Sei sulla pagina 1di 21

Composites

ELSEVIER

PII:

SO266-3538(96)00167-4

Science and Technology 57 (1997) 415-435 0 1997 Elsevier Science Limited Printed in Northern Ireland. All rights reserved 0266-3538/97/$17.00

DRY

SLIDING

WEAR

OF ALUMINIUM REVIEW

COMPOSITES-A

R. L. Deuis;

C. Subramanian

& J. M. Yellupb

Ian Wark Research Institute, University of South Australia, SA 5095, Australia hCSIRO, Division of Manufacturing Technology, Adelaide Laboratory, SA 5012, Australia
(Received 25 February 1996; revised 30 August 1996; accepted 14 October 1996)

Abstract
Aluminium-silicon alloys and aluminium-based metalmatrix composites have found application in the manufacture of various automotive engine components such as cylinder blocks, pistons and piston insert rings where adhesive wear (or dry sliding wear) is a predominant process. Materials possessing high wear resistance (under dry sliding conditions) are associated with a stable tribolayer on the wearing surface and the formation of fine equiaxed wear debris. For adhesive wear, the influence of applied load, sliding speed, wearing surface hardness, reinforcement fracture toughness and morphology are critical parameters in relation to the wear regime encountered by the material. In this review contemporary wear theories, issues related to counterface wear, and wear mechanisms are discussed. Other areas of research relevant to adhesive wear of Al-5 alloys and aluminium composites containing discontinuous reinforcement phases, such as the role of the reinforcement phase, are also presented. 0 1997 Elsevier Science Limited

Keywords: metal-matrix composites, silicon alloys, dry sliding wear

aluminium-

1 INTRODUCTION Metal-matrix composites (MMCs) exhibit the ability to withstand high tensile and compressive stresses by the transfer and distribution of the applied load from the ductile matrix to the reinforcement phase. In situ composite structures, such as aluminium containing silicon in amounts above the solubility limit, are formed by solidification. However, the compositions and relative amounts of the two phases are limited to a narrow range, controlled by growth kinetics and equilibrium conditions. Artificial composite structures do not exhibit these limitations in composition. These MMCs are fabricated by the addition of a reinforcement phase to the matrix by the use of several techniques such as powder metallurgy, liquid metallurgy and squeeze-casting. The reinforcement phase is generally one of the following: continuous boron or graphite fibres, or hard particles such as Sic
415

and A&O, in discontinuous particulate or whisker morphology. The volume fraction of reinforced particles or whiskers is generally within the range lo-30%. Aluminium alloys, such as the 2000, 5000, 6000 and 7000 alloy series, are the most commonly utilised materials in composite fabrication. Aluminium composites are widely employed in the aerospace industry.lm3 Hyper-eutectic Al-S1 based composites such as A356 (A1,7Si,0_3Mg) that contain A1203, ZrO, particles4 or Sic particles are used in the fabrication of automotive engine components. Wear resistance and operating properties of aluminium cast diesel pistons are enhanced by the use of aluminium-based composite piston ring inserts. These composite inserts are reinforced with A&O3 whiskers or a combination of these whiskers (12 vol.%) and carbon fibres (9 vol.%). Aluminium-based composites have also been considered as substitute materials for use in the fabrication of brake rotors, pistons, cylinder liners and cylinder heads. A comprehensive review of the current and potential application for cast aluminium matrix composites in the automotive industry has been provided by Rohatgi. In this review, the terms fibre, whisker and particle reinforcement are denoted by the symbols f, w and p respectively. Unless otherwise indicated quantities expressed as a percentage imply weight percentage. 2 ADHESIVE WEAR

Adhesive wear is defined as the transfer of material from one surface to another during relative motion by a process of solid-phase welding or as a result of localised bonding between contacting surfaces. Particles that are removed from one surface are either permanently or temporarily attached to the other surface.X 2.1 Classical theory According to Rabinowicz, adhesive wear occurs when surfaces slide against each other and the pressure between the contacting asperities is sufficiently high enough to cause local plastic deformation. Hardness

416

R. L. Deuis, C. Subramanian,

J. M. Yellup

of a material determines the real area of contact between asperities of contacting materials. Asperity hardness is therefore considered to be more important than bulk hardness. The adhesive wear theory stated by Archard] defined wear volume as a function of sliding speed, normal load and material hardness. However, this theory ignored the effect of the materials microstructure on wear and was limited to idealised sliding conditions. This theory was based on a mechanism of adhesion at the asperities and the material removal process was related to a cohesive failure of asperities. The processes of crack nucleation and subsequent growth were disregarded. With the assumption that wear particles could be described as hemispherical particles of the same radius as the contact area, Archard developed the following expression for wear rate, W (volume of material worn):
W=KdP 3H

where K = wear coefficient, d = sliding distance, P = applied normal load and H = bulk hardness of the material. Archard concluded that the wear rate was proportional to the applied load (assuming that the average size of the contact areas and the wear particles were constant) and that the wear rate was independent of the apparent area of contact. The theory predicted that enhanced wear resistance was associated with increase in hardness. However, this conclusion has not always been found to be valid. The proposed proportionality between the applied load and wear rate was not always observed in sliding wear systems in studies by Archard and Hirst.12 2.2 Delamination wear theory Sub proposed that at low sliding speeds, wear debris formation could be described by a delamination wear theory. Wear processes such as adhesive wear, fretting and fatigue were all related to this same mechanism. Suh stated that wear occurred by the following sequential steps: Cyclic plastic deformation of surface layers by normal and tangential loads. Crack or void nucleation in the deformed layers at inclusions or second-phase particles. Crack growth nearly parallel to the surface. Formation of thin, long wear debris particles and their removal by extension of cracks to the surface. The rate-determining mechanism of wear showed dependence on the metallurgical structure. When sub-surface deformation controlled the wear rate, hardness and fracture toughness were both considered to be major influencing factors. Jahanmir and Suhi4 showed that for microstructures

containing hard second-phase particles, if sufficient plastic deformation occurred during sliding wear, crack nucleation was favoured at these particles. In this situation, where inter-particle spacing is an important variable, crack propagation controlled the wear rate. Void formation was primarily attributed to the plastic flow of the matrix around these hard particles. Void formation occurred very readily around the hard particles but crack propagation occurred very slowly. The depth at which the void nucleation was initiated and the void size tended to increase with increased friction coefficient and applied load. Subs delamination theory proposed that voids only nucleated at a defined depth below the sliding wear surface. Void formation was related to the hydrostatic pressure which existed directly under a contact region. Voids therefore nucleated below a level where hydrostatic pressure was not large enough to suppress their formation and above a depth where plastic deformation was sufficient to nucleate voids around the hard particles. The existence of this critical depth dictated the resultant dimensions of the wear particle debris, especially the thickness of plate-like particles. The existence of a depth below the worn surface where void formation and eventual shear instability occurred, was confirmed by Rosenfield. In a study of dry sliding wear of dispersionhardened alloys, delamination wear theory was identified as a major wear process.h Hardness and friction coefficient played a major role in the overall wear process. The wear resistance decreased with increased volume fraction of the oxide phase, even when hardness was increased. Crack propagation was considered to be the wear-rate controlling factor. Cracks were initiated at the particle/matrix interface or by fracture of the particles. For crack nucleation at particle/matrix interfaces, the following conditions were necessary:
l

Tensile stress across the interface should exceed the interfacial bond strength. Elastic strain energy released upon decohesion of the interface should be sufficient to account for the surface energy of the crack created.

Argon suggested that for second phase particles to be a significant factor in the nucleation of a void or micro-crack, they must possess a diameter in excess of 2.5 pm in order to satisfy the energy conditions. Therefore, for good wear resistance it was found to be desirable to have an alloy reinforced with a large volume fraction of very small coherent particles. Zhang and Alpas examined the extent of plastic deformation that occurred below the contact surfaces as a result of dry sliding wear in an Al-7Si alloy. In this study, profiles of both the shear strain and microhardness distributions were determined in the worn region. For dry sliding, it was found that the

Dry sliding wear of aluminium composites-a

review

417

magnitude of the plastic strains and the depth of the heavily deformed sub-surface zone increased with both sliding distance and applied load. The wear process was observed to follow Suhs delamination wear theory. Delamination cracks formed by the growth and coalescence of voids at a critical depth (=lO-20pm) below the wearing surface. Zhang and Alpas emphasised this critical depth was a result of a balance between plastic strain that enhanced void growth and the hydrostatic pressure that opposed it. These findings supported earlier work reported by Alpas and Zhang.20 In this earlier study, the dry sliding wear behaviour of aluminium alloy A356 reinforced with SiCp (O-20 vol.%) was assessed for a range of loads (l-150 N) and sliding velocities load range (0*16-O+ m s-). For an intermediate (lo-95 N), the predominant wear mechanism was identified to be delamination wear. Crack nucleation within the wearing MMC was believed to be associated with void nucleation around the SIC particles. Cracks were observed to propagate at a depth of lo-20 pm below the surface and resulted in the formation of plate-like wear debris. In reference to the matrix alloy, crack nucleation was reported to be associated with void nucleation around Si particles. Delamination wear and the associated nucleation of voids at the SiCp/matrix interface during the dry sliding of an MMC pin against a steel counterface was also reported by Venkataraman and Sundararajan2 2.3 Friction Bowden and Tabor proposed that the two main contributing factors to friction generated during sliding wear could be described by an adhesion term and a ploughing term.22 A new theory for friction was reported by Suh and Sin2 in which a third term was introduced, namely deformation. These factors are summarised below: .
l l

adhesion of flat regions of the sliding surfaces; ploughing by wear particles and hard asperities; deformation of the surface asperities.

The relative contributions of these components depended on the conditions of sliding, the materials involved and the environment. In general, the contributions to the overall friction coefficient by ploughing and the deformation of asperities were greater than that by the adhesion term.2 For the determination of the friction coefficient for the dry sliding of an Al-Si-Cu alloy against an AISI 52100 steel counterface (Fe,l.3-1.6Cr,0.95-1.1C,O.20.35Si,0.25-0.45Mn,0.025P,O.O25S), Blau24 employed a rule of mixture approach. The coefficient of friction for the composite, Il~ix, was given by the following expression:
Pmix fSiPuSi +fAlPN (2)

where .fsi = area fraction

of Si phase on the sliding

psi = friction coefficient of the Si phase, of Al phase on the sliding surface, and pa, = friction coefficient of the Al phase. This relationship predicted the resultant friction coefficient with reasonable accuracy. A limitation of this approach is that it can predict only the initial run-in behaviour, since as sliding progresses a compacted layer with different microstructure forms on the wearing surface. Rana and Stefanescu25 studied the friction coefficient existing between an Al-l.S%Mg alloy reinforced with SiCp and a tool steel counterface. The value of the friction coefficient decreased with an increase in volume fraction of SIC particles. When a constant volume fraction of SIC was maintained, a decrease in Sic size resulted in a reduced coefficient of friction. In this study, the coefficient of friction was found to be independent of the sliding velocity, though this variable range was too narrow (0.11 and 0.21 m s-) for the relationship to be accepted without further investigation. A study of the coefficient of friction and wear on AI-Li alloys reinforced with 10 wt% SiCp (13 and 30 ,um) using scratch testing was conducted by Zhang et a1.2h Scratch tests were performed at both room temperature and 200C using a scratch speed of 6 mm so and a load range of l-20 N. A diamond pyramidal indenter with an apex angle of 136 was used. Friction coefficient for the composite containing 30p.m SiCp was found to be independent of load. However, the coefficient of friction for the 13 pm reinforced material increased with the applied load. The ratio of penetration depth to particle diameter2 appeared to significantly influence the wear rate and the friction property. The wear at elevated temperature was larger and increased more notably with load due to MMC softening and was associated with a greater indenter penetration depth. At 200C matrix cracking was also more extensive and produced more discontinuous chips that increased the contact area of the indenter resulting in a larger coefficient of friction. Saka et a1.28 examined friction and wear in Cu reinforced with Al,O,p (size 45-53 pm, lo-40 vol.%). For steady state wear, the friction coefficient decreased with increase in alumina content. A similar observation was reported by Rana and Stefanescu, as previously mentioned. This reduction in friction coefficient was attributed to the higher reinforcement/counterface contact area encountered as volume fraction increased. Zhang et a1.29 stated that the friction coefficient of an MMC in scratch testing was made up of the following contributions: ploughing, adhesion and particle fracture. Friction coefficient increased with increasing reinforcement volume fraction but was independent for the scratch speeds (l-10 mm SK) and applied indenter load (l-20 N). The material evaluated in this study was Al 6061 alloy reinforced with
fA, = area fraction

surface,

418

R. L. Deuis, C. Subramanian,

J. M. Yellup

SiCp (10, 2Ovol.% and 1.8 pm) and A1203p (10, 20 vol.% and 4.5, 8.8 ,um). A model to describe the wear volume, V, produced from a single scratch was presented:
V=ApL=

fLP
k,&+k,v+00

(3)

where A,, = cross-sectional area of scratch groove, k, and k2 = experimental constants, f = coefficient of friction, V, = reinforcement volume fraction, Ki, = MMC fracture toughness, d = reinforcement particle size, P = normal load, Hv = Vickers microhardness and L = sliding distance. This theoretical model assumed the following: no particle debonding, uniform distribution of reinforcement phase, an average reinforcement particle size and a rigid diamond indenter. In reference to this model, wear rate increased proportionally with increasing sliding distance, load and coefficient of friction. Increases in the Vickers hardness or the ratio of Kit/d* resulted in a reduction in volumetric wear rate. This model exhibited good correlation with the experimental data as displayed in Fig. 1. 2.4 The tribolayer Jiang et a1.30 reviewed the mechanism involved in dry sliding wear of like metals (nickel alloy/nickel alloy sliding couple) and commented that the observed transition from severe to mild wear after a defined sliding time, was attributed to the formation of a stable, compacted layer (tribolayer). Wear debris formed this layer by a process of transfer from one surface to another, which resulted in the comminution and consolidation of these plastically deformed and oxidised particles into a hard, protective layer that reduced the overall wear rate. It was reported that this transition in the wear regime was more significant at elevated temperatures, where the wear scar developed

a very smooth tribolayer. These researchers identified the transition temperature required for the tribolayer to form. The material could be transferred back and forth several times during sliding and eventually produce wear debris particles. It was suggested that the formation of these wear particles could be a direct result of their work-hardening ability. It was further suggested that a critical transfer layer thickness existed for a given sliding situation.* At this critical value, wear debris particles were thought to form by delamination at or near the interface between the transferred material and the base material. Subramanian3 reported that for an Al-12.3%Si alloy sliding against a copper counterface at low speed (0.1 m s-) and high pressure (>l*O MPa), a tribolayer formed on each wear surface. This layer was composed of a mechanical mixture of the Al alloy and copper. He further concluded that the formation of the tribolayer depended on sliding wear parameters such as applied pressure and sliding speed. In a wear mechanism map developed by Subramanian,34 a tribolayer was favoured in regimes A and B (see Fig. 2). However, for conditions characterised by high sliding speeds and high pressures, it was suggested that fewer transfer and back transfer cycles occurred. Iwai et a1.35 noted that a tribolayer was formed on an Al 2024 alloy reinforced with SiCw after a sliding distance of only 50 m (40N load and sliding speed O-1 m s-). During dry sliding of Al-Si alloy/AISI 4340 (Fe, O-38-0*43C, 0.2-0*35Si, O-6-0.8Mn, 0*04P, 0.4S, 0*7-0*9Cr, 0.2-0=3Mo, 1.652aONi) steel systems, it was implied by Antoniou et al.6 that a finely dispersed amphorous iron oxide formed on the wearing surfaces. This was due to the oxidation of a significant proportion of the steel counterface. This phase helped stabilise the tribolayer on the alloy surface by pinning dislocations. Several points regarding this tribolayer phenomenon were highlighted by Heilmann et al.:31
.

. .

transfer layers were common in both dry and lubricated sliding wear processes; transfer layers developed very early, even before loose wear debris was noted; the composition of these layers consisted of an intimate (mechanical) mixture of materials derived from both sliding materials; the loose wear debris had the same structure and composition as the transferred layer.

Normal Fig. 1.

load, N

Scratch groove cross-sectional area as a function of indenter load for Al/Al,O, MMC (experimental: (m) 10 and (0) 2Ovol.%; theory: (0) 10 and (0) 2Ov01.%).~

It should be noted that both the classical adhesive and delaminated theory of wearI wear theory neglect the formation of the tribolayer in their treatment of wear. Rice et al. characterised the sub-surface zones of materials subjected to sliding and impact wear. Three material zones were identified and are shown in Fig. 3:

Dry sliding wear of aluminium composites-a

review

419

.
l

Zone l-represents the undisturbed base material. Zone 2-describes the deformed intermediate region of the base material. At the Zone l/2 interface, elastic-plastic deformation was noted. No constituents from the counterface or environmental reaction products were evident. Zone 3-this region, including the worn surface and usually containing counterface material and any oxide formed, has been labelled by other researchers as the tribolayer.

Fine cquiaxad pnrticle formation (region a)

Compact dolsmination

Crack and void formation were generally associated at the Zone 213 interface and dictated the dimensions of the wear debris formed, as previously discussed.13 The extent and compositional features of these sub-surface zones were found to depend on the conditions of sliding wear, material and environment. Rice et aZ.37 also indicated that these sub-surface zones developed quickly under dry sliding wear conditions. In relation to these highly strained regions (Zones 2 and 3), a standard technique was developed for estimation of the plastic strain distribution.38 In this method the effective deformation was calculated by determining the change in grain size dimension arising from the severe wear occurring. The effective deformation, 6, was given by:

(region b)

&AL

c[3]

Plastic delamination (region c)

Gross matcrlal transfer

(region e)

Fig. 2. Wear mechanism map for aluminium alloys and schematic drawing of the sub-surface structure developed in Al-Si alloys as a result of different wear conditions depicted in this map.

where D = original grain size and c = deformed grain size. Values for c and D were determined by the line intercept method. Application of this method in relation to dry sliding of an A356 alloy against 52100 steel, revealed shear strains greater than five were present at a depth of 10 pm below the worn surface. It was proposed that the flow stresses inherent in Zones 2 and 3 could be calculated from indentation hardness measurements.3 In relation to a fully work-hardened material, the corresponding yield stress, Y, was given by the following expression:

contact surface

direction of count&ace

-W

Fig. 3. Diagram of the sub-surface zones found beneath wearing surfaces, modified from Rice et al.

where P is the indentation hardness. Examination of the subsurface beneath the worn surfaces of Al reinforced with varying fractions of SiCp (O-40 vol.%, size 2.5 pm) subjected to adhesive wear resulted in the identification of distinct layers4 On the worn surface of MMCs a mechanically mixed layer (MML) was present. This layer exhibited a hardness approximately 6 times that of the bulk composite. No MML layer was evident on the worn surface of aluminium. Beneath this layer (but present as the surface layer on the worn aluminium) was the shear layer. Within this shear layer high shear gradients existed. This shear layer was composed of two sub-layers identified as a soft layer and a

420

R. L. Deuis, C. Subramanian,

J. M. Yeilup

hardened layer. The soft layer was located beneath the MML formation and was caused by the generation of voids at the SiCp/matrix interface as a result of high shear strains. These voids prevented efficient load transfer from the matrix to the particles and lowered the MMC strength. Within this region, the flow stress decreased with increasing plastic shear strain. Adjacent to this soft layer was the hardened layer. Within this layer the flow stress increased with increasing strain. The next region was termed the plastically deformed layer and was associated with bulk plastic deformation. Beneath this layer was the undeformed material. The thickness of each layer was dependent on both the SiCp volume fraction and the applied load. These subsurface layers observed in the dry sliding of these MMCs were similar to the subsurface regions described by Rice et a1.37 Comparing these two studies the MML layer could by described as the tribolayer and the shear layer as the refinement layer. These researchers explained the formation of the MML layer with the following rationale. Initially the MMC material experienced bulk deformation and then shear deformation. With increasing deformation the reinforcement particles at the wearing surface fragmented and the number of voids nucleated at the SiCp/matrix interface increased. When the void density reached a critical value, shear instability, as postulated by Rosenfield, was initiated at local subsurface regions. This resulted in the occurrence of turbulent plastic flow, during which time iron and iron oxide debris was mixed in with the surface MMC material, resulting in the formation of the MMC layer. Venkataraman and Sundararajan40 stated that the presence of this MML layer most probably controlled the wear rate. The formation of wear debris was directly related to delamination wear active within this region. 2.5 Aluminium-silicon
2.5.1 Addition of silicon alloys

linearly with applied pressure but was independent of sliding velocity. The value of the friction coefficient was found to be insensitive to applied pressure, Si content and sliding velocity. The fact that no transition in wear mechanism was observed with increased pressure, as reported by Sarkar,43 could be due to the narrow range explored (0.105-1.733 MPa). These findings were contradicted in a study published by Clarke and Sarkar,44 where it was reported that the wear resistance of Al-Si alloys improved with Si content up to a near-eutectic composition. Furthermore it was indicated that Si additions improved seizure resistance, with the optimum Si content corresponding to the near-eutectic value as indicated in Fig. 4. Reddy et a1.4s investigated the wear and seizure behaviour of Al-Si alloys containing up to 23% Si, using a pin-on-disc rig and a wide load range (15-200 N). It was suggested that seizure occurred at a definitive temperature which was labelled as the critical seizure temperature. Additions of Si (>7 %) increased the seizure load and this temperature value. Eyre46 commented that the etched surface of Al-Si alloys exhibited anti-wear and anti-seizure characteristics. This was a result of the protruding Si phase contacting the steel counterface and thereby avoiding contact between the aluminium matrix and the counterface. This etched surface minimised the smearing of aluminium on the wearing surface and therefore improved frictional performance.47 Howell and Ba114#proposed that the addition of Sic particles (20 vol.%) to A357 (A1,6*2Si,0*42Fe,0.24Mg) alloy supported and protected the matrix from the counterface material and minimised the establishment of severe adhesive surface shear strain that was associated with the unreinforced matrix.

8k,

In a review paper concerning topographical features of worn Al-Si alloys in dry sliding, Clarke and Sarkar4 made the following statements:
1.

2.

The mutual transfer of material between the wearing Al-Si and the steel counterface appears to be a feature of all wear regimes and becomes more significant as load increases. The transition from mild to severe (metallic) wear is associated with the existence of a delamination wear process.
12 16 20 Percentage Sillcon Fig. 4. Wear
alloys rate as a function of Si additions for Al-Si subjected to dry sliding against a hard steel bush, applied load (kgf): 1 (0), 1.5 (A),2 (O), 2-S (o).~~

Although Pramila Bai and Biswas42 reported that Si additions (4-24%Si) improved wear resistance of aluminium, no relationship between wear rate as a function of Si content was found. Wear rate increased

Dry sliding wear of aluminium composites-a

review

421

Modi4Y postulated that reinforcement phases reduced the area fraction of the matrix in contact with the counterface. This would lead to relatively smaller temperature increases at the sliding surface of the composite, compared to that of the alloy, and therefore explain improved seizure resistance. This observation was supported in a dry sliding wear study conducted by Narayan et a1.50 In this study, an Al 2024 alloy reinforced with 15 vol.% A1203p (18 pm) was shown to have better seizure resistance (in the peak-aged condition) than the matrix alloy. Both materials were manufactured using a liquid metallurgy technique. It was also noted that the composite material that was further processed by extrusion exhibited a lower wear resistance compared to the matrix alloy. This was attributed to reinforcement damage incurred in the extrusion technique. Reddy et aZ.l further stated that the frictional force and the shear resistance of AI-Si alloys, were the two main controlling factors determining the initiation of seizure. For a given alloy and sliding temperature, seizure was initiated at a sub-surface depth where the shear stress equalled the resistance to shear. The presence of silicon was said to provide strengthening of this sub-surface region and to help inhibit bulk shear. Narayan et a/. reported that seizure was associated with the following features: a rapid increase in wear rate; transfer of MMC pin material to the counterface; and increased noise and vibration. 2.6 Aluminium-based
2.61 Reinforcement composites size and fraction

The following factors were considered to be significant in influencing wear rate of composites? second phase particle dimension, interparticle spacing, and particle/matrix interfacial bond strength. It was stated that wear rate decreased with an increase in alumina volume fraction. The predominant wear mechanism was identified as delamination, due to poor particle/matrix interfacial bonding. One of the earliest papers that reported the tribological properties of composites reinforced with various particulate types was that of Sato and Mehrabian.* Matrix alloys of Al-4Cu,A1,4Cu, 0.75Mg and 6061 were reinforced with SiC, TIC, Si3N4, A1203, silica sand, MgO, mica, glass beads and B4N. The particulate size range (0.06-840 pm) and volume fraction (l-30%) investigated were extensive. Adhesive wear tests (pin-on-disc) and the associated coefficient of friction data for a sliding velocity range of 0.05-0.46 m s- were tabulated. In general, the friction coefficient decreased with increased sliding velocity for all the composite combinations. For composites reinforced with 210% of either Sic, TIC, Si3N4, AlzOs, or silica sand, the wear rate was less than that of the unreinforced matrix. The correlation of particulate volume fraction and size with these

tribological properties by Sato and Mehrabiar? was unfortunately not conducted in a very systematic manner. Therefore, the nature of the relationship between sliding wear and reinforcement volume fraction and size remained undefined in this study. Surappa et a1.53 noted that aluminium reinforced with 5 % alumina possessed an adhesive wear rate comparable to that of Al-11.8Si or AI--S1 hypereutectic alloys. Other work published by the same workerss4 involving Al, Al-11.8Si, Al-16Si alloys and Al reinforced with A1203p (5%) indicated that increased silicon content reduced the wear rate. Hoskings et al. reported a decrease in adhesive wear rate with increasing particle content (at constant particle size) and dimension (at constant volume fraction) for Al 2014 and 2024 alloys reinforced with Al,O,p and SiCp (1-142 pm) of various weight fractions (2-30 %). SIC was shown to be more effective than alumina in resisting wear, when tested in a ball-on-disc rig. These findings were in disagreement with those reported by Surappa.54 However, it should be noted that the former work involved only a small reinforcement content (5 %). The effect of reinforcement particle size and volume fraction on the dry sliding wear resistance was also provided by Chung and Hwang.56 In this investigation the matrix consisted of commercially pure aluminium and the reinforcement phase was SiCp of the following size fractions: 2-5, 15-25 and 70-85 pm. The reinforcement content varied up to 30 wt%. For a constant SiCp size fraction, wear resistance increased with increasing SIC content. It was also stated that for a constant SiCp volume fraction, MMCs containing coarser particles exhibited higher wear resistance. Delamination wear was considered the predominant wear mechanism. The coarser SiCp were reported to provide greater resistance to the propagation of subsurface cracks as compared to the finer Sic fractions. These findings presented by Chung and Hwangs6 are particularly significant since the use of aluminium as the matrix alloy avoided the effects of ageing and the formation of interfacial compounds within the microstructure. The adhesive wear rate for an Al-1OZn alloy reinforced with corundum (5-40 %) obtained by Anand and Kishore, supported the findings of Hoskings et a/.5 The optimum particle content for adhesive wear resistance of this composite was considered to be within the range of 25-35%. Evidence for the delamination wear mechanism,13 was observed in composites that contained more than 15 % of the reinforcement phase. Pramiia Bai et a/ studied dry sliding wear of an A356 alloy reinforced with SiCp (43 pm, 15 and 25 %), using a pin-on-disc machine, for various applied pressures up to 26 MPa. With increasing applied pressure the wear behaviour of the unreinforced alloy was dominated by extensive plastic flow

422

R. L. Deuis, C. Subramanian,

J. M. Yellup

of the alloy surface and significant wear debris formation. The addition of Sic reduced the wear for the applied pressure range examined. SIC particles were reported to minimise this plastic deformation on the wearing surface and promoted the formation of an iron-rich layer on the composites surface. It should be noted that these researchers found no indication of a transition in the wear mechanism as a function of applied pressure. They did, however, comment that at low pressure, abrasive and delaminated wear processes were probably active. Venkataraman and Sundararajan21 also studied sliding wear rate and friction as a function of reinforcement volume fraction. Composite pins composed of an aluminium matrix reinforced with a range of SiCp content (O-40 vol.%) were slid against a hardened steel disc. Two loads were studied (52 and 122 N) and the sliding velocity was 1 m s-l. From this research it was concluded that wear rate decreased with increasing reinforcement volume fraction. Zhang et a1.s9 employed Vickers microhardness measurements and scratch testing to explore the influence of reinforcement particle size and content on friction and wear rate in an Al 6061 matrix containing either SiCp or A1203p (O-20 vol.%). This study utilised the following conditions: a scratch velocity of 6 mm s-, 6 mm wear track and a load of 10 N. MMCs reinforced with large particles revealed a lower wear rate. The effect of increasing reinforcement volume fraction resulted in an overall higher composite hardness and friction coefficient but a reduced wear rate. Composites containing small particles showed wear rates that were more sensitive to hardness, whereas the wear rates for materials containing large particles were more dependent on both hardness and particle fracture. The ratio of penetration depth to particle size (as defined by Wang and Rack27) was believed to be an important factor governing the tendency for the reinforcement particles to be removed by the indenter. Iwai et al. reported that increasing additions of SiCw (diameter 0.3-0.6 ,um, length 5-15 pm) improved the wear resistance of an Al 2024 alloy in both severe and mild wear regimes. Composites containing up to 16 vol.% of whiskers were studied via a pin-on-disc rig (counterface 0*45%C steel, 40 N load and sliding speed 0.1 m SK). The sliding distance corresponding to the transition from severe wear to mild wear was observed to decrease with increasing volume fraction. This was stated as being a direct result of surface hardening due to higher amounts of SiCw and oxidation. The significance of the oxidation in this transition mechanism remains doubtful, since detection of aluminium oxide on the wearing surface was achieved not by X-ray diffraction (XRD) but by X-ray photoelectron spectroscopy (XPS). The detected oxide was described as being an extremely thin layer attached to some fragments of SiCw.

Roy et a1.60 characterised the friction and wear behaviour of Al composites reinforced with the following particulates: Sic, Tic, TiB2 and B4C (10, 20 vol.%, size l-9-7.0 pm). Dry sliding wear tests were conducted using a pin-on-disc machine at two loads (80 and 160N). Results indicated that the friction coefficient of the composites was 30 % lower than that of the aluminium matrix, while the wear rates for the composites were lower. The type and size of the reinforcement phase were reported to have a negligible influence on the wear rate and friction coefficient. Volume fraction of the reinforcement had a marginal effect on the wear rate. Composites containing TIC exhibited a high wear rate due to the existence of much lower shear strain at the Tic/matrix interface. The introduction of the reinforcing particles appeared to increase the load corresponding to transition from mild to severe wear. With reference to dry sliding wear, Modi et a1.4 showed that reinforcement/matrix interfacial strength and the dispersoid shape both greatly influenced composite wear rate. Composites reinforced with Sic particles revealed a lower wear rate than those containing Sic fibres. The SIC fibres were associated with poor interfacial bonding that promoted dispersoid pullout, leading to void formation, a higher probability of crack nucleation at these voids and higher wear rate due to three-body abrasion effects. Alpas and Embury6 stated that in situations where dry sliding wear occurred, a delamination wear was active with sub-surface cracks mechanism nucleated at the SIC particle/matrix interface. Both composite and unreinforced alloy showed similar wear rates. These researchers implied that the second phase particle size played a significant role in determining the size of the wear debris.

2.6.2 Wear of counterface During dry sliding of aluminium-silicon alloys and aluminium-based composites, wear of the counterface is usually evident. Antoniou and Borlandh2 showed that under conditions of mild wear a tribolayer was formed on the surface of Al-Si alloys. This layer was reported to consist of an ultrafine mixture of aluminium, silicon and a-iron. The formation of this surface resulted from the transfer of material from both the Al-Si alloy pin material and the steel counterface. Dry sliding wear studies employing Al-Si based composites (A356 reinforced with 20 vol.% SiCp) have also shown that iron from the counterface is deposited onto the composite surface.20 The extent of this deposition was, however, found to be more significant for MMC/steel couples, when compared to Al-Si/steel sliding couples, due to the micromachining effects of the SIC particles on the counterface. Iron detached from the steel counterface due to the abrasive action of the SIC particles, was

Dry sliding wear of aluminium composites-a

review

423

believed to be oxidised during its transfer to the composite surface. Sliding wear studies of an Al 6061 alloy containing 20 % AlzOsp, performed at elevated temperatures (25-5OOC), have been published recently.h3 In this work Singh and Alpas also reported that iron from the counterface oxidised and was then deposited on the composite surface. It was proposed that this resulted in an improvement in wear resistance in the mild wear regime (ring-on-disc, sliding speed O-2 m s-, applied load lo-50 N). This reduction in wear rate due to the formation of the iron oxide phase on both sliding surfaces, was attributed to the fact that iron oxides can act as solid lubricants.64,65 However, the micromachining effect of the reinforcement phase, as previously stated, can also effectively remove these oxide particles via three-body abrasive processes during sliding wear. In a study of dry sliding wear of an Al 2124 (A1,4~65Cu,l~69Mg,O~llFe,0~92Mn) alloy reinforced with SiCw, against a counterface of stainless steel 17-4 PH (Fe,0~04C,16~5Cr,4*25Ni,O~25Nb,3~6Cu) Wang and Rack6 noted that an increase in reinforcement volume fraction resulted in a higher counterface wear rate. This observation was reported for both the initial run-in and steady-state (see Fig. 5) wear periods. The coefficient of friction for the composite-steel arrangement was found to be much lower compared with the unreinforced matrix-steel system. These researchers also stated that a reduction in surface roughness for both sliding surfaces produced lower wear rates in the initial run-in period for both surfaces but had little influence on the steady-state wear rate of either the counterface or the pin material. Zhang et a1.67 also reported that the counterface wear rate increased for increasing values of both composite reinforcement volume fraction and particle size.

In another study, Wang and Rack6* commented that cluster formations of the reinforcement phase within a composite acted as an asperity possessing a higher hardness than the counterface surface. This resulted in a higher wear rate for the counterface relative to the composite under dry sliding conditions. 2.7 Wear mechanisms of aluminium-silicon alloys In this section, studies relating to transition in wear rate as a function of variations in sliding wear parameters, such as applied load and sliding distance, are discussed. In a study of dry sliding wear of two Al-Si alloys (10.9 and 22.1%Si) against a steel counterface, two wear regimes were observed.43 The first wear regime was described as a mixed mode of elastic-plastic contact where Archards lawl was obeyed. The second regime, activated at a critical applied load, was characterised by gross plastic flow where the wear rate was not directly proportional to the load. It was also stated that the hyper-eutectic alloy experienced a higher wear rate than the hypo-eutectic alloy. This last observation has been supported by findings reported by Clarke and Sarkar.44 Work on the dry sliding of an Al-17Si alloy by Krishna Kanth et a1.69 confirmed the existence of a transition in wear rate related to applied pressure. The applied pressure range investigated was l-30 MPa. Regime I was characterised by a worn surface that contained deep abrasion grooves, craters and no evidence of a transferred iron-rich layer. In this wear regime, abrasion and delamination wear processes were considered predominant. At an applied pressure of 13.1 MPa the wear process was described by a Regime II mechanism. In this regime, wear rate increased and the wearing surface developed extensive plastic deformation and a smooth topography. Shivanath et al. classified the two wear mechanisms in the dry sliding of Al-Si alloys as oxidative and metallic wear:
l

174 PH &kc (clmnt~ca) ,cix / %,,---a

fL-&x+

Volume Percent SIC, Fig. 5. Dry sliding wear rate of an Al 2124/SiCw MMC

against stainless steel 17-4 PH as a function of SiCw v01.%.~~

Oxidative wear occurred at low applied loads. In this wear process an aluminium oxide layer (lo-80 pm) formed on both the wearing Al-Si surface and the counterface. Wear occurred firstly by oxidation of the asperities and then secondly by fracture and compaction of the oxidised wear debris into this film. The wear rate was low, due to the amount of metal removed being confined to the thickness of this oxide formation. Some localised deformation of the substrate and fracture of the Si particles was observed. Oxidative wear was considered to be generally independent of the Si content or Si particle size. Metallic wear became the predominant wear process at higher applied loads. The Al-Si wear surface was characterised by plastic deformation

424

R. L. Deuis, C. Subramanian,

J. M. Yellup
+mnll debris r-Large laminate debris

and fracture, significant transfer of material between the sliding surfaces and wear debris formation. The amount of plastic deformation and the higher wear rate prevented the formation of an oxide layer. A higher Si content was reported to increase the load at which this wear regime was active. It should be mentioned in this section that the oxidative wear process defined above was supported by Razavizadeh and Eyre.71 These authors proposed that during dry sliding wear of Al-S1 alloys, oxidation of asperities first occurred, and was then followed by asperity fracture and compaction into the wearing surface. A critical surface temperature was assumed that corresponded to the formation of an oxide film and prevented surface metal asperity contact and, therefore, minimised adhesive wear. Antoniou and Borlandc2 reported that oxidation of aluminium did not play a significant role in the dry sliding wear of Al-Si alloys. It was also stated that the observed iron-rich layer, formed on the Al-S1 surface due to mutual material transfer between the sliding surfaces, was composed of an ultrafine mixture of Al, Si and a-Fe (co.1 pm). Furthermore, it was argued that as wear continued, the deposition reached a stage where these transferred layers could not adhere and became detached. At low loads the wear debris was equiaxed, laminar and consisted of a large iron content. At high loads the higher shear stresses in the surface layers removed wear debris after a shorter cycle time with the wear particles being described as laminar with a granular under-surface. At these higher loads a delamination wear mechanism was inferred. For both an Al-12*3%Si alloy and Al-Sic MMC/AISI 4340 steel sliding systems, a-Fe20, was found to be present in the wear debris,72 indicating that oxidation of the counterface had occurred. The porous structure of a-Fe203 provided poor oxidation protection and contributed to a noticeable transfer oxidised under layer thickness. The counterface relatively low pressure and at low sliding speeds. It was further suggested that the presence of this oxide reduced wear rate by separating the contacting metallic asperities and thereby reducing the frictional forces. Oxidation of aluminium was not considered to be a major factor in the wear process. The existence of an extremely thin aluminium oxide layer on the wearing surface of aluminium-based composites has been detected (by XRD and TEM techniques) in other studies,73,74 but only reported on a qualitative basis. The relative proportion of this phase compared to matrix material and oxidised steel counterface debris (Fe203) within the tribolayer has not been cited. The aluminium oxide formed was, however, characterised as consisting of very fine particles.73 The significance of this oxide layer within the wear process, therefore, remains doubtful.
Crac

Debris

(b)
rMaterial transferred to the counterface

75pm 1

(9
Fig. 6.

Schematic diagram of the following wear processes in AI-Si alloys: (a) mild wear, (b) severe wear and (c) seizure, as reported by Reddy et a1.45

In research reported by Reddy et aZ.45three distinct wear regimes were identified as a function of applied load. These wear regimes are discussed below and presented in Fig. 6:
l

Mild wear. At low loads, the worn surface was characterised by the formation of an iron-rich compacted debris layer, supporting the observations of Antoniou and Borland. The presence of an oxidation film in this wear regime, as proposed by Razabizadeh and Eyre,71 was not detected. Wear debris formed in this regime was a result of abrasion and cracking of this protective layer with a significant proportion of this debris again added to the compacted layer. Severe wear. As load increased a delamination type wear mechanism was operative in which sub-surface deformation and cracking caused fragmentation of the Si particles and removal of the iron-rich protective layer. Wear debris dimensions were related to the extent of the tribolayer formed and were in accordance with delamination wear theory,13 resulting in a high wear rate. Seizure wear. In this wear regime, near surface temperatures were high enough to lower the shear strength in the sub-surface layer and promoted extensive material transfer from the wearing alloy to the steel counterface. Shivanath

Dry sliding wear of aluminium composites-a et aZ.Oindicated .

review

42.5

that seizure of these alloys was enhanced under load conditions where the iron-rich compacted layer was removed. Additions of graphite (3-10 vol.%) to composite materials (A356 reinforced with 20 vol.% SiCp) also improved seizure resistance and reduced counterface wear by minimising the frictioninduced surface heating associated with dry sliding wear.
l

2.8 Wear mechanisms composites

of aluminium-based

2.8.1 Effect of applied load

Alpas and Zhang 20,76investigated the effect of Sic particle reinforcement on the dry sliding wear of an A356 alloy under different applied loads (0*9-150N). In this study the wear behaviour of an A356 alloy reinforced with SK 2Ovol.% (5-10p.m) was compared to that of the unreinforced alloy. Three distinct wear regimes are identified in Fig. 7 and are summarised below. . Regime I (<lON): in this wear regime, the composite exhibited a lower wear rate than the unreinforced alloy. The wearing composite surface also possessed an iron-rich transferred layer (5-15 pm) that indicated a material transfer process was active. Wear proceeded by spalling of the iron-rich layers. The Sic particles acted as a load bearing phase. The addition of the SIC phase impeded the transition to Regime II. Regime II (lo-95 N): wear rates for both materials were similar in this regime with surface morphologies characterised by severe plastic deformation. The applied load produced stresses higher than the fracture strength of the SIC phase, therefore the matrix was in direct contact with the steel counterface and large

plastic strains resulted in the formation of a tribolayer. Delamination wear was the predominant wear mechanism. In the case of the sub-surface cracks occurred via composite, decohesion at the Sic/matrix interface and for the A356 alloy, sub-surface crack nucleation was associated with the Si particles. Wear rate was generally independent of the size and volume fraction of the SIC particles. Regime III (~98 N): at these high applied loads, the wear rate of the A356 alloy was about two orders of magnitude greater than in Regime II. Severe wear was associated with adhesion of aluminium to the steel counterface and the occurrence of thick irregular plate-like wear debris particles (200-500 pm). Plastically deformed zones penetrated to a depth of 300-400pm. The SIC phase appeared to be responsible for improving the wear resistance of this alloy at high loads by improvement in the thermal stability of the matrix. It was also argued that composites had the ability to retain room temperature strength during the frictional heating up to a critical temperature. Once this temperature value had been exceeded material near the contact surface softened and the worn surface layers became detached and transferred to the counterface. This concept of thermal stability discussed by Alpas and Zhang76 clearly requires more research in order to elucidate the role of SIC within this wear regime.

The fracture strength of the SIC particles directly influenced the magnitude of the transition load between Regimes I and II.2o With the assumption that the applied load, F, was supported by the carbide particles on the composite surface and particle fracture strength was dependent on the presence of a critical dimension defect inherent in the particles, the following relationship was given:
f 5 = ~z(G,~E)~ p (2r)12

1
Fig. 7.

10 Load

10 N

Log wear rate versus log load diagram for an A356 alloy (0) and A356-20% SIC MMC (H).

where A = nominal contact area, H = constant that relates the proportionality between critical defect size and particle size, GIc = fracture toughness of the SiCp, E = Youngs modulus of the SiCp, fP = particle volume fraction (areal) and r = average radius of SiCp. Moustafa77 reported that the addition of alumina fibres (26 vol.%) to an Al-22%Si alloy increased the transitional load (between mild and severe wear) by more than a factor of three and reduced the coefficient of friction. The near surface temperatures associated with this transition were measured as 160 f 10C and 110 * 10C for the MMC and the matrix alloy, respectively.

426

R. L. Deuis, C. Subramanian,

J. M. Yellup

Alpas and Zhang commented that increasing the SIC volume fraction would increase the transition load; a transition load of approximately 0.9 N for A356-10% SiC and 17 N for A356-20% Sic. Another study of dry sliding wear of A356 reinforced with varying volume fractions of SiCp (10, 15 and 20%), confirmed the relationship between wear transition and applied load.78 This study was limited to an applied load range of 15-50 N with these researchers indicating a Regime I to II transition as described by Alpas and Zhang. Zhang and Alpas proposed a similar wear regime and transitions for an Al 6061 alloy reinforced with 20 vol.% A&O,p (15 pm) as identified previously.76 These experiments were carried out using a block-on-ring rig, at a sliding velocity of 0.2 m SK and an applied load range of l-350N. In comparing the two studies, Zhang and Alpas showed that the transitional load from Regime I to II was higher for composites that contained alumina particles It was concluded that this was due either to a higher fracture toughness inherent in these particles or a higher matrix-reinforcement interfacial bond strength. The transition from one wear mechanism to another as a function of applied pressure has also been documented by Zhang et a1.67 This study involved the dry sliding of Al 6061 containing SiCp or AlzOlip (10, 20 vol.%) against a mild steel disc, at a sliding speed and for a defined pressure range of lms- (0.3-3 MPa). Three wear mechanisms were identified for the MMCs/steel sliding system studied. Initially, sliding was described as an abrasive wear (running-in period). During this time period reinforcement particles exposed on the surface caused material removal from the steel counterface in the form of continuous chips. The associated coefficient of friction was high. For low pressures, and with continuing sliding distance the wear process was transformed into an oxidative steady state wear mechanism. Oxidative wear was attributed to the rise in sliding surface temperature due to the increase in sliding distance. It was noted that in this wear regime, the oxidised wear debris particles promoted a reduction in friction coefficient and wear rate. With an increase in pressure, a severe wear regime occurred which corresponded to a critical pressure value. Within this wear regime delamination wear was predominant. Zhang et a1.67 stated that increases in reinforcement volume fraction and particle size or an effective particle/matrix interfacial bond all contributed to increasing this transitional pressure value. In summary, the transition from abrasive to oxidative wear was considered to be a function of sliding distance, whereas the transition from oxidative to adhesive wear was dependent on the applied pressure. The relationship between wear rate and applied normal pressure (expressed as normal stress) for Al 6061 reinforced with SiCp and Al,O,p is shown in Fig. 8,

31 0.2
Fig. 8.

0.3

0.5
Applied

1 normal

2
stress, MPa

Relationship between wear rate and applied normal stress for SIC and A&O, MMCs: (U) lOvol.% Sic, (m) 20 vol.%, (0) 10 vol.% A&OX,(0) 20 vol.% A1203.h7

where symbols B and C define the existence of oxidative and severe wear regimes, respectively. Venkataraman and Sundararajan reported that the addition of a small content of SiCp to aluminium (10%) prevented the transition from mild to severe wear during dry sliding. Wear studies involving the dry sliding of an MMC pin of Al 6061 alloy reinforced with either SiCp or A1203p (10 and 20 vol.%) against a steel counterface were cited by Zhang et a1.80 A wear model was presented to describe the volumetric wear, V, present in the adhesive wear regime:
c, LSV,

v = (1 - v&f

where c1 = experimental constant, A = mean free path, L = normal pressure, S = sliding distance, V, = reinforcement particle volume fraction and d = reinforcement particle diameter. In reference to this wear model, wear rate was considered to be proportional to the applied pressure and sliding distance and also increased with particle content and decreased with reinforcement dimension. The transition pressure (indicating the change from mild wear regime to adhesive wear regime) increased with increasing particle volume fraction and also increased with increasing reinforcement particle size, for a given volume fraction. Finally, these researchers concluded that particle size influenced the transition pressure more significantly than the effect of particle volume fraction. 2.8.2 Effect of sliding speed Wang and Rack81 studied dry sliding wear behaviour of an Al 7091 (A1,5.6Zn,2Mg,l*lCu) alloy reinforced with SiCp and SiCw and compared the results to the unreinforced alloy. At a sliding velocity of ~1.2 m s- they described the wear mechanism as fatigue-related surface cracking. Wear debris was small and generally metallic and the wearing surface was covered with a

Dry sliding wear of aluminium composites-a

review

427

tribolayer. At these low sliding velocities, the reinforcement did not appear to influence the wear rate and the wear rate for the composites and matrix alloy was similar. At higher sliding speeds, a transition in the wear process occurred which was associated with a breakdown of the tribolayer, with wear being controlled by sub-surface cracking-assisted adhesive transfer and by abrasion. In steady state sliding the choice of reinforcement type had no influence on the wear rates of the composite. Similar transitions in wear behaviour with increase in applied load were reported by Lim et al. for an A356 alloy reinforced with SiCp. At a slow sliding velocity (0.5 m s-l), the wearing surface was covered with a transferred material layer. At higher speeds an oxide-like transferred layer formed at the sliding interface and reduced direct metallic contacts. This resulted in a lower wear rate. At very high speeds thermal softening of the matrix was reported. The oxide layer was observed to break down and allow greater direct metallic contact during sliding and Sic particles became dislodged and three-body abrasive wear was predominant. Further evidence of the wear transition with sliding speed in the range of 0*08-1.98 ms- was presented by Park.* This study involved the sliding wear of an Al 6061 alloy reinforced with SiCw and SiCf. At low sliding speeds (0.08 m s-) the wear process proceeded with the removal of material by fracture of the reinforcement and the matrix due to high frictional forces. At an intermediate speed (0.94 m s-l), the wear rate was reduced. The predominant wear process was the formation of micro-grooves. These grooves were observed to form in fractured regions of the reinforcement and matrix. It was suggested that adhesive and abrasive wear were the predominant mechanisms at low and intermediate sliding speeds. At high sliding speeds (1.98 m s-l) Park reported localised melting on the wearing surface. Kwok et al.n3 observed the adhesive wear behaviour of composite Al-4~5Cu-15 vol.% SiCp as a function of sliding speed and applied load. The wear mechanism corresponding to a load <50N and a sliding speed of =5 m s- was a combination of abrasive, delamination and melt wear. For sliding conditions associated with a 50 N load and a speed of 5 m s-l severe adhesive wear was identified as the main mechanism. For these sliding conditions significant regions of composite material were transferred onto the coupled steel disc. Melt wear was reported to be the predominant wear mechanism for sliding speeds >10 m s-. This increased sliding speed resulted in the formation of molten matrix material that was transferred onto the steel disc in the form of small lumps. The resultant worn MMC surface, in this case was characterised by a dense layer of compacted Sic particles. Subramanians4 investigated the dry sliding wear

behaviour of an Al-12.3%Si alloy as a function of sliding speed using a pin-on-disc rig. He reported that the wear rate of the alloy decreased with increasing sliding speed up to a critical speed, beyond which it increased. This critical speed was dependent on the applied load, thermal diffusivity and hardness of the wear surfaces. Temperature due to frictional heat showed an increase with increasing sliding speed. The rate of increase above the critical speed was greater than below this value. A transition in the wear mechanism was associated with a change in wear debris morphology as the sliding speed was increased.
l

Low speeds: equiaxed wear particles were formed. Critical speed: delamination of the compacted particles occurred. Above critical speed: plastically deformed material delaminated as flake debris.

Subramanian concluded that below the critical speed, the increase in strain rate with increasing speed led to an increase in hardness and a lower wear rate. with higher speeds the increase in However, temperature due to frictional heat softened the material, increasing the true area of contact which resulted in a higher wear rate. Below the critical speed the influence of strain rate was more pronounced than the friction-induced temperature effect. Above this speed the temperature effect was believed to dominate. The influence of sliding speed on wear rate is demonstrated in Fig. 9. 2.8.3 Effect o,f temperature In one of the very few studies published regarding the effects of temperature on the tensile properties of an
16 1

, , , , ,

I,,

IL

Pressure

0.5 MPa

Sliding Speed rns-


Fig. 9.

Wear behaviour of an Al-12,3%Si


different counterface

alloy slid against materialsM

428

R. L. Deuis, C. Subramanian,

J. M. Yellup

aluminium composite (Al 2014 reinforced with Al,O,p), Zhao et a1.85 reported that the failure mechanism was dependent on temperature. Two failure modes were identified. Particle fracture predominated at low temperatures, while interparticle void formation was considered the main failure mode at elevated temperatures. A transition temperature was said to indicate the change in failure mode from particle cracking to interparticle voiding and the value of this parameter was stated to increase with reinforcement volume fraction and size. It is possible that a similar rationale may be applicable to the failure mechanism in dry sliding wear of composites at elevated temperatures. Singh and Alpas studied the dry sliding wear of an Al 6061 alloy and composite Al 6061-20 vol.% Alz03 at a range of temperatures (25-500C). Results of their work are shown in Fig. 10. Both materials exhibited a mild to severe wear transition when the environment temperature was increased above a certain value. This transition occurred at a higher temperature in the composite. At high temperatures severe wear was reported in the form of gross plastic deformation and material transfer to the counterface, indicating that softening of the matrix material significantly influenced the wear rate at these temperatures. In the mild wear regime, oxidation of the iron counterface and its ability to act as a solid lubricant64 resulted in lower wear rates, as previously stated. These researchers attempted to simulate the deformation behaviour present in this alloy and MMC, at elevated sliding temperatures, using uniaxial tests. The reported flow stress and compression work-hardening rate trends (displayed as a function of temperature), however, can only be viewed as approximations.

2.9 Wear mechanism maps Wear mechanism maps, as developed by Lim and Ashbys6 for a variety of steels, help to classify wear information and show the relationship between the different wear processes in dry sliding situations. Antoniou and Subramanian, following the work of Lim and Ashby, developed a wear mechanism map for aluminium alloys, with each mechanism identified by direct observation of the wearing surfaces of both the aluminium alloy and the steel counterface, and the associated wear debris morphology, and constructed the map displayed in Fig. 2. These wear tests were performed using a pin-on-disc rig. Antoniou and Subramanian characterised the following regions: Region a. At low sliding speeds and applied pressures, the worn aluminium alloy exhibited a smooth surface, containing granular wear scars and dark regions associated with a high iron content. These dark areas confirmed that material transfer had occurred between the steel counterface and the aluminium. Fine equiaxed wear debris (~5 pm) consisting of Al, Si and a-Fe6 was also evident. Region b. In this wear regime, the wear debris showed a flake-like morphology with a smooth upper surface and granular under surface. It was proposed that this wear debris formed by compaction of the fine equiaxed particles into depressions on the wearing surface and then a delamination mechanism produced the flakes. Composition of this wear debris and the wearing surface topography of the aluminium were similar to that observed in Region a. Region c. With increased load and sliding velocity, the delamination wear mechanism was the dominant wear process and produced flake-like wear debris that consisted of Al and Si with smooth upper surface and under surface, indicative of ductile shear. Material transfer was a more notable process. The wearing aluminium surface revealed smooth and scarred areas that were caused by ductile shearing. Region d. At very high loads and sliding velocities, there was a significant transfer of material between the sliding surfaces. Seizure within this wear regime was probable. Region e. At very high applied loads these researchers postulated that melt wear was the predominant wear process. From work reported by Bowden and Persson this wear process was characterised by a low coefficient of friction and removal of the material by the formation of an interfacial molten film between the sliding surfaces. Howell and Ba11,48characterised melt wear in composites as a layer of molten MMC trapped between a surface of unmelted MMC

1Ci3

200

400

Temperature

Fig. 10. Wear rate as a function of temperature for Al 6061-20% A&O3 MMC tested at 10N (=), Al 6061-20% Al,O, MMC tested at 50N (El), and alloy Al 6061 tested at 10N (0).63

Dry sliding wear of aluminium composites-a

review

429

and the counterface material. Under these sliding conditions wear rate was minimised and a stable coefficient of friction was obtained. With reference to Fig. 2, the mechanism boundaries for Regions a, b and c are justified by the findings of previous experimental data. However, the positioning of Regions d and e as functions of normalised pressure and velocity can only be regarded as approximate. However, the existence of these two wear mechanisms at some definitive combination of load and sliding velocity is not disputed. Another wear mechanism map for aluminium alloys was proposed by Liu et a1.89In this case the dominant wear regimes were constructed with reference to the wear mechanism models derived by Lim and AshbyXh and wear constants reported from empirical studies. Four predominant wear mechanisms were reported and are displayed in Fig. 11:
l

a $j 60-I4 1 z

30-

1.54

Sliding

2.75 Speed

3.75 m/s

5.00

Fig. l2.Dry sliding wear mechanism map for an as-cast Al 6061 alloy reinforced with SiCw.

Oxidation dominated wear. This wear mechanism was reported at relatively low sliding velocities and loads. Delamination wear and severe plastic deformation wear. These two wear mechanisms were based on the delamination wear theory postulated by Sub. Melt wear. This wear regime was stated to be predominant at high loads and sliding velocities, where a high wear rate was characterised by gross material transfer, due to severe plastic deformation or ejection of the melt.

A third wear mechanism map has been presented by Wang et a1.73,90with the regimes defined as a function of normal load and sliding speed. This map corresponds only to data relevant to the dry sliding wear of an Al 6061 alloy reinforced with SiCw (2Ovol.%) and is displayed in Fig. 12. From this map, it was concluded that increases in sliding speed and/or normal load resulted in a transition in wear behaviour from mild wear to severe wear and ultimately to seizure. At low loads, where a mild wear regime was reported, the wearing surface was described as relatively smooth and wear debris as small and brittle. Oxidation of aluminium was considered to be an important part of this wear mechanism even though no quantitative data regarding oxidation was given. In the severe wear regime a delamination wear process was stated. Formation of wear debris and associated transfer of material onto the steel counterface was related to the poor whisker/matrix interfacial bond strength. Comparing the first two maps presented here (Figs 2 and ll), Liu et a1.89supported their map with more experimental data, so the boundaries between the different wear regimes were better defined than those published by Antoniou and Subramanian.* However, the use of the mechanism models derived by Lim and Ashby= are at best approximate in defining these regions due to use of the empirical model constants employed by Liu et aZ.x9 At present, these maps can only provide a guide to the nature of dry sliding wear in aluminium alloys under the stated operating conditions. Application of such maps in the study of composite wear behaviour would have to incorporate the influence of intrinsic tribological factors, such as reinforcement volume fraction, morphology and size. 3 SUMMARY
3.1 Adhesive wear

Nomalised

Velocity

Fig. 11.Wear mechanism map for aluminium alloys.8

The worn surface produced by adhesive wear, is normally characterised by transfer of material from

R. L. Deuis, C. Subramanian,

J, M. Yellup

one surface to the other, resulting in a tribolayer. This layer generally consists of plastically deformed and oxidised particles that reduce the wear rate of the material. Studies of the tribolayer, reveal that it is composed of a mechanical mixture derived from the wearing alloy and the counterface. The extent of its formation depends on the sliding wear parameters such as applied pressure and sliding speed. Conditions of low speed and high pressure promote a large number of transfer and back-transfer cycles, resulting in a noticeable layer thickness. Hardness of the wearing surface also influences the adhesive wear rate, by directly determining the real area of contact between asperities of both contacting surfaces. Both the classical adhesive wear and delamination wear theories neglect a tribolayer formation in their discussion of dry sliding wear. The importance of the materials microstructure with respect to wear behaviour is disregarded by the classical theory, so this hypothesis can only apply to idealised sliding conditions. For conditions of low sliding speeds, delamination wear theory was formulated and helps describe the formation and dimensions of wear debris. This mechanism incorporates the influence of the microstructure, hardness and fracture toughness. The theory emphasises that wear debris is formed at a defined depth below the wearing surface by delamination cracks formed due to the growth and coalescence of voids. This critical depth results from a balance between the hydrostatic pressure that opposes void formation and plastic deformation effects that promote void nucleation. Increases in sliding distance or the magnitude of the applied load favours a greater critical depth. Findings involving the relationship between wear rate and volume fraction of second phase particles within the wearing material, remain indefinite at this stage. Wear rate decreases for composites as the volume fraction of the reinforcement phase is increased above a certain value ( = 15%). Evidence of the existence of the delamination wear mechanism, has been reported for MMCs containing these higher fractions. Composites fabricated with lower particulate or fibre contents, display higher wear resistance than their respective matrix alloy. Reinforcement morphology directly influences wear rate. Fibres, as compared to particles, promote greater crack nucleation and offer a lower reinforcement/matrix interfacial strength. The second phase particles play a significant role in determining the size of the wear debris. For steady state wear conditions, the coefficient of friction decreases with increasing reinforcement content. Increased reinforcement content reduces the area fraction of the matrix in contact with the counterface, thus minimising the smearing effect of aluminium on the counterface surface and results in smaller temperature increases at the sliding interface. Friction

coefficients for composites are reported to be 30% lower compared to the aluminium matrix. Wear of the counterface is noted, with reference to the dry sliding of Al-Si alloy/steel and MMC/steel systems. Findings indicate that a proportion of the counterface material removed by the micro-machining effects of the reinforcement phase, is subject to oxidation (even at room temperature). This oxidised material is present as Fe,O, in the tribolayer of both wearing surfaces. Furthermore, this oxidised phase helps stabilise the tribolayer on the MMC surface by pinning dislocations. Sliding wear studies (at elevated temperatures) report that the oxidised counterface material transferred onto the MMC surface, temporarily reduces the wear rate (for mild wear conditions) by acting as a solid lubricant. However, the micro-machining action of the second phase particles within the MMC, then enhanced the removal rate of these iron oxide particles. This finding, supported by other studies, indicates that increased reinforcement fraction results in a higher counterface wear rate. The additions of silicon (up to the eutectic composition) increase the wear resistance of Al-Si alloys by improving the seizure resistance for these alloys. The influence of reinforcement content on coefficient of friction could be used here to explain the associated with silicon higher seizure resistance content. A definite term for seizure was not cited in the compilation of this review paper. The only information regarding this wear phenomena, was that prior to the onset of seizure, gross material transfer from the wearing alloy to the steel counterface material may be observed. In summary, a composite material that possesses superior adhesive wear resistance is associated with a stable tribolayer. The creation of this layer depends on the magnitude of the applied load, sliding speed, operating temperature and composition of the MMC. In reference to the microstructure of the composite, reinforcement volume fraction appears to be a critical parameter. Achievement of a lower wear rate for the counterface, would involve the utilisation of materials exhibiting high hardness and adequate fracture toughness. The micro-machining action of the MMCs reinforcement phase, towards the counterface, could be minimised by selecting reinforcement particles of low fracture toughness. However, the wear resistance of the composite is then reduced. Clearly, a compromise is required between wear rates expected for both the composite and the counterface materials. 3.2 Wear mechanisms Concerning the dry sliding wear of aluminium-silicon alloys, the wear mechanism varies as a function of applied load, with low wear rates being recorded at low loads. Conflicting theories, however, have been published regarding the wear mechanism within this

Dry sliding wear of aluminium composites-a

review

431

low load range and this is perhaps due to the fact that varying speeds and different counterface materials were used. Early studies reported that the oxidation of aluminium plays a significant role in formation of the wear debris and hence the tribolayer. The term oxidative wear was introduced to describe this wear process. More recent findings, confirmed by X-ray diffraction analysis, indicate that this oxidation process is not extensive. Wear debris within this regime has been described as equiaxed, laminar and of high iron content. The tribolayer is usually observed to be composed predominantly of an ultrafine mixture of Al, Fe and c-u-Fe (CO.1 pm). Within the low load range, the stability of this protective layer is high. At higher applied loads, high wear rates are observed. The wearing surface is characterised by a significant transfer of material between the sliding surfaces. A delamination wear mechanism has been inferred for this wear regime, where the tribolayer is removed by sub-surface plastic deformation and fragmentation of the silicon particles. At very high loads, the near-surface temperature generated is high enough to reduce the shear strength in the sub-surface layer promoting extensive material transfer from the AI-Si alloy to the steel counterface. Distinct wear regimes have been proposed for composite materials as a function of applied load. At low loads, composites exhibit low wear rates and the existence of an iron-rich tribolayer is noted (for MMC/steel sliding couples). Within this pressure range, the reinforcement particles act as load bearing elements. Fracture toughness of this phase is a critical parameter. Where the magnitude of the load creates stresses higher than the fracture strength of the reinforcement phase, the next wear regime commences. In this regime, matrix material and the counterface are in direct contact and delamination wear is the predominant wear mechanism. At higher pressures, the wear rate is considerable and associated with noticeable adhesion of the aluminium onto the steel counterface. Delamination wear is still the rate-determining mechanism. Transitions in the wear behaviour of MMCs with variations in sliding speed are documented. The wearing surface at low speeds exhibits a low wear rate and a stable tribolayer. In a defined sliding speed range, the extent of the tribolayer formation becomes significant and a reduction in direct sliding surface contact results in a lower wear rate. At very high speeds, thermal softening of the matrix indicates a breakdown in the tribolayer, corresponding to a high wear rate. A critical sliding speed has been observed for this transition. Wear rate of an MMC decreased with increasing speed up to this critical speed, and then increased. Below the critical speed the influence of strain rate is the determining factor, while above this value, friction-induced temperature effects dominate the wear process. For speeds both above and

below this critical value, a change in the wear debris morphology is evident. With increase in temperature, the wear rate of composites changes from mild to severe. At high temperatures, severe wear is associated with gross plastic deformation and material transfer to the counterface. A transition temperature is implied, but more research is required in this area. Wear mechanism maps, similar to those developed for wear encountered in the dry sliding of steel, have been reported for Al-S1 alloys. One research effort concentrated on the construction of a map by direct observation of the wearing surfaces of both the alloy and the counterface and the corresponding wear debris structure. In another study, a map was developed by modifying existing wear mechanism models for steel and incorporating wear constants for Al-Si alloys obtained from empirical data. As already mentioned within Section 2.9, these maps can only offer a guide to the actual wear process operating at defined sliding speeds and applied pressures. Dry sliding parameters (speed, load and temperature) that create the conditions for a stable tribolayer, and the formation of fine equiaxed wear, both promote a low wear rate. Sliding conditions characterised by low speeds and loads possess these attributes. Collection and analysis of wear debris and the examination of the wearing surfaces, are important investigative tools in attempting to identify the nature of the wear mechanism. 4 CONCLUDING
REMARKS

The purpose of this paper was to highlight the current research focus involving dry sliding wear of aluminium composites containing discontinuous reinforcement particles. A significant body of experimental data has been reported in the area of dry sliding wear behaviour, as indicated in Table 1. The two most common wear testing rigs used are pin-on-disc and block-on-ring. However, comparison between the tests is difficult due to the wide range of wear parameters and counterface materials used. Influences of reinforcement size, volume fraction and morphology on wear behaviour were investigated by many researchers, e.g. Bhansali and Mehrabian,96 Ma et al, ,97 Sato and Mehrabian,* Hoskings et al.,55 Chung and Hwang,56 and Rana and Stefanescu.25 However, none of these investigations were based on a systematic approach. Little information regarding particulate size and volume fraction related to abrasive or dry sliding wear processes is actually available. The data collated by these workers were restricted to only two reinforcement types, namely silicon carbide and alumina. Future research should incorporate the wear behaviour of other potential reinforcement materials such as Tic, TiB, and Si02. An optimum reinforcement content has been

Table 1. Summary
Counterface Test configuration HR, 63 Block-on-ring 0.9-150 0.5 pm 32 Pin-on-disc 2.9 Pin-on-disc Pin-on-disc 3.4 l-3 MPa 0.11-0.21 2.68 0.05-0.46 25 49 52 0.16-0.8 20 Sliding velocity (m s-l) Load (N) Ref.

of dry sliding wear parameters

for bulk MMCs and MMC coatings

Matrix

Reinforcement

Material 4 10-20 vol. % AISI 52100 both surfaces Tool steel EN 25 steel AISI 52100 50, 100 --23 =12 diam. 5-100 13, 15 10 vol.% 10 vol.% l-30

m
Material, roughness

Wt%

A356

SiCp

Al-1.5Mg LMll

SiCp SiCp SiCf

k ?

Al,Osp, SiCp, MgO, glass beads, S&N, Tic, silica sand, B,N l-142 16 Ball-on-disc AISI 52100

Al@,p SiCp

0.5-3.9

0.1 55 57 58 60 ; ,G 9

Alzo,~ SiCp SiCp, TIC, TiB,, SiCw SiCp 43 2-7 0.5 diam. 15

Al-4Cu, Al,4Cu,0.75Mg 6061 2014, 2024, as-cast 201b Al-1OZn A356 Al 2124 A356 6061 2-30 5, 20 5-40 15, 25 10-20 vol.% O-20 vol. % 10-20 vol.% 20 vol. % 63 50 45 35 63 63 Pin-on-disc Hv~~.shgr =345 52 HR, = 82 Universal wear (Block-on-ring) Pin-on-disc 19 58 tester tester tester Pin-on-disc Pin-on-disc Pin-on-disc Pin-on-disc Pin-on-disc Block-on-ring Cr steel EN 24 steel (CLA 0.3) Low alloy steel Stainless steel (CLA 0.25) AISI 01 steel (CLA 0.05) AISI 52100, both surfaces 0.5 Frn Stainless steel

B,C

Al,O,p

39 2-26 MPa 80, 160 14.2 15-50 l-350 0.43 MPa 31

5 0.5 1 3.6 0.5-2.5 0.2 0.36-3.6 0.08-1.98

66 78 79 81 82

$ 2 3 :. ,g 4 is 19.6 207.8 9.8 1 l-4 0.6-4.36 91 92 d 2 F +u

7091 JIS SCM 4 AISI 304 stainless steet Both surfaces <1.8pm Tool steel JIS S45CF steel SUJ2 steel

6061

SiCp SiCw SiCw 20 vol.% 20 vol. % 12-20 vol. %

5 0.5 diam. 0.45 diam. 3.5 2.50 vol. % 10 vol.% 61-147 88-246 43-150

Ti,6A1,4V

5083

Al AC2B

Al@,f SiCw + Al,O,f Ti,6A1,4V + W,C Ti,6A1,4V + Cr,C, NbC, TiC Sic Cu, Ni, Cr

93

5083 5083

NbC + (Cu, Ni) Si

50-150 60-150 30-70 vol.% 567

JIS S45C steel SUJ2, both surfaces polished are

58

Universal wear Block-on-disc Universal wear block-on-disc Block-on-disc Universal wear block-on-disc

tester

~207.8 10

l-4 4.36

94 95

Compositions for aluminium alloys and counterface materials not already mentioned in the text Fe,O.9C,l.OMn,O,SW,O.5Cr; 0~5Mn.0~15-0~55Mg,0~15-0~35Ti; Fe,O.39C,O.3Si,0.6Mn,0.8Cr,0.4Mo,l.4Ni; l,OSi,18-20Cr,8-10.5Ni; CFe,0.45C,0.3Si,0.7Mn; A1,5~0-7~0Si,l~OFe,2~0-4~0Cu,0~5Mn,O~5Mg,l~OZn,0~2Ti,O~3Ni;

listed here: Fe,0.3C,0.7Cr,2.5Ni,O.5Mo: Fe,0.41C,0.3Si,O.7Mn,l.lCr,O.2Mo; Fe,0~95-1~1C,0~4Si,0~5Mn,0~9-1~2Cr.

hA1,0~1Si,0~15Fe,4~0-5~2Cu,0~2fFe,0~08Cmax,2~0Mn,0~045P,0~03S,

Dry sliding wear of aluminium composites-a

review

433

coatings. For example, the PTA surfacing method allows the deposition of a macro-coating (3-5 mm thick). These MMC coatings exhibit microstructural and tribological properties characteristic of a bulk composite material but independent of substrate effects. Therefore, the information reviewed in this paper has direct relevance to the wear behaviour of these composite coatings.

ACKNOWLEDGEMENTS R.L.D. wishes to express his thanks to the of South Australia and the Commonwealth Industrial Research Organisation (CSIRO) cial assistance and to Mr J. R. Manuel for this manuscript. University Scientific for finanreviewing

Volume Fraction
Fig. 13. Ductility of copper-based composites as a function of second phase volume content.

loosely defined where the adhesive wear rate corresponds to a minimum. However, this value is dependent upon dry sliding conditions and composite microstructure considerations and, therefore, is not a universal constant. What can be confirmed is that increases in reinforcement volume fraction have been associated with reduced ductility and yield strength of composites,98 as indicated in Fig. 13. Therefore, composites that contain a high content of the reinforcement phase would tend to behave differently under wear conditions than those with a lower amount of reinforcement. The partnership between the wear mechanism map and wear debris analysis allows the determination of optimum sliding conditions that favour a stable tribolayer and, hence, low wear rates. Information of this nature is clearly beneficial to the design engineer. Other issues that need to be mentioned concern the following areas. The significance of aluminium oxide formation regarding the formation of wear debris and the overall wear process still remains controversial. With ever-increasing demand for MMCs suitable for high temperature sliding applications, knowledge in this area still remains in its infancy. The influence of thermal softening of the wearing composite material and its material transfer to the counterface are both significant at elevated temperatures (~~500C) and, therefore, warrant further study. Finally, more data concerning the wear of the counterface material would also enhance the selection criteria for MMC/counterface sliding couples for a given sliding situation. Generally in the reporting of sliding wear rates, emphasis is placed on the wear behaviour of the composite with little regard to the state of the counterface material. This review has concentrated on the study of bulk composite materials. Techniques such as plasma spraying, laser cladding and plasma transferred arc (PTA) surfacing are used to fabricate composite

REFERENCES
1. Terry, B. and Jones, G., In Metal Matrix CompositesCurrent Developments and Future Trends in Industrial Research and Applications. Elsevier Advanced Techno-

logy, 1990. for 2. Charles, D., Metal matrix composites-ready take-off? Met. Mater., 1990, 6, 78-82. 3. Zedalis, M. S., Bryant, J. D., Gilman, P. S. and Das, S. K., High-temperature discontinuously reinforced aluminum. J. Met., 1991, 43, 29-31. 4. Sanders, R. N. and Valdo, A. R., Eyhyl Corp., Aluminum-Silicon Composite, US Patent no. 3961945. 5. Vaccari, J. A., Cast aluminum MMCs have arrived.
Amer. Machinist, 1991, 135,42-46.

of the metal matrix 6. Feest, E. A., Exploitation composites concept. Met. Mater., 1988, 4, 273-278. composites for 7. Rohatgi, P., Cast aluminum-matrix automotive applications. J. Met., 1991, 43, 10-15. 8. Glossary of Terms and Definitions in the Field of Friction, Wear and Lubrication (Tribology ), Research Group on Wear of Engineering Materials, OECD, Paris, 1969. 9. Rabinowicz. E., Friction and Wear of Materials. Wiley, New York, 1965. C., Some considerations towards the 10. Subramanian, design of a wear resistant aluminium alloy. Wear, 1992, 155, 193-205. 11. Archard, J. F., Contact and rubbing of flat surfaces. J.
Appl. Phys., 1953,24,981-988.

12. Archard, J. F. and Hirst, W., The wear of metals under unlubricated conditions. Proc. R. Sot. Lond., 1956,
A236,397-410.

13. Suh, N. P., The delamination


1973,25, 111-124.

theory of wear. Wear,

14. Jahanmir, S. and Suh, N. P., Mechanics of subsurface void nucleation in delamination wear. Wear, 1977, 44,
17-38.

15. Rosenfield, A. R., A shear instability model of sliding wear. Wear, 1987,116,319-328. 16. Saka, N., Pamies-Teixeira, J. J. and Suh, N. P., Wear of two-phase metals. Wear, 1977, 44, 77-86. 17. Argon, A. S., Formation of cavities from nondeformable second-phase particles in low temperature ductile fracture. Trans. ASME, Ser. H, J. Engng Mater.
Technol., 1976,98,60-68.

434

R. L. Deuis, C. Subramanian,

J. M. Yellup

18. Saka, N. and Suh, N. P., Delamination wear of dispersion-hardened alloys. Trans. ASME, Ser. B, J. Engng Indust., 1977,99,289-294. 19. Zhang, J. and Alpas, A. T., Delamination wear in ductile materials containing second phase particles. Mater. Sci. Engng, 1993, A160, 2.5-35. 20. Alpas, A. T. and Zhang, J., Effect of Sic particulate reinforcement on the dry sliding wear of aluminiumsilicon alloys (A356). Wear, 1992, 155, 83-104. 21. Venkataraman, B. and Sundararajan, G., The sliding wear behaviour of Al-Sic particulate composites-I. Macrobehaviour. Acta. Metall., 1996, 44, 451-460. 22. Bowden, F. P. and Tabor, D., The Friction and Lubrication of Solids, Vols I and II. Oxford University Press, Oxford, 1964. 23. Suh, N. P. and Sin, H.-C., The genesis of friction. Wear, 1981, 69, 91-114. 24. Blau, P. J., Test of a rule of mixtures for dry sliding friction of 52100 steel on an Al-Si-Cu alloy. Wear, 1982,81, 187-192. 25. Rana, F. and Stefanescu, D. M., Friction properties of Al-l .SPct Mg/SiC particulate metal-matrix composites. Metall. Trans. A, 1989, 2QA, 1564-1566. 26. Zhang, Z. F., Zhang, L. C. and Mai, Y.-W., Scratch studies of Al-Li alloy reinforced with SIC particles. Proc. 4th Int. Tribology Confi (Austrib 94), 5-8 Dec. 1994, Perth, Australia, pp. 249-254. 27. Wang, A. G. and Rack, H. J., Abrasive wear of silicon carbide particulateand whisker-reinforced 7091 aluminum matrix composites. Wear, 1991, 146, 337-348. 28. Saka, N. and Karalekas, D. P., Friction and wear of particle-reinforced metal-ceramic composites. Proc. Conf: Wear of Materials. ASME, New York, 1985, pp. 784-793. 29. Zhang, Z. F., Zhang, L. C. and Mai, Y.-W., Modelling friction and wear of scratching ceramic particlereinforced metal composites. Wear, 1994, 176, 231-237. 30. Jiang, J., Stott, F. H. and Stack, M. M., A mathematical model for sliding wear of metals at elevated temperatures. Wear, 1995,181-183,20-31. 31. Heilmann, P., Don, J., Sun, T. C., Rigney, D. A. and Glaeser, W. A., Sliding wear and transfer. Wear, 1983, 91,171-190. 32. Don, J., Unlubricated friction and wear in the Cu-Be system. PhD thesis, Ohio State University, 1982. 33. Subramanian, C., On mechanical mixing during dry sliding of aluminium-12.3wt%silicon alloy against copper. Wear, 1993,161,53-60. 34. Subramanian, C., Dry sliding wear of aluminium alloys: Wear mechanism maps and effects of the counterface. PhD thesis, University of Melbourne, 1989. 35. Iwai, Y., Yoneda, H. and Honda, T., Sliding wear behavior of SIC whisker-reinforced aluminum composite. Wear, 1995,181-183,594-602. 36. Antoniou, R. A., Brown, L. R. and Cashion, J. D., The unlubricated sliding of Al-Si alloys against steel: Mijssbauer spectroscopy and X-ray diffraction of wear debris. Acta Metall. Mater., 1994, 42, 3545-3553. 37. Rice, S. L., Nowotny, H. and Wayne, S. F., Characteristics of metallic subsurface zones in sliding and impact wear. Wear, 1981-1982,74,131-142. 38. Dautzenberg, J. H. and Zaat, J. H., Quantitative determination of deformation by sliding wear. Wear, 1973,23,9-19. 39. Marsh, D. M., Plastic flow in glass. Proc. R. Sot. Lond., 1964, A279,420-435. 40. Venkataraman, B. and Sundararajan, G., The sliding wear behaviour of Al-SIC particulate composites-II. The characterization of subsurface deformation and

41.

42.

43. 44.

45.

46.

47.

48.

49.

50.

51.

52.

53.

54.

55.

56.

57. 58.

59.

60.

61.

correlation with wear behaviour. Acta. Metall., 1996, 44, 461-473. Clarke, J. and Sarkar, A. D., Topographical features observed in a scanning electron microscopy study of aluminium alloy surfaces in sliding wear. Wear, 1981, 69, l-23. Pramila Bai, B. N. and Biswas, S. K., Characterization of dry sliding wear of Al-Si alloys. Wear, 1987, 120, 61-74. Sarkar, A. D., Wear of aluminium-silicon alloys. Wear, 1975,31,331-343. Clarke, J. and Sarkar, A. D., Wear characteristics of as-cast binary aluminium-silicon alloys. Wear, 1979, 54, 7-16. Reddy, S. A., Pramila Bai, B. N., Murthy, K. S. S. and Biswas, S. K., Wear and seizure of binary Al-Si alloys. Wear, 1994,171, 115-127. Eyre, T. S., Wear characteristics of castings used in internal-combustion engines. Met. Technol., 1984, 11, 81-90. Prasad, S. V. and Mecklenburg, K. R., Friction behavior of ceramic fiber-reinforced aluminum metal-matrix composites against a 440C steel counterface. Wear, 1993,162-164,47-56. Howell, G. J. and Ball, A., Dry sliding wear of particulate-reinforced aluminium alloys against automobile friction materials. Wear, 1995, 181-183, 379-390. Modi, 0. P., Prasad, B. K., Yegneswaran, A. H. and Vaidya, M. L., Dry sliding wear behaviour of squeeze cast aluminium alloy-silicon carbide composites. Mater. Sci. Engng, 1992, A151,235-245. Narayan, M., Surappa, M. K. and Pramila Bai, B. N., Dry sliding wear of Al alloy 2024-A&0, particle metal matrix composites. Wear, 1995, 181-183, 563-570. Reddy, S. A., Pramila Bai, B. N., Murthy, K. S. S. and Biswas, S. K., Mechanism of seizure of aluminiumsilicon alloys dry sliding against steel. Wear, 1995, 181-183,658-667. Sato, A. and Mehrabian, R., Aluminum matrix composites: Fabrication and properties. Metall. Trans. B, 1976, 7B, 443-451. Surappa, M. K., Prasad, S. V. and Rohatgi, P. K., Wear and abrasion of cast Al-alumina particle composites. Wear, 1982, 77, 295-312. Surappa, M. K. and Rohatgi, P. K., Preparation and properties of cast aluminium-ceramic particle composites. J. Mater. Sci., 1981, 16, 983-993. Hosking, F. M., Portillo, F. F., Wunderlin, R. and Mehrabian, R., Composites of aluminium alloys: Fabrication and wear behaviour. J. Mater. Sci., 1982, 17, 477-498. Chung, S. and Hwang, B. H., A microstructural study of the wear behaviour of SiCp/Al composites, Tribol. Znt., 1994,27,307-314. Anand, K. and Kishore, On the wear of aluminiumcorundum composites. Wear, 1983, 85,163-169. Pramila Bai, B. N., Ramasesh, B. S. and Surappa, M. K., Dry sliding wear of A356-Al-Sic, composites. Wear, 1992,157,295-304. Zhang, Z. F., Zhang, L. C. and Mai, Y.-W., Particle effects on friction and wear of aluminium matrix composites. J. Mater. Sci., 1995, 30, 5999-6004. Roy, M., Venkataraman, B., Bhanuprasad, V. V., Mahajan, Y. R. and Sundararajan, G., The effect of particulate reinforcement on the sliding wear behavior of aluminum matrix composites. Met&l. Trans. A, 1992, 23A, 2833-2847. Alpas, A. T. and Embury, J. D., Wear mechanisms in

Dry sliding

wear of aluminium

composites-a

review

435

62.

63.

64. 65.

66.

67.

68.

69.

70.

71. 72. 73.

74.

75.

76.

77. 78.

79.

80.

particle reinforced and laminated metal matrix composites. Proc. Conf on Wear of Materials, ASME, New York, 1991, pp. 159-166. Antoniou, R. and Borland, D. W., Mild wear of Al-S1 binary alloys during unlubricated sliding. Mater. Sci. Engng, 1987,93,57-72. Singh, L. and Alpas, A. T., Elevated temperature wear of A16061 and A16061-20%Al,O,. Scripta Metall. Mater., 1995, 32, 1099-1105. Welsh, N. C., Frictional heating and its influence on the wear of steel. J. Appl. Phys., 1957, 28, 960-968. Smith, A. F., The influence of surface oxidation and sliding speed on the unlubricated wear of 316 stainless steel at low load. Wear, 1985, 105,91-107. Wang, A. and Rack, H. J., Dry sliding wear in 2124 Al-SiC,/17-4 PH stainless steel systems. Wear, 1991, 147,355-374. Zhang, Z. F., Zhang, L. C. and Mai, Y.-W., Wear of ceramic particle-reinforced metal-matrix composites: Part I. Wear mechanisms, J. Mater. Sci.. 1995, 30, 1961-1966. Wang, A. and Rack, H. J., A statistical model for sliding wear of metals in metal/composite systems. Acta Metall., 1992, 40, 2301-2305. Krishna Kanth, V., Pramila Bai, B. N. and Biswas, S. K., Wear mechanisms in a hypereutectic aluminum silicon alloy sliding against steel. Scripta Metall., 1990, 24,267-272. Shivanath, R., Sengupta, P. K. and Eyre, T. S., Wear of aluminium-silicon alloys. Br. Poundryman, 1977, 70, 349-356. Razavizadeh, K. and Eyre, T. S., Oxidative wear of aluminium alloys. Wear, 1982, 79, 325-333. Antoniou, R. A. and Subramanian, C., Oxidation during dry sliding of aluminium alloys, unpublished research. Wang, D. Z., Peng, H. X., Liu, J. and Yao, C. K., Wear behaviour and microstructural changes of SiCw-Al composite under unlubricated sliding friction. Wear, 1995,184, 187-192. Bai, M.. Xue, Q., Wang, X., Wan, Y. and Liu, W., Wear mechanism of Sic whisker-reinforced 2024 aluminum alloy matrix composites in oscillating sliding wear tests. Wear, 1995,185,197-202. Ames, W. and Alpas, A. T., Wear mechanisms in hybrid composites of graphite-20 Pet SIC in A356 aluminum alloy (Al-7 Pet Si-0.3 Pet Mg). Met. Mater. Trans. A, 1995,26A, 85-98. Alpas, A. T. and Zhang, J., Wear rate transitions in cast aluminum-silicon alloys reinforced with Sic particles. Scripta Metall., 1992, 26, 505-509. Moustafa, S. F., Wear and wear mechanisms of Al-22%Si/Al,O,, composite. Wear, 1995, 189, 189-195. Lim, S. C., Liu, Y. and Tong. M. F., The effects of sliding condition and particle volume fraction on the unlubricated wear of aluminum alloy-SiC particle composites. Proc. Conf on Processing Properties and Applications of Metallic and Ceramic Materials, Birmingham. 7-10 Sept. 1992, pp. 485-490. Zhang, J. and Alpas, A. T., Wear regimes and transitions in A&O, particulate-reinforced aluminum alloys. Mater. Sci. Engng,1993, A161,273-284. Zhang, Z. F., Zhang, L. C. and Mai, Y.-W., Wear of

81.

82.

83.

84.

85.

86. 87.

88.

89.

90. 91.

92.

93.

94.

95.

96. 97.

98.

ceramic particle-reinforced metal-matrix composites: II. A model of adhesive wear. J. Mater. Sci., 1995, 30, 1967-1971. Wang, A. and Rack, H. J., Transition wear behavior of Sic-particulateand Sic-whisker-reinforced 7091 Al metal matrix composites. Mater. Sci. Engng, 1991, A147,211-224. Park, H. C., Wear behavior of hybrid metal matrix composite materials. Scripta Metall., 1992, 27, 465470. Kwok, J. K. M., Goh, H. S. and Lim, S. C., Effects of mechanical alloying on the friction and wear characteristics of Al-4.5Cu-15SiC particulate composites. Proc. 4th Int. Tribology Conf. (Austrib 1994), 5-8 Dec. 1994, Perth, Australia, pp. 241-247. Subramanian, C., Effects of sliding speed on the unlubricated wear behaviour of AI-12.3wt%Si alloy. Wear, 1991,151, 97-110. Zhao. D., Tuler, F. R. and Lloyd, D. J., Fracture at elevated temperatures in a particle reinforced composite. Acta Metall. Mater., 1994, 42, 2525-2533. Lim, S. C. and Ashby, M. F., Wear-mechanism maps. Acta Metall., 1987, 35, l-24. Atoniou, R. and Subramanian, C., Wear mechanism map for aluminium alloys. Scripta Metall., 1988, 22, 809-814. Bowden, F. P. and Persson, P. A., Deformation, heating and melting of solids in high-speed friction. Proc. R. Sot. Lond., 1961, A260, 433-458. Liu, Y., Asthana, R. and Rohatgi, P., A map for wear mechanisms in aluminum alloys. J. Mater. Sci., 1991, 26, 99-102. Wang, Y., Cao, L., Liu, G. S. and Yao, C. K., Acta Metall. Cornp. Sin., 1991, 8, 85. Takahashi, W., Nakanishi, M. and Kuwayama, T., Development of plasma transferred arc hard facing on titanium alloys using a new wear-resistant titanium alloy. Sumitomo Search, 1989, 39, 67-72. Matsuda, F., Nakata, K., Shimizu, S. and Nagai, K., Carbide addition on aluminum alloy surface by plasma transferred arc welding process. Trans. JWRI, 1990, 19, 81-89. Matsuda, F., Nakata, K., Lee, K.-C. and Lee, Y.-H., Formation of thicker hard alloy layer on surface of aluminum alloy by PTA overlaying with metal powder. Trans. JWRI, 1991,20,77-88. Matsuda, F., Nakata, K., Shimizu, S. and Nagai, K., Surface alloying of aluminium using plasma transferred arc welding with powder addition. Weld. Int., 1991, 5, 823-827. Matsuda, F., Nakata, K., Park, S. and Hashimoto, T., Improvement of wear resistance of aluminum alloyed surface by forming Si alloyed layer with PTA process. Trans. JWRI, 1993, 22,21-30. Bhansali, K. J. and Mehrabian, R., Abrasive wear of aluminum-matrix composites. /. Met., 1982, 32, 30-34. Ma, Z., Bi, J., Lu, Y., Shen, H. and Gao, Y., Abrasive wear of discontinuous SIC reinforced aluminum alloy composites. Wear, 1991,148,287-293. Edelson, B. I. and Baldwin, W. M., The effect of second phases on the mechanical properties of alloys. Trans. ASM, 1962,55,230-250.

Potrebbero piacerti anche