty about whether IMCs are more appropriately considered as a subset of MMCs or as part of the CMC family. The intermetallic systems which have received most attention are the aluminides, particularly Ti3Al, Ti2AlNb, TiAl, Ni3Al, and NiAl. Full details of these and the IMC systems produced with them are given in Chapter 3.28, this volume. There is also interest in certain metal silicides, such as MoSi2, which have excellent oxidation resistance. In general, the commercial exploitation of IMCs has remained at a low level, although there have been extensive demonstrator trials. This may be partly because in many cases optimization of the microstructural features exhibited by the intermetallics themselves is not yet complete. This is quite a complex issue, with substantial changes in properties often arising from rather subtle microstructural modifications, and the presence of reinforcement often tends to disturb the evolution of microstructure during processing. There are also problems of added cost and complexity, given that most intermetallics are themselves rather expensive and difficult to process. Viable applications for IMCs may nevertheless emerge in due course, but further development is clearly necessary. 3.01.2.5.5 Graded MMC coatings and structures

A class of material which has attracted attention recently is that in which metallic and ceramic constituents are incorporated into planar structures with continuous or discrete changes in the proportions of the two. These are often in the form of thin coatings on massive substrates, although it is also possible to build layered structures up into bulk laminated material of the type outlined in Section 3.01.2.4 and covered in detail in Chapters 3.10, 3.11, and 3.12, this volume. Reviews are available which cover processing aspects (Mortensen and Suresh, 1995) and also thermomechanical characteristics (Suresh and Mortensen, 1997) of such systems. In many cases, the aim is to combine desirable features of a ceramic, such as good wear resistance or thermal insulation properties, with the

20

An Introductory Overview of MMC Systems, Types, and Developments type of processing route employed. In general, melt-based routes are economically more attractive for high-volume production than powder-based routes, since they allow easier handling of bulk quantities, removal of impurities, etc. Procedures have been developed (see Chapter 3.21, this volume) which allow the efficient incorporation of ceramic particulate into aluminum melts and MMC material produced in this way is available commercially at a price which is not enormously higher than for conventional aluminum (perhaps greater by a factor of the order of 2 or 3). Of course, such prices are sensitive both to economies of scale during production and to market forces and the perceived demand (Kevorkijan, 1999). Powder route Al-based particulate MMCs, on the other hand, tend to be significantly more expensive but often offer superior properties to melt route material, as a consequence of the improved microstructural control which can be achieved. Further development work on optimization of processing conditions, so as to reduce the costs, while maintaining the necessary microstructural quality, is likely to be useful for both melt and powder routes over the coming years. Other types of MMC fall into two broad categories in terms of economic aspects. First there are various short fiber and high-volume content particulate MMCs, which are becoming increasingly competitive for certain relatively specialized applications. These are commonly made by some type of infiltration or reactive processing technique. Examples are described in Chapters 3.20 and 3.25, this volume, and in Sections 3.01.2.5.2 and 3.01.2.5.3. It would also be possible to include metallic foams (see Chapter 3.30, this volume) in this category. Raw material costs are again fairly low, but processing is in general more complex and costly than for competing materials. On the other hand, the simultaneous generation of material and component (which is characteristic of many types of composites) may offer scope for processing effects not readily obtainable with conventional materials, such as tailored distributions of reinforcement or penetration of material into prepared cavities as it is formed. Advances are being made on process economics in these areas and this will probably assist in the commercialization of a range of MMC products. Finally, there are some types of MMC which are inherently expensive to produce. These include most MMCs incorporating long fiber reinforcement, particularly monofilaments. The cost of the fibers themselves is often relatively high. Furthermore, processing is in general rather cumbersome and expensive, since it is often difficult to envelop long fibers with

toughness and adhesion to a substrate which are often more readily achievable with metallic materials. There may also be objectives related to the establishment of control over residual stress distributions by using constituents with selected thermal expansivities or creep characteristics. An example is provided by thermal barrier coating systems (Miller, 1997), which conventionally consist of a metallic bond coat and a ceramic top coat. Attempts have been made (Jamarani et al., 1992) to improve the mechanical stability of such coatings by grading the composition between entirely ceramic at the free surface and entirely metallic at the substrate, although in this particular case there is often a problem of enhanced oxidation rates of the metallic constituent with such structures. However, in general it is often found that tailored distributions of metal/ceramic proportions through the thickness of a graded structure can lead to improvements in the achievable property combinations exhibited by the system (Suresh and Mortensen, 1997).

3.01.3

SOME CURRENT PROCESSING AND PERFORMANCE ISSUES FOR MMCS Process Economics

3.01.3.1

A recurrent concern when exploring the viability of MMC usage is that of cost, particularly as it relates to processing. Details of how costs of MMC products may be broken down and analyzed are given in Chapter 3.29, this volume. The raw material costs depend very much on the type of MMC being considered. Most of the applications being explored and developed for relatively high-volume usage are oriented towards particle-reinforced aluminum MMCs. There have been very few clear indications to date that major benefits are to be expected from the use of specially formulated aluminum alloys as MMC matrices. Cost issues relating to supply of matrix material will therefore be similar in nature to those of conventional unreinforced aluminum. The use of particulate reinforcement does add to the raw material cost, although for standard SiC and Al2O3 powders this is not a very dramatic effect. If more specialized types of powders are required, in terms of particle size, shape, or composition, then the cost will in general rise and could become highly significant. It should, however, be recognized that, while the raw material costs of some types of MMC are often rather similar to those of competing monolithic materials, the processing costs tend to be somewhat higher. This is sensitive to the

Some Current Processing and Performance Issues for MMCs metal effectively while maintaining a uniform spatial distribution. Vapor deposition of metal onto fibers, followed by hot pressing, can be very effective in this respect for titanium-based composites (see Chapter 3.24, this volume), but this is inherently rather slow and requires high capital cost equipment. Such composites may find use in very high value-added applications (see Chapter 3.27, this volume), but it is difficult to see how major reductions could be effected in the processing costs while maintaining the necessary microstructural quality. 3.01.3.2 Microstructural Quality Control

21

A recurrent issue for manufacturers of MMC material and products concerns identification of appropriate quality control parameters, particularly relating to microstructural features. Some of these can be drawn from standard practice for unreinforced metals. For example, levels of impurity or porosity in the matrix can be fairly easily measured and specification maxima for these can be defined. While there may be problems in establishing the precise effects of these parameters on the performance of particular types of MMC, there are no conceptual difficulties in setting specifications of this type. However, it is becoming clear that there are certain specific microstructural characteristics peculiar to the presence of the reinforcement which can have a strong influence on the behavior of MMCs. For example, not only is the size, shape, and volume fraction of the reinforcing constituent often significant, but also its spatial distribution and the nature of the interfacial bonding. Moreover, there are two separate types of problem associated with such attributes: first their effect on the performance of the material may be incompletely understood and second the way in which they are best characterized may be far from obvious. It is now fairly well established that inhomogeneous distributions of reinforcement, particularly the presence of clustering (giving rise to ceramicdepleted and ceramic-rich areas), can have a deleterious effect on the mechanical properties. This is particularly true for the ductility and fracture toughness. This area is covered in detail in Chapter 3.06, this volume. That chapter also summarizes the approaches which can be adopted to characterization of the severity of clustering. Systematic analysis requires the application of tesselation procedures, in which a metallographic section is divided up into a series of polygons, each containing a single reinforcing particle or fiber. In the present state of MMC development, it would be unusual for such procedures to be systematically used as a

quality control tool, although with the advance of automatic image analysis facilities they are becoming easier to implement and they may be used in the future. A similar problem relates to the interfacial bond strength, except that it is quite difficult to even approximately estimate this via a study of the microstructure. In general, it is now clear that the properties of most MMCs are enhanced by the promotion of a high interfacial bond strength and toughness (Clyne, 1996). Further details are given in Chapter 3.05, this volume. Such good bonding is certainly present in successful high ceramic volume fraction MMCs, such as cermets (see Section 3.01.2.5.2), and there are many indications that good bonding is beneficial for lower ceramic content MMCs (see Chapters 3.04, 3.05, 3.06, and 3.07, this volume). It is also fairly clear that the presence of brittle interfacial reaction products, which tend to lower the toughness of the interfacial region, is normally undesirable. While this can be monitored by microstructural examination, quantitative evaluation of bond strength requires some sort of mechanical interrogation of the material. A variety of techniques have been developed (Le Petitcorps et al., 1989; Kallas et al., 1992; Roman and Aharonov, 1992; Kalton et al., 1994, 1998) for doing this, most of which are much more easily applied to long fiber systems (particularly monofilament-reinforced) than to particulate or short fiber MMCs. The development of a simple and reliable method for assessing the interfacial strength in particulate MMCs has proved to be rather difficult, but there is little doubt that its availability would be of potential value for quality control purposes. 3.01.3.3 Damage, Ductility, and Fracture

Perhaps the area of greatest concern in the performance of MMCs is that there is a tendency for the ductility and toughness to be relatively low. In fact, MMCs can exhibit fracp ture toughness values of well above 20 MPa m, which is certainly adequate for many critical load-bearing applications, and ductilities of at least 5% or so are routinely obtainable. However, many MMCs exhibit lower values than these. Furthermore, the toughness of nominally similar types of MMC, perhaps made by different routes or with different processing conditions, can vary significantly. Of course, materials engineers have long been familiar with such variations in the context of metallic alloys, in which microstructural changes during production or as a consequence of thermomechanical heat treatments can lead to substantial

22

An Introductory Overview of MMC Systems, Types, and Developments rapid degradation in water (Saxena et al., 1987). In other cases there are galvanic effects which promote aqueous corrosion in MMCs. Protection is, however, usually quite straightforward providing the characteristics of the system concerned are well understood. Details are given in Chapter 3.18, this volume. Similarly, high-temperature oxidation and surface degradation of MMCs, which are covered in Chapter 3.17, this volume, usually follow similar trends to those shown by the parent metal. In some cases, however, problems arise from interfacial chemical reaction at high temperature, which can be accentuated in long fiber MMCs by penetration along the interface of species from the environment. While environmental degradation characteristics of MMCs are usually rather similar to those of the corresponding metals, the presence of the reinforcement can cause difficulties during recycling. This is usually tackled for metals by melting the scrap, often while virgin metal is being generated at the same time. In general, it is not feasible to separate ceramic particles or fibers from the molten matrix in such a way that they can be reused and, indeed, some problems can arise in simply removing the reinforcement so that the metal can be recovered. Details of this are given in Chapter 3.30, this volume. It may, however, be noted that there is a possibility of MMC scrap being recycled into metallic foam material, which is starting to find a number of industrial applications. This is also outlined in Chapter 3.30, this volume: the mechanical properties of metallic foams are described in Chapter 3.31, this volume. 3.01.4 MMC CONTACTS, NETWORKS, AND SOURCES OF INFORMATION

changes in mechanical properties. However, while the mechanisms responsible for these changes are well established and understood for unreinforced metals, cause and effect are in some cases rather unclear for MMCs. In particular, while there has certainly been extensive work in these areas (Flom and Arsenault, 1989; Lewandowski et al., 1989; Povirk et al., 1991; Manoharan and Lewandowski, 1992; Singh and Lewandowski, 1993; Marchal et al., 1996; Maire et al., 1997; Wilkinson et al., 1997; Murphy et al., 1998), there is still a degree of uncertainty surrounding the effects of reinforcement size, degree of clustering, interfacial bond strength, residual stress distribution, and matrix toughness on the ductility and fracture toughness of MMC systems. Details are given in Chapters 3.06, 3.07, and 3.08, this volume. Broadly speaking, it is clear that relatively fine particle size, uniform spatial distribution of reinforcement, and strong interfacial bonding all lead to enhanced fracture toughness. These effects can be qualitatively explained in terms of favored crack paths within the material and absorption of energy occurring predominantly by matrix plastic deformation. For fibrous MMCs, while situations can arise with relatively brittle metallic matrices when fiber pullout contributes significantly to the net energy absorption during fracture (Dellis et al., 1991), this is in general rather unusual and factors controlling the matrix plasticity constitute the most important consideration for the toughness of all types of MMCs. Work remains to be done on various aspects of this, including the precise role of residual stresses within the matrix (see Chapter 3.13, this volume) and the effect of having bimodal distributions of particle size. 3.01.3.4 Environmental Stability and Recycling

Issues related to the whole life cycle of products have assumed considerable importance as awareness of pollution and the value of recycling strategies has increased. There is thus some interest in the factors which affect the generation of by-products during manufacture of MMCs, the degradation occurring in service, and the scope for recovery and recycling of scrap MMC material. In general, the environmental stability of an MMC is commonly rather similar to that of the parent metal matrix. Of course, there are a number of effects which can occur in particular cases which may alter this picture. For example, certain reinforcements can react with aluminum matrices during processing to form aluminum carbide, which is strongly hygroscopic and leads to

A wide range of MMC applications have been explored and, in some, cases taken to full commercial exploitation. Some details are given in Chapters 3.26 and 3.27, this volume. The current state of development of MMCs is such that it is common for researchers, technologists, designers, and engineers to have a keen interest in the possibilities offered by this class of materials, but to experience some difficulties in obtaining key technical information or material suitable for R&D work. In fact, there is now a wide network of possible sources, both for information and for MMC material of various types. Under these circumstances it is often helpful to use the Internet to track down appropriate sources. In Table 2 a list is given of website addresses which may be useful in this context. These include independent organiza-

Mmc Contacts, Networks, and Sources of Information


Table 2

23

A selection of organizations involved in the dissemination of information concerning MMCs. Type of organization European Union-funded network Trade society Trade society Professional institute Consultancy network Website address mmc-assess.tuwien.ac.at/ http://www.adtek.u-net.com/ http://www.almmc.com/ http://www.instmat.co.uk /index.htm http://metalexpertsintl.com/ Contact e-mail adress mmc assess @ewkmmc.tuwein.ac.at bdodd@adtek.u-net.com jsimon5100@aol.com Admin@materials.org.uk yodonna@aol.com

Organization name MMC-Assess Adtek Aluminium Metal Matrix Composites Consortium Institute of Materials (MMC Committee) Metal Experts International

Table 3 A selection of commercial suppliers of MMC material. Organization name 3M Allied Signals Composites Inc. Lanxide Corporation Chesapeake Composites Corp. Alyn Corporation Atlantic Research Corporation Textron Systems Ceramics Process Systems Ametek Advanced Forming Technology Brush Wellman Duralcan Metal Matrix Cast Composites Inc Inco SPP DWA Aerospace Metal Composites ERG Corporation Stillen Website address http://www.3m.com/market/industrial/mmc/ http://www.dlcomposites.com/ http://www.lanxide.com/ http://www.chesapeakecomposites.com/ http://www.alyn.com/ http://www.arcmaterials.com/ http://www.systems.textroncom/mm.htm http://www.alsic.com/ http://www.ametek.com http://pcc-aft.com/tech/mmc/mmc.html Types of MMC currently available Long fiber reinforced Al Reactively-processed and melt-infiltrated systems Reactively-processed and melt-infiltrated systems Infiltrated Al- and Mg-based composites with high volume fractions of submicron particles Range of advanced products Ti MMCs Ti MMCs AlSiCp, for electronic applications AlSiCp, for electronic applications AlSiCp, for electronic applications

http://www.brushwellman.com/www/Technical Be- and Cu-based MMCs for /Ceramics/EMaterials.html electronic applications http://bricad.com/aluminium/dur/ Cast Al-based particulate MMCs http://www.mmccinc.com/ Cast Al-based particulate MMCs http://www.incospp.com /incospp/new_metal_composites.htm http://207.178.131.112/ http://www.amc-mmc.co.uk/ Cast Al-based particulate MMCs Powder route Al-based particulate MMCs Powder route Al-and Fe-based particulate MMCs MMC foams MMC brake pads Cermet-type MMCs B4C-reinforced MMCs for nuclear fuel storage AlSiCp MMCs

http://www.ergaerospace.com/corp.htm http://www.stillen.com/Stillen%20 Folder/Stillen%20Docs/brakepads.html Sinter Metal http://www.sinter-metal.com/english Technologies /indexen.htm Advanced Refractory http://www.art-inc.com/products Technologies Inc. /carbides.html Goodfellow http://www.goodfellow.com/static /A/AL41.html

tions and networks, as well as associations of commercial firms. A number of the websites listed in Table 2 provide tools designed to facilitate searches for MMC material or component suppliers. A selection of MMC supplier website addresses is

given in Table 3. While this is by no means an exhaustive list, it does indicate that there is now a relatively wide range of commercial sources for MMC material. This in turn illustrates the recent growth in industrial interest and suggests that MMC technology may be approaching

24

An Introductory Overview of MMC Systems, Types, and Developments


J. D. Birchall, Trans. J. Br. Ceram. Soc., 1983, 82, 143 145. J. D. Birchall, D. R. Stanley, M. J. Mockford, G. H. Pigott and P. J. Pinto, J. Mat. Sci. Lett., 1988, 7, 350 352. N. Bouaouadja, M. Hamidiuche, H. Osmani and G. Fantozzi, J. Mat. Sci. Lett., 1994, 13, 1719. R. J. Bucci, L. N. Mueller, L. B. Vogelesang and J. W. Gunnink, J. Metals, 1987, 39, A58. G. R. Cappelman, J. F. Watts and T. W. Clyne, J. Mat. Sci., 1985, 20, 21592168. T. Christman, A. Needleman and S. Suresh, Acta Metall., 1989, 37, 30293050. L. Christodoulou, P. A. Parrish and C. R. Crowe, in `High Temperature/High Performance Composites', Reno, NV, eds. F. D. Lemkey, S. G. Fishman, A. G. Evans and J. R. Strife, MRS, Pittsburgh, PA, 1988, pp. 2934. W. J. Clegg, K. Kendall, N. M. Alford, J. D. Birchall and T. W. Button, Nature, 1990, 347, 455457. T. W. Clyne, Key Eng. Mats., 1996, 127-131, 8198. T. W. Clyne and J. F. Mason, Metall. Trans., 1987, 18A, 15191530. T. W. Clyne and P. J. Withers, `An Introduction to Metal Matrix Composites', Cambridge University Press, Cambridge, UK, 1993. G. S. Daehn and G. Gonzlez-Doncel, Met. Trans., 1989, 20A, 23552368. G. J. Davies and S. Zhen, J. Mat. Sci., 1983, 18, 1899 1911. M. A. Dellis, J. P. Keustermans, F. Delannay and J. Wegria, Mat. Sci. andEng., 1991, 135A, 253257. H. E. Deve, A. G. Evans and R. Mehrabian, MRS Symp. Proc., 1990, 170, 3338. H. E. Deve and M. J. Maloney, Acta. Metall. Mater., 1991, 39, 22752284. T. Donomoto, K. Funatani, N. Miura and N. Miyake, `Ceramic Fibre Reinforced Piston for High Performance Diesel Engines', SAE paper 830 252, 1983. I. Dutta, J. D. Sims and D. M. Seigenthaler, Acta. Met., 1993, 41, 885908. H. E. Exner, Int. Metals Reviews, 1979, 4, 14973. Y. Flom and R. J. Arsenault, Acta Met., 1989, 37, 2413 2423. F. Folgar, W. H. Kreuger and J. G. Goree, Ceram. Eng. Proc., 1984, 5, 643648. H. Fusheng, Z. Zhengang and G. Junchang, Metall. Trans., 1998, A29, 24972502. V. Gergely and T. W. Clyne, in `Porous and Cellular Materials for Structural Applications', eds. D. S. Schwartz, D. S. Shih, A. G. Evans and H. N. G. Wadley, MRS, San Francisco, 1998, pp. 139144. V. Gergely and T. W. Clyne, in `MetFoam '99', Bremen, Verlag, eds. J. Banhart, M. F. Ashby and N. Fleck, Verlag MIT Publishing, Berlin, 1999, pp. 8389. L. J. Gibson and M. F. Ashby, `Cellular Solids', Cambridge University Press, Cambridge, UK, 1997. R. M. German, `Sintering Theory and Practice', John Wiley, New York, 1996. C. G. Goetzel, `Metals Handbook: Powder Metallurgy', ASM, Materials Park, OH, 1984. F. H. Gordon, S. P. Turner, R. Taylor and T. W. Clyne, Composites, 1994, 25, 583592. H. J. Grabke, M. Brumm and M. Steinhorst, Mat. Sci. & Tech., 1992, 8, 339344. I. W. Hall, J. Mat. Sci., 1991, 26, 776781. I. W. Hall, G. R. Brewer and A. Magata, J. Mat. Sci. Lett., 1989, 8, 343345. S. H. Hong, O. D. Sherby, A. P. Divecha, S. D. Karmarkar and B. A. MacDonald, J. Comp. Mat., 1988, 22, 102123. S. J. Howard, S. K. Pateras and T. W. Clyne, Mat. Sci. & Tech., 1998, 14, 535541.

maturity. While this may be true in certain respects, it is clear that the development of improved products, and indeed of the background understanding needed for this to take place efficiently, is far from complete. Promulgation of reliable and up-to-date information about MMCs will thus be very important for their industrial development over the next decade or so. Use of the web will probably be central to this dissemination. Of course, the list of addresses given in Tables 2 and 3 will date rapidly and it should be recognized that many of them will probably cease to be of value within a relatively short period. Nevertheless, some of them should at least remain useful for several years as starting points for web searches. Background knowledge about the processing and performance characteristics of MMCs will date much less quickly than information about products and suppliers, so it is anticipated that the current volume will serve as a comprehensive and authoritative reference for all aspects of MMCs over an extended period.

ACKNOWLEDGMENTS The author is grateful to Professor Tony Kelly for extensive discussions and collaborations concerning composites in general, evolution of the present volume on MMCs, and the contents of the current introductory chapter in particular. Thanks are also due to Dr. D. Mari, of the EPFL in Switzerland, for his cooperation in providing a pre-publication copy of a review article of his concerning cermets.

3.01.5

REFERENCES

M. K. Aghajanian, J. T. Burke, D. R. White and A. S. Nagelberg, SAMPE Q., 1989a, 20, 4346. M. K. Aghajanian, N. H. MacMillan, C. R. Kennedy, S. J. Luszcz and R. Roy, J. Mat. Sci., 1989b, 24, 658670. M. K. Aghajanian, M. A. Rocazella, J. T. Burke and S. D. Keck, J. Mat. Sci., 1991, 26, 447454. M. F. Ashby, Acta. Met., 1993, 41, 13131335. C. Badini, F. Marino and A. Tomasi, Mat. Sci. & Eng., 1991, 136A, 99107. J. Banhart and J. Baumeister, J. Mat. Sci., 1998a, 33, 14311440. J. Banhart and J. Baumeister, in `Porous and Cellular Materials for Structural Applications', eds. D. S. Schwartz, D. S. Shih, A. G. Evans, and H. N. G. Wadley, MRS, San Francisco, CA, 1998b, pp. 121132. A. Bardal and R. Hier, in `Proceedings of the 12th Ris International Conference on Metal Matrix CompositesProcessing, Microstructure & Properties', eds. N. Hansen, D. J. Jensen, T. Leffers, H. Lilholt, T. Lorentzen, A. S. Pedersen, O. B. Pedersen and B. Ralph, Ris National Laboratory, Roskilde, Denmark, 1991, pp. 205210. C. Y. Barlow and N. Hansen, Acta Metall. Mater., 1991, 39, 19711979.

References
D. Hull and T. W. Clyne, `An Introduction to Composite Materials', Cambridge University Press, Cambridge, UK, 1996. K. L. Hwu and R. Derby, Acta Mater., 1999, 47, 545 563. J. A. Isaacs and A. Mortensen, Met. Trans., 1991, 23A, 12071219. L. Jacobs, M. Hyland and M. D. Bonte, in `10th National Thermal Spray Conference', Indianapolis, ed. C. C. Berndt, ASM, Materials Park, OH, 1997. F. Jamarani, M. Korotkin, R. V. Lang, M. F. Ouellette, K. L. Yan, R. W. Bertram and V. R. Parameswaran, Surf. Coat. Technol., 1992, 54/55, 5863. W. B. Johnson and B. Sonuparlak, J. Mater. Res., 1993, 8, 11691173. Y. Kagawa and E. Nakata, J. Mat. Sci. Lett., 1992, 11, 176178. M. N. Kallas, D. A. Koss, H. T. Hahn and J. R. Hellman, J. Mat. Sci., 1992, 27, 38213826. A. F. Kalton, S. J. Howard, J. Janczak-Rusch and T. W. Clyne, Acta Mater., 1998, 46, 31753189. A. F. Kalton, C. M. Ward-Close and T. W. Clyne, Composites, 1994, 25, 637644. C. G. Kang, B. C. Koh, J. Kim and K. H. Kim, in `Composites: Design, Manufacture and Application (ICCM-8)', eds. S. W. Tsai and G. S. Springer, SAMPE, Covina, CA, 1991, p. 17I. H. Kaufmann and A. Mortensen, Metall. Trans., 1992, 23A, 20712073. V. M. Kevorkijan, J. Metals, 1999, 51(11), 5458. M. Khan, `Materials Science and Metallurgy', Cambridge University Press, Cambridge, UK, 1997. M. S. A. Khan and T. W. Clyne, in `Thermal Spray: Practical Solutions for Engineering Problems. Proceedings of the 9th Nat. Thermal Spray Conference', ed. C. C. Berndt, ASM, Novelty, OH, 1996, pp. 113122. M. S. A. Khan, T. W. Clyne and A. J. Sturgeon, in `Thermal Spray: A United Forum for Scientific and Technological Advances. Proceedings of the 1st United Thermal Spray Conference', ed. C. C. Berndt, ASM, Indianapolis, 1998, pp. 681690. R. R. Kieschke and T. W. Clyne, Mat. Sci. & Eng., 1991, 135A, 145149. K. S. Kumar and G. Bao, Comp. Sci. & Technol., 1994, 52, 127150. V. Lavaste, M. H. Berger, A. R. Bunsell and J. Besson, J. Am. Ceram. Soc., 1995, 30, 42154225. Y. Le Petitcorps, R. Pailler and R. Naslain, Comp. Sci. & Technol., 1989, 35, 207214. A. P. Levitt, `Whisker Technology', Wiley, New York, 1970. J. J. Lewandowski, C. Liu and W. H. Hunt, Mat. Sci. & Eng., 1989, A107, 241255. C. J. Li, A. Ohmori and Y. Harada, J. Thermal Spray Technol., 1996, 5, 6973. H. N. Liu, H. Miyahara and K. Ogi, Mat. Sci. & Technol., 1998, 14, 292298. D. J. Lloyd, Int. Mat. Rev., 1994, 39, 123. D. J. Lloyd, H. Lagace, A. McLeod and P. L. Morris, Mat. Sci. & Eng., 1989, A107, 7380. T. J. Lu, H. A. Stone and M. F. Ashby, Acta Mater., 1998, 46, 36193635. R. C. Lueth, in `Fracture Mechanics of Ceramics', ed. R. C. Bradt, Plenum, New York, 1974, pp. 791806. E. Maire, D. S. Wilkinson, J. D. Embury and R. Fougeres, Acta Mater., 1997, 45, 52615274. M. Manoharan and J. J. Lewandowski, Mat. Sci. & Eng., 1992, 150A, 179186. Y. Marchal, F. Delannay and L. Froyen, Scripta Mater., 1996, 35, 193198. D. Mari, in `Encyclopaedia of Materials: Science and Technology', ed. A. Mortensen, Elsevier, Oxford, 2001, in press.

25

J. F. Mason, C. M. Warwick, P. J. Smith, J. A. Charles and Clyne, T. W. Clyne, J. Mat.Sci., 1989, 24, 3934 3946. H. Matsubara, Y. Nishida, M. Yamada, I. Shirayanagi and T. Imai, J. Mat. Sci. Lett., 1987, 6, 13131315. P. W. May, C. A. Rego, R. M. Thomas, M. N. R. Ashfold, K. N. Rosser, P. G. Partridge and N. M. Everitt, J. Mat. Sci. Lett., 1994, 13, 247249. R. A. Miller, J. Thermal Spray Technol., 1997, 6, 3542. A. Mortensen and S. Suresh, Int. Mat. Rev., 1995, 40, 239265. A. Mortensen and T. Wong, Metall. Trans., 1990, 21A, 22572263. J. D. Muhly, in `The Beginning of the Use of Metals & Alloys', ed. R. Maddin, MIT Press, Cambridge, MA, 1988, pp. 220. A. M. Murphy, S. J. Howard and T. W. Clyne, Mat. Sci. & Techol., 1998, 14, 959968. V. C. Nardone and J. R. Strife, Metall. Trans., 1987, 18A, 109114. M. Nathan and J. S. Ahearn, Mat. Sci. & Eng., 1990, A126, 225230. J. Nerz, B. Kushner and A. Rotolico, in `High Performance Ceramic Films and Coatings', ed. P. Vincenzini, Elsevier, Amsterdam, 1991, pp. 2736. M. S. Newkirk, H. D. Lesher, D. R. White, C. R. Kennedy, A. W. Urquhart and T. D. Claar, Ceram. Eng. Sci. Proc., 1987, 8, 879885. S. Nourbakhsh, F. V. Liang and H. Margolin, Metall. Trans., 1989, 20A, 18611866. S. R. Nutt, J. Am. Ceram. Soc., 1988, 71, 149156. A. Parvizi-Majidi, in `Structure and Properties of Composites', ed. T.-W. Chou, VCH, Weinheim, 1993. S. K. Pateras, S. J. Howard and T. W. Clyne, Key Eng. Mats., 1997, 127-131, 127131. S. K. Pateras, M. C. Shaw, W. J. Clegg, A. C. F. Cocks and T. W. Clyne, in `10th International Conference on Composite Materials (ICCM-10)', eds. A. Poursartip and K. Street, Whistler, BC, Canada, Woodhead Publishing, Cambridge, UK, 1995, pp. 11271136. J. Pelleg, Mat. Sci. & Eng., 1999, 269A, 225241. A. J. Phillipps, W. J. Clegg and T. W. Clyne, Acta Metall. et Mater., 1993, 41, 805817. J. R. Pickens and J. Gurland, Mat. Sci. & Eng., 1978, 33, 135142. N. Popovska, H. Gerhard, D. Wurm, S. Poscher, G. Emig and R. F. Singer, Mats. & Design, 1997, 18, 239242. G. L. Povirk, A. Needleman and S. R. Nutt, Mat. Sci. & Eng., 1991, 132A, 3138. V. Ramnath and N. Jayaraman, Mat. Sci. & Technol., 1989, 5, 382388. K. T. V. Rao, G. R. Odette and R. O. Ritchie, Acta Met. Mater., 1992, 40, 353361. K. S. Ravichandran, Acta Metall. Mater., 1994, 42, 143 150. R. H. Read, Ph.D. Thesis, Pennsylvania State University, (1955). H. Ribes and M. Suery, Scripta Met., 1989, 23, 705709. I. Roman and R. Aharonov, Acta Metall. Mater., 1992, 40, 477485. A. Santhanam, P. Tierney and J. L. Hunt, `Properties and Selection: Non-Ferrous Alloys and Special Purpose Materials', ASM, Materials Park, OH, 1990. M. Saxena, O. P. Modi, A. H. Yegneswaren and P. K. Rohatgi, Corros. Sci., 1987, 27, 249256. S. Schamm, R. Fedou, J. P. Rocher, J. M. Quinisset and R. Naslain, Metall, 1991, 22A, 21332139. J. Schijve, in `Fracture '93', eds. J. P. Bailon and J. I. Dickson, Engineering Materials Advisory Service, Warley, UK, 1993, pp. 320, chap. 285. H. Scholz and P. Greil, J. Mat. Sci., 1991, 26, 669677. H. Scholz, R. Gunther, J. Rodel and P. Greil, J. Mat. Sci. Lett., 1993, 12, 939942.

26

An Introductory Overview of MMC Systems, Types, and Developments


J. R. Tinklepaugh and W. B. Crandall, in `Cermets', eds. J. R. Tinklepaugh and W. B. Crandall, Reinhold Publishing Corporation, New York, 1960. S. P. Turner, R. Taylor, F. H. Gordon and T. W. Clyne, Int. J. Thermophysics, 1996, 17, 239252. A. Urquhart, Mat. Sci. & Eng., 1991, A144, 7582. A. W. Vasek, J. Polak and V. Kozak, Mat. Sci. & Eng., 1997, A234, 621624. L. B. Vogelesang and J. W. Gunnink, in `4th International SAMPE Conference', Bordeaux, France, eds. G. Jube, A. Massiah, R. Naslain and M. Popot, SAMPE, Covina, CA, 1983, pp. 8192. J. Warren, D. M. Elzey and H. N. G. Wadley, Acta Metall. et Mater., 1995, 43, 36053619. C. M. Warwick and J. E. Smith, in `12th Ris International Symposium on Metals & Materials Science', Ris, Denmark, eds. N. Hansen, D. J. Jensen, T. Leffers, H. Lilholt, T. Lorentzen, A. S. Pedersen, O. B. Pedersen and B. Ralph, Ris National Laboratory, Roskilde, Denmark, 1991, pp. 735740. R. G. Wendt, W. C. Moshier, B. Shaw, P. Miller and D. L. Olson, Corrosion, 1994, 50, 819826. D. S. Wilkinson, E. Maire and J. D. Embury, Mat. Sci. & Eng., 1997, 233A, 145154. J. T. Wood, in `Metal Foams', eds. J. Banhart and H. Eifert, MIT-Verlag Publishing, Stanton, DE, 1997, pp. 3135. P. Xiao and B. Derby, J. Am. Ceram. Soc., 1993, 77, 17611770. R. M. K. Young, Mat. Sci. & Eng., 1991, A135, 1922.

H.-W. Seeliger, in `Metal Foams', eds. J. Banhart and H. Eifert, MIT-Verlag, Berlin,1997, pp. 7989. R. A. Shatwell, Mat. Sci. Techol., 1995, 10, 552557. M. C. Shaw, D. B. Marshall, M. S. Dadkhah and A. G. Evans, Acta Metall Mater., 1993, 41, 33113322. H. R. Shetty and T.-W. Chou, Metall. Trans., 1985, 16A, 853864. L. S. Sigl and H. F. Fischmeister, Acta Metall., 1988, 36, 887897. A. E. Simone and L. J. Gibson, Acta Mater., 1998a, 46, 39293935. A. E. Simone and L. J. Gibson, Acta Mater., 1998b, 46, 21392150. J. R. Singh, Future Fab., 1999, 2. P. M. Singh and J. J. Lewandowski, Met. Trans., 1993, 24, 25312543. M. Skibo, P. L. Morris and D. J. Lloyd, in `Cast Reinforced Metal Composites', eds. S. G. Fishman and A. K. Dhingra, ASM, Metals Park, OH, 1988, pp. 257 261. N. S. Stoloff and D. E. Alman, MRS Bull., 1990, 15, 47 53. J. Subrahmanyam, M. P. Srivastava and R. Sivakumar, Mat. Sci. & Eng., 1986, 84, 209214. S. Suresh and A. Mortensen, Int. Mats. Rev., 1997, 42, 85116. M. Suzuki, S. R. Nutt and R. M. Aiken, Mat. Sci. & Eng., 1993, A162, 7382. M. Taya and R. J. Arsenault, `Metal Matrix Composites: Thermomechanical Behaviour', Pergamon, Oxford, 1989.

Copyright # 2000 Elsevier Science Ltd. All rights reserved. No part of this publication may be reproduced, stored in any retrieval system or transmitted in any form or by any means: electronic, electrostatic, magnetic tape, mechanical, photocopying, recording or otherwise, without permission in writing from the publishers.

Comprehensive Composite Materials ISBN (set): 0-08 0429939 Volume 3; (ISBN: 0-080437214); pp. 126

Вам также может понравиться