You are on page 1of 47
Sponsored by ‘The Research Committee on Lubrication An ASME Centennial Research Project The American Society of Mechanical Engineers United Engineering Center 345 East 47th Street New York, New York 10017 Wear Theory and Mechanisms J.F. Archard The University Leicester, England : = + a-ring is worn thin next to the finger with continual rubbing. Dripping water hollows a stone, a curved ploughshare, iron though it is, dwindles imperceptibly inthe furrow. We see the cobble stones of the highway worn by the feet of many wayfarers. The bronze statues by the city gates show thei right hands worn thin by the touch of all travellers who have greeted them in passing. We see that al these are being diminished since they are worn away. But to perceive what particles drop off at any particular time is a power grudged to us by our ungenerous sense of sight.” Lucretius (95-55 B.C.) De rerum natura, I NOMENCLATURE ‘Area of true contact ‘Area of apparent contact Cross-sectional area of plastically deformed res [Eq. (20)} Area of true contact at one asperity contact (Eq. (13a)] [ei =A] Radius of circular area of contact Lateral dimension of model asperity (Fig. 7) Depth of wear (Eq. (16)] Ratio of roughness to lubricant film thickness [Eq. (29b)] ‘Young’s moduli of two contacting bodies Reduced elastic modulus [Eq. (1d)] Frictional force Indentation hardness (D.P.N.) ‘Numerical constant [Eq. (11¢)] Coefficient of wear Coefficient in adhesive wear theory, K, = 3K, [Eqs. (14b), (15)] = Coefficient in abrasive wear theory, [Eqs. (24), (15)] K, = Coefficient in corrosive wear theory, [Eqs. (28), (28a)} 35 WEAR CONTROL HANDBOOK L = Sliding distance 4L = Increment of sliding distance for one asperity contact [Eq. (13¢)] Lana. Power law relation between area and load [Eq. (9b)] Pressure Maximum Hertz contact pressure [Eqs. (1b), (2)] Mean Hertz contact pressure (Eq. (1b)] Apparent contact pressure [Eq. (16)] Radial coordinates Radius of curvature = Relative radius of curvature [Eq. (1e)} ‘Shear strength of junction (Eqs. (7a, b)] ‘Time of sliding (life of bearing) (Eq. (17)) Transition loads in the wear of steels (Fig. 9) Velocities of surfaces in the kinematics of lubrication (Fig. 13) Volume of wear Incremental volume of wear from one asperity contact, V=E6V, (Bq. (130) Volume of plastically deformed material [Eq. (20)] = Sliding velocity ‘Advance of rigid surface in deformation of single model asperity Fig.) Normal load Increment of load borne by a single asperity contact (Eq. (13a)] Critical loads in elastic/plastic transition in contact mechanics Eqs. (4), (6a, b) and Fig. 5] Coordinates (Fig. 3) Yield stress in uniaxial tension (Eq. (5)] Depth of abrasive indentation (Fig. 12) Approach of bodies in elastic contact [Eq. (1e)] Correlation distance: lateral scaling factor in surface topography related to average wavelength (Fig. 3 and Eq, (12b)] Viscosity | Controlling viscosity in ehd lubrication (viscosity at atmospheric pressure and at ambient temperature of the surfaces) Semi-angle of abrasive indentor (Fig. 12) Mean value of Cot # for all abrasive grits Critical thickness of film in corrosive wear theory [Eq. (26b)] Ratio of ubricant film thickness to surface roughness [Eq. (29a)] Poisson’s ratios of contacting bodies | rms value of heights in a surface profile ms value of asperity heights in a model surface Alternative statements of the Plasticity Index (Eqs. (12a, b)] representing the probability of plastic deformation of a surface in its contact rity why reric ‘a +o WEAR THEORY AND MECHANISMS. ah rots | aes Bee] Has (a) ay —— tion | Le wear ei cS (SES) fasta [Eatery “Sat-acing ras| |reatgrna.| | Pager zise)| (MGteeatar| ellen] [lastonygraayrani)] era) wflegtion FIG.1 BASICTRIBOLOGY 1 INTRODUCTION Wear is one of a number of processes which occur when the surfaces of engineering components are loaded together and are subjected to sliding and/or rolling motion. Together with the allied subjects of friction and lubrication, it can be regarded as an integral part of the science and technology of interacting surfaces which has, in recent years acquired the collective title of tribology. Much of this science and technology is of recent origin, the result of work carried out since World ‘War Two. The main purpose of this first chapter is to present a summary of existing fundamental knowledge of the subject both as an introduction to the handbook as @ whole and as a suitable background for the chapters which follow. Figure 1, therefore, presents a simplified “*birds-eye view"’ of the subject of tribology within which context the subject of Wear will be examined. Figure 1 will first be used to provide a guide to the whole field. All engineering surfaces are rough. In tribological terms this is their first im- portant characteristic; it means that when they are loaded together they touch only ‘over a very small part of their apparent area of contact. Tribology therefore involves the mechanics of contact and it is usual to represent surface protuberences (or “asperities”” to use the common parlance of the subject) as parts of spheres. We shall therefore outline below the mechanics of the contact of spherical bodies. y WEAR CONTROL HANDBOOK TABLE 1 UNLUBRICATED WEAR RATES AND COEFFICIENTS OF FRICTION FOR VARIOUS MATERIALS Load 3.9 (0.4 keh) Speed 1.8 m/sec. The Stated Value ofthe Hardnes Is That ofthe Softer (Wearing) Material in Each Example. The Table Also Shows the Values of the Wear Cosicient(X) Calealsat - from Fa (1.15) MATERIALS Wear Rate Hardness Coefficients Coelicents — o-f mm? per (DPN) of Friction of Wear Wearing Su sliding) kg/mm? w. Mild Steet, Mil Se LsT x10" 186 x10 60/40 Leaded Brass Too! Stel 24x10 9s 6x 10-4 PTFE Too Stel 20% 5 2 x 10-5 Stelle 2 690 53x 10-5 Ferrite Stainless a 250 17x 10 Stel Polyethylene Tool Ste! 3 v 0s 13x10 Tungsten C Tungsten Carbide 02 1300 03s ix 10" Because the true area of contact is small, it follows that the contact stresses are high. It is this feature which gives tribology its special flavour and, indeed, its major scientific interest. We are concerned with events which are small in size but intense in scale; through the understanding of such events, one can interpret the processes of friction and wear, The second important characteristic of engineering surfaces is that they are contaminated. On earth (as distinct from outer space) perfectly clean surfaces can be prepared only by special techniques in very high vacuum systems. Naturally oc- curring contaminant films range from a single layer of molecules adsorbed from the atmosphere to much thicker oxide and other films formed by chemical reactions between the surface and its environment. The existence of naturally occurring contaminant films is crucially important in tribology; we shall see that even in unlubricated rubbing they reduce considerably the wear and damage from that which could occur if the contact were between perfectly clean rubbing surfaces. It has been implied above that the significant event in tribology is the contact of two asperities on the opposing surfaces and, most importantly, their contribution during the contact to the support of the load between the surfaces. The details of what happens during these contacts will be discussed later. However, it is important to note here that it is now generally agreed that the major force of friction is that necessary to shear such junctions; this is the adhesion theory of friction. A further result of asperity contacts is that during such events (or as a subsequent consequence of them) material is removed from the surfaces. The removal of such particles can ‘occur in a number of different ways by mechanisms which have probabilities varying over a very wide range. This is wear. Seen in the context of the whole field of tribology, shown in Fig. 1, the subject of ‘wear has a number of important characteristics. First, its magnitude can vary over ‘an enormous range. Table 1 is taken from early studies of the subject in which the uunlubricated wear rates and coefficients of friction for a range of materials were ‘measured. It will be seen that whilst the wear rates vary over a very wide range (almost 10°), the variations of the coefficients of friction are, by comparison, in significant (a factor of less than 4). This wide range is increased still further when lubrication and other wear reduction techniques, discussed later, are considered (see 38 are ajor sof nbe athe ions ring nin that ctof ation Is of tant that reher vying, xtof over hthe were range 1, ine when A Gee WEAR THEORY AND MECHANISMS, Table 3 below). Second, because of its wide range, in terms of engineering practice and of economics, wear is often the tribological subject of greatest significance, Energy losses arising from high friction, although important, are often of minor concern compared with the increased maintenance costs due to high wear rates or the expensive ““down-times” of large engineering assemblies caused by catastrophic wear failures. Finally, as the very brief introduction to the fundamentals of wear given at the end of the last paragraph implies, the fine details of the ways in which material is removed in rubbing is a complex subject and remains as one of the most difficult intellectual challenges facing the fundmental researcher in tribology. As will be shown, significant advances have been made and a clear framework for the subject has been established. Nevertheless, despite the passage of two thousand years, and the development of sophisticated physical instruments of high resolution, the closing sentence of the quotation at the head of this chapter still represents, in some significant areas of the subject, the essential problem for our basic un- derstanding. ‘Since this handbook is mainly concerned with wear, itis appropriate, at this stage, to discuss one further important feature of the wear measurements shown in Table 1. It will be seen that the rates of wear have been expressed as the volume removed ‘per unit sliding distance, Expressed in this way, the wear rate has dimensions of ‘area; since the true area of contact is the central concept in dry rubbing we can see, at once, that this is probably the correct method of expressing wear rates in fun- damental terms. Indeed, one can go further. An appropriate measure of the severity of wear is, therefore, the ratio of the wear rate, expressed in this way, to the true area of contact. This is the coefficient of wear, K, shown in the last column of Table 1h its theoretical derivation will be given and discussed later in this chapter. This concept of the wear coefficient will also persist throughout this handbook. Although practical situations occur in which high friction, or even high wear, is desired, the most common aim is their reduction. The major aim of this volume is the reduction of wear; lubrication, in its many forms, is the common method of achieving this. If we define lubrication in this rather broad fashion, it seems relevant to remark, at the outset, that the most common lubricants met in everyday life are the oxide films present upon the surfaces of metals used in common practice. Because of the nature of the interatomic metallic bond, clean metal surfaces almost always weld together in rubbing contact and are severely damaged. Thus the existence of oxide and other naturally occurring surface films represents the tribologist’s first wear-reducing agent. ‘At the other extreme, the ideal method of protection is the complete separation of the surfaces by a fluid film (Liquid or gas). Fluid film lubrication is classified in Fig, 1 and illustrated in Fig. 2. As shown in Fig. 1, a fluid film can be produced in two main ways. First, the fluid can be injected between the surfaces by a pump; this is hydrostatic or externally pressurized lubrication (Fig. 2a), a system which is par- ticularly valuable for the protection of stationary or slowly moving surfaces. Second, if a fluid exists between the surfaces, a pressure can be generated within it by the relative motion of the surfaces; this pressure can support the applied load. ‘Osborne Reynolds first showed, nearly a century ago, how this mechanism arose from the relative motion of the surfaces dragging fluid into a converging gap. In simple terms, the pressure thus generated increases with the velocity and with the viscosity of the lubricating fluid in the film separating the surfaces. This is the 9 WEAR CONTROL HANDBOOK £1G.2_ MECHANISMS OF FLUID FILM LUBRICATION (@) Hydrostatie (Externally Presurised) Lubrication, () Plain Journal Bearing (©) Tited Pad Bearing tation of Noaconforming Machine Element. {@) Close-up pf () with EHD Film Examples (b) and c) llasrateClasieal Sell-Acting Hydrodynamic Lubrication () Repr ‘mechanism of self-acting hydrodynamic lubrication. The most common ap- plications of both of these types of fluid film lubrication theory occur in situations {such as a plain journal bearing (Fig. 2b) or a tilted pad thrust bearing (Fig. 2c)], where the surfaces nominally conform and, in essence, the shape of the uid film is known. It is convenient, for what follows, to call such applications classical lubrication theory. In the terms of this handbook, the first essential design question for these situations is whether the lubricant film thickness is adequate to allow complete separation of the surfaces or whether some form of solid contact occurs mitigated, perhaps by some components of the lubricating fluid. A very different situation exists in another class of engineering component such as cams and their followers, gears and rolling element bearings. Here, because of the essential geometry of the system, the surfaces do not conform, but diverge rapidly from a single contact region which, under dry contact conditions and a negligibly small load, would be either a line of contact (say, spur gears represented in sim- plified form as the contact of parallel cylinders) or a point of contact (say, ball bearings represented in simplified form as the contact of spheres) (see Fig. 2d), When classical lubrication theory is applied to these situations, it forecasts values of film thickness of a magnitude comparable with the length of a single oil molecule. During the past 20 years, the theory of elasto-hydrodynamic Iubrication (ehd lubrication) has been developed and largely confirmed by experiments. In essence, it forecasts significantly thicker films than those given by classical theory (typical revised ehd values of film thickness are of the order to 1 jum or 40 qin,). The 40 a ap ations 20), film is assical estion allow pecurs uch as of the apidly ligibt y, ball 1. 2d). ues of lecule. a (ehd ance, it typical ). The WEAR THEORY AND MECHANISMS enhancement of the film thickness arises from a significant local deformation of the Surfaces under load and large increases in lubricant viscosity under the con: Sequential high local pressures (see Fig. 2«). Once again, in terms of this handbook, the essential preliminary question is whether the thickness of the lubricant film is fadequate to prevent significant solid contact and wear or only mitigated solid contact. Many situations occur when the ideal form of fluid film lubrication is impossible either because the presence of a fluid is inappropriate or because the conditions are too severe to support a fluid film. We call these conditions mitigated solid contact. To simplify our classification, we consider mitigation of the severity of these ‘contacts by the intervention of either a solid lubricant or by some component of the fluid. In recent years solid lubricants have played an increasing role in the solution of tribological problems; their particular, but by no means exclusive, fields of application are extremes of temperature, where lubricating fluids are unstable, and in lightly loaded long life components where renewal of a fluid lubricant is tiresome or costly. The mitigation of solid contact by naturally occurring components of fluid lubricants or by the addition of special additives has a much longer history and is included under the umbrella title of “boundary lubrication.”” Historically, this term was first used to explain lubricant efficiency under conditions where the classical Mechanisms of fluid film lubrication were thought not to occur. In more recent Years, the term boundary lubrication has come to be used mainly for those situations Where the molecular rather than the bulk properties of the lubricant play @ significant role. In earlier theories of boundary lubrication, physical adsorption of monolayers of the boundary lubricant species upon the surfaces was designated the major role. Later chemisorption, oxide films and multimolecular layers were all Shown to be significant. At the same time lubricant technologists, armed with these basic ideas, have developed lubricants sular, additives) for a range of par practical situations where lubrication was found to be difficult. E.P. (extreme pressure) additives were developed for conditions such as the automotive hypoid Tear axle drive, where the local contact pressures are high; in fact, it is now generally ‘agreed that it is the generated high temperature (not the high pressure) which is the important factor in the initiation of e.p. lubricant action, The array of such ad- ditives has increased with the arrival, upon the scene, of new lubrication problems » and ‘‘anti-wear” additives. As of and now includes such descriptions as “mild e. has been indicated above, the important feature of these additives is the pro increased protection of the surfaces against failure through severe damage or wear at the expense of the inclusion within the lubricant of a material of potentially higher ‘chemical reactivity with the surfaces. As we shall see, the design question is the obvious one of balance between the prevention of failure, on the one hand, and on the other, the producton of increased wear caused by the higher chemical reactivity of the environment. In what follows, we shall concentrate upon a more detailed exposition of those aspects of tribology of particular relevance to wear. The subject will be developed by considering first the dry contact of surfaces followed by a consideration of rubbing ‘under unlubricated conditions. This leads directly to a consideration of unlubricated Wear; at this stage the role of oxide and other naturally occuring surface films cannot be avoided. Quantitative theories of wear, which attempt to express rates of al WEAR CONTROL HANDBOOK | TPesmoertectleg Tum (al FIG.3. TYPICAL PROFILE OF A GROUND SURFACE, (4) As Obtained witha Stylus Profilomecer with Distorion othe RMS of the Helght Distibution (Wertcal Scaling Factor); 6°; Auto-Correation Function (Horizontal Sealing Factor, Related to the Average Wavelength) () Steepest Aspeity from (a (Indicated by Arrow) Reproduced Without Distortion, Also Shown Approximately to Scale, are Grain Size of Subsurface Metallic Stuctute (Distorted by Mechanical Working near the Surface) the Thiknes ofa Typical Oxide Fil is Represented by the Thickness ‘ofthe Line Representing the Surface ‘material removal in terms of experimental or design conditions will be considered and each of the major mechanisms of wear will be considered in turn. This 1.0, then plastic flow (albeit beyond the onset criterion of IY; in Fig. 5) occurs most commonly. "As an illustration of the physical meaning of the geometric aspects of the theory of the last paragraph, consider the extreme example of the two surface profiles shown in Fig. 8. These two profiles have the same distribution of heights (the same rms ot R, values). However, profile (b) has a structure of much shorter wavelengths and much smaller asperity curvatures, It should be clear that, in contact with another surface, a surface having profile (b) will have significantly higher contact pressures (and will be much more liable to damage) than a more gently curving surfee having a profile such as that shown in (a). ‘One important aspect of the concept of the plasticity index, and the theory associated with Eqs. (Il, 12) is that it suggests that an appropriate definition of Surface finish for tribological function requires more than one parameter. Since the introduction of the stylus profilometer, some 40 years ago, it has become traditional 3 WEAR CONTROL HANDBOOK to define surface finish simply in terms of the height distribution (rms or R, value). As will be seen later, this may be acceptable when considering the likelihood of damage or wear in lubricated systems for which a theoretical value of the lubricant film thickness can be calculated; at the design stage, this value can be compared with the value of the surface roughness, However, for dry and boundary lubricated systems the traditional definition of surface finish has been found to have its limitations. Two examples may now be cited to illustrate the potential utility of the new approach represented by the plasticity index of Eqs. (12a, b). Boundary lubricated systems usually fail by an abrupt transition from smooth sliding to severe metallic welding. Surface finish is one factor which affects the conditions for this transition and the traditional measure of surface finish (say, R,) does not provide a guide to this safe/fail boundary. However, it has been shown that, in simple laboratory experiments, the plasticity index of Eq, (12b) does provide a definition of this boundary. Clearly, in more practical engineering situations where failure under boundary lubrication conditions is likely, this approach to the role of surface finish merits consideration in design. In the wear of soft polymers on hard metallic counterfaces, the surface finish of the metal has an influence upon the wear rate. Again, the conventional definition of surface finish (say R,) does not provide an adequate explanation of the role of surface topography; it is found that the radius of curvature of the asperities has an influence, the wear rate being reduced as the radius of curvature of the asperities increases. Once again, this is in accord with the concepts of this section and points the way to a more meaningful consideration of surface topography in design. ‘As has been already indicated, the subject of the last part ofthis section is under active investigation at the time of writing. During the lifetime of this handbook, further developments will occur and therefore an attempt has been made to provide some pointers for the future. There is one further important consideration in this connection. Production engineers are showing marked interest in these new methods for the specification of surface finish, in part because they emerge from stylus profilometers linked to digital computers which are becoming increasingly available. In the future, designers are likely to be presented with these more complex alter- natives in the specification of surface finish. Therefore, it has been our purpose to present here such guidance as now exists upon the relationship of these specifications to tribological behavior. 6 CLASSIFICATION OF WEAR MECHANISMS ‘The major problem in the study of wear, both asa subject for academic research and in its consideration as an engineering problem, is the answer to the question: “how aze the worn particles removed.” The problem arises, in part, because of the microscopic nature of the processes which are often involved, However, much progress has been made and, in broad outline, there is general agreement about the major mechanisms which are involved in wear. Following upon the pioneer work for Burwell [8], who first attempted such a classification of wear mechanisms, and drawing upon subsequent work we shal use (both here and later inthis handbook) the classification set out in Table 2. This is fairly self-explanatory, but some further details of some of these mechanisms will be provided later. In what follows we shall briefly outline the nature of fundamental knowledge, s4 value), ood of bricant ed with vricated ave its r of the ay, R,) shown provide uations 1 to the ish of ition of role of has an perities points s under dbook, provide in this rethods 1 stylus ailable. x alter ose to f these esearch testion: 2of the + much ‘out the ork for as, and book) further wledge, WEAR THEORY AND MECHANISMS. TABLE2 CLASSIFICATION OF WEAR MECHANISMS Ts Adhesionand Transfer Materials weld at sliding asperity tps, is wansfered to the harder member, possibly grows in subsequent encoutes and is eventually removed by Fracture, fatigue or corrosion 2. Corrosion Fim Wear A fil formed by reetion withthe envionment or the lubricant i removed, by sliding 2. Cui [A sharp particle or asperty cuts chip. 4 Plastic Deformation The surface is worked plastically. Cracks form, grow, and coalesce form: ing wear particles 5. Suctace Fracture ‘Tf nominal stress exceeds the fracture stress of a brite material, particles can be formed by fracture. 6, Surface Resstions One material dissolves or diffuses into another, 7 Teacing Elastic material cen be torn by asharp indenter 8. Making High generated temperatures can cause wear by melting 9, Electrochemical ‘The difference in potential onthe surface due to moving Mid can cause a material topo into solution. 10, Fatigue The surface is worked elastically. Microcracks form, grow, and colasce forming wear partles using the example of two combinations of materials which have been studied in some detail. In part, using these examples as illustrations, we shall then draw certain general conclusions about mechanisms of wear, their dependence upon operating conditions, and the nature of transitions in wear behaviour. In this preliminary discussion, the major emphasis will be upon unlubricated wear since knowledge of this subject is more complete. Perhaps the most closely studied example of unlubricated wear is that of a 60/40 leaded brass pin on a hard ring of stellite or tool steel [9]. The soft brass was transferred to the ring and the wear debris consisted of fairly large particles of metallic brass. Using radioactive tracers and other techniques, the mechanism of wear was shown to be as follows: a. Discrete particles of brass were transferred to the ring, in time forming a fairly continuous layer. b. ‘There was no significant back-transfer of brass from this layer to the pin. c. On repeated rubbing the transferred brass particles formed aggregates in which form they were finally removed from the ring, probably by a fatigue mechanism, d. The whole wear process was in dynamic equilibrium so that the rate of wear (the rate of production of loose debris) was equal to the rate of transfer of brass from pin to ring; transfer (which was one of a sequence of events resulting in the final production of a worn particle) thus appeared to be the rate determining process for the wear mechanism as awhole, This particular example has been outlined in some detail because it gives very reproducible results, it has been subjected to detailed study and is reasonably well understood, It is not of any particular immediate engineering importance; indeed the magnitude of the surface damage and the rate of wear are excessive in terms of what is acceptable in practice. However, there are some further features of the wear of this combination of materials which have wider significance. First, Fein [10] has shown that under boundary lubrication conditions the wear mechanism is essentially similar but the wear rate is reduced to values more acceptable in engineering practice, Second, the regime of severe metallic wear which has been described operates over a central range of speeds. At the two extremes of speed the wear rate 3s WEAR CONTROL HANDBOOK Load that Wear yp mr siging 10 FIG.9 _ UNLUBRICATED WEAR RATES OF ANNEALED PLAIN CARBON STEEL (0.52 So C, 268 DPN) AS A FUNCTION OF THE LOAD The Characters of the Worn Volume Versus Sliding Distance Graphs fr Each ofthe Three Main Regimes of Wear Are Shown Inset. The Effect of Increasing the Hardness I Alo Shown falls by two orders of magnitude and the mechanism of wear is reduced to @ mild regime in which the worn particles are largely oxidised. At low speeds this occurs because the time between asperity contacts is sufficiently long to allow the regeneration of the protective oxide film. At high speeds essentially the same transition occurs, but in this situation it arises from the higher frictional tem- peratures and the consequential higher chemical reaction rates. Figure 9 shows, in somewhat simplified form, the wear rate of a pin of a plain carbon steel rubbing against a ring of the same material [9]. Both rubbing members were prepared in the soft annealed condition. This graph is a double logarithmic plot of wear rate versus load and is a type commonly used in this type of work since it allows the presentation of a very wide range of behaviour. The values of the wear rates presented here correspond to the values obtained after a period of sliding sufficiently long to establish equilibrium steady state conditions in which the worn volume increased linearly with the sliding distance. The character of the worn volume versus sliding distance graphs for each of the three main regimes of this example are also shown inset in Fig. 9. These results, although of some interest in their own right, are introduced here to illustrate a number of important features which are common to the wear of many metals. At low loads, below the transition load, 7, (Fig. 9), the wear rate (after an initial running-in period when equilibrium surface conditions were established) remains at @ relatively low level. In this regime, the surfaces remain smooth and relatively 56 mild curs the tem- plain abers ree it wear iding worn worn this ast in tures nitial ins at tively WEAR THEORY AND MECHANISMS undamaged, the worn debris is finely divided (say, 10 nm-1 zm) and almost entirely oxidised. This mechanism of wear has been called “mild wear.”” In discussing the allied problem of fretting, where essentially similar mechanisms occur, Waterhouse [9] lists the possible mechanisms for the removal of material as follows. ‘a. Removal of metal in finely divided form with possibly subsequent oxidation (mechanism 1 of Table 2). 'b. Removal of metal which when oxidised gives abrasive powder which con- tributes to the action (mechanism 3 of Table 2). . Oxidation of the metal surface and the removal of the oxide which exposes fresh metal for oxidation (mechanism 2 of Table 2). The last of these three mechanisms is one which most workers have thought to be significant in mild wear; indeed, some fairly elaborate theories for this process have been developed. The second mechanism has often been cited as important when metal oxides have a hard abrasive structure (AlO,, as an example). The first mechanism is more rarely thought to be significant under typical mild wear conditions since, as will be seen shortly, the existence of appreciable metal-to-metal contact usually results in the growth of these metallic junctions and the appearance af gross damage. Returning to Fig. 9, we see that at the transition load 7, there is a change in the ‘wear rate by more than two orders of magnitude. Above 7, the worn debris consists of relatively large metallic particles visible to the unaided eye, and the rubbing surfaces are severely torn and damaged. This is the severe metallic adhesive wear already discussed in connection with the wear of brass. Mild to severe wear trans- itions are common in the wear of metals ‘At a still higher load (T;, Fig. 8), a further transition in wear behaviour occurs; above T, although the inital wear is severe, mild wear is soon established. It has been shown that this mild wear regime is associated with the appearance, upon the rubbing surfaces of a hard layer caused by the high transient frictional “flash temperatures.” However; what is more relevant in the present connection is the consequence of changing the initia! hardness of the rubbing members by hardening and tempering (it will be recalled that in the experiments just described the specimens were in the annealed condition). As the inital hardness of the rubbing ‘members was increased the severe wear regime, between T, and T;, was reduced in its range until eventually, as shown in Fig. 9, when the hardness reached a value between 400 and 450 DPN the severe wear regime disappeared and mild wear persisted over the whole range of loads. This illustrates the role of material hardness in wear. With this brief and selective review, illustrating some aspects of more detailed investigations of wear, as a background we can now summarise the more important conclusions drawn from our present knowledge of wear; in particular, this must be seen against the background of the classification of Table 2. a. Transitions in wear behaviour can occur with changes in the operating con- ditions. These transitions can not only result in a change in the mechanism of wear, but sometimes can cause catastrophic increases in the rate of wear. In terms of engineering practice a complete change to severe adhesive metallic wear should be described as failure; the first objective of design should be to avoid this. b. The avoidance of severe wear of metals is essentially the avoidance of severe and persistent welding between the two rubbing members. Their compatibility (of which more will be said later) is therefore crucial. Clearly the rubbing of similar 7 WEAR CONTROL HANDBOOK materials should be avoided where possible; some particular materials (e.g., austenitic stainless steel) rubbing on themselves are known to be particularly susceptible to severe wear. c. Although quite common, sharp and clear-cut transitions in the wear ‘mechanism is not a universal pattern of behaviour. Regimes of mixed wear, e.8.. even mixed mild and severe wear, can occur; changes in the conditions (e,g., the environment) can then push the behaviour towards dominance of one or another mechanism. 4. Within any well defined regime of wear, the most common form of behaviour that the wear rate increases proportionally with the load. e. An increase in surface hardness is normally beneficial (in particular, with metals). I¢ is more likely to be helpful in the avoidance of severe wear than in any marked reduction in the rate of mild wear (see Fig. 9). £, With metals the role of oxide and other naturally occurring protective surface films is also very significant in the avoidance of severe wear. g. The environment in which wear occurs is therefore highly significant in avoiding failure through chemical reaction and the formation of protective surface films; however, the achievement of optimum conditions is more complex than this implies. The provision of an environment which is chemically too reactive can cause a marked increase in the rate of mild wear by the ready production and removal of the products of chemical reaction (mechanism 2 of Table 2). In these terms the formulation of some lubricant additives can be seen as a compromise between, on the one hand, the avoidance of severe wear and, on the other, the danger of in- creased corrosive wear. hh, Most discussion of wear mechanisms is concerned with the behaviour after the establishment of steady state equilibrium conditions. The nature of thse conditions often involves significant modification of the nature of the rubbing surfaces and these changes are almost always a vital part of the “rubbing system.”” The relatively gross changes involved in the establishment of a transferred film of brass in the example discussed earlier is an obvious example. The importance of running-in is also well known. In lubricated systems which approach the ideal of a full-fluid-film, modification of the surface profile, so as more closely to approach this ideal, is probably the major factor. But changes in the rubbing surfaces which occur during the establishment of steady state conditions are sometimes far more complex than in the two simple examples aready cited. The changes which occur in the copper surface of copper/graphite current collection devices may be cited as an example of great complexity. Nor, should it be emphasised, are these changes immediately obvious. In the rubbing of polymers (and also of dry bearing materials which in- clude polymers) on metals, a very thin film of polymer, invisible to the naked eye, is first established upon the metal surface; the presence, and indeed the nature, of this thin transferred film can play a crucial role in the subsequent wear behaviour. i. As explained in (d) above, within one mechanism of wear the most common (and as will be explained later, the expected) behaviour is that the wear rate is proportional to the load. Deviations from this pattern occur particularly when the ‘equilibrium surface conditions are changed. The most common cause of such deviations is the effect of frictional heating. The results are rarely as dramatic as the TT, transition in Fig. 9. A fuller discussion of frictional heating is given in Section 8 below. 58 Is B., ticularly ae wear ar, €., eg, the another ar, with nin any : surface icant in surface han this, an cause noval of sms the veen, on or of in after the ditions tees and slatively sin the ing-in is film, ideal, is = during than in copper mple of diately hich in- Jeye, is + of this ommon rate is hen the af such cas the zxtion 8 WEAR THEORY AND MECHANISMS. FIG.10 ADHESIVE WEAR: A SIMPLIFIED CONCEPT OF A UNIT EVENT IN RUBBING 7 ADHESIVE WEAR THEORY We now continue with the development of a quantitive theory of rubbing. The assumptions of the adhesion theory of friction given in Section 4 will be used to develop a similar theory of adhesive wear. Consider one event in rubbing, shown in Fig. 10 as the contact of two asperities. It will be assumed that during this contact the maximum area of contact is Bana? = 8/H (3a) where 697 is the maximum load support contributed by this asperity contact to the total load, W¥. Similarly, 64 is the maximum contribution to the total true area of contact, A. It is also assumed, as in theory of Section 4, that the asperity defor- mation is plastic. ‘Assume that the asperity contact results in a worn particle of volume 8¥. We expect that dimensions of this worn particle will be directly proportional to the size of the contact, Physical examination shows that, in general, wear debris is equi-axed in shape (ie., Worn particles are usually lumps of material roughly equal lengths in three dimensions rather than, say, layers). Thus, we expect 6V to be proportional to «@*: making the specific convenient assumption that the particle is hemispherical, ov: (138) Finally, we need to make some assumption about the duration (sliding distance) of this asperity contact. This sliding distance, 6L, will clearly be proportional to the size of the contact, a. A specific convenient assumption based upon a number of dif- ferent models is, a (130) It may be noted here that this equation can be taken to imply that the asperity contact exists for a total sliding distance 4a; a distance 24 the load is growing to its maximum value, 6; and a distance 2a in which the load declines again to zero. ‘We can now deduce the contribution of this asperity contact to the wear rate expressed, as explained earlier, as worn volume per unit sliding; thus using Eqs. (13a, b, ©) 59 WEAR CONTROL HANDBOOK 1 tage th 140) aa = Bearing in mind the essential need for conservation of total load support, i.e., the conservation of total true area of contact, we can write that for all areas of contact the total wear rate is V_yarv_t 77 lyre (aby Into Eq. (14b) has been introduced a constant of proportionality K, . Its significance will be discussed shortly; i will suffice here to explain that itis introduced to provide ‘agreement between theory and experiment. Examination of the derivation of Eqs. (14a, b) from Eqs. (13a, b, c) shows that the factor of one-third in Eq. (14b) is a simple consequence of the specific assumptions about the volume of the worn particle (Eq. (13b)] and the operative sliding distance (Eq. (130)]. Other assump- tions, within the essential spirit of the theory (ie., 6V « a!; AL & a), would Produce a value slightly differing from one-third. Therefore we put K = one-third K, and write w KAsky as) ‘The K of this equation thus becomes the wear coefficient as defined in the in- troduction and tabulated (as deduced from experiments) in Table 1. The magnitudes of typical values of K shown in this table shows very clearly that the difference, by a factor of one-third, between the formulations of Eqs. (14b) and (1S) is of no real significance, The very wide range of wear rates, which is so important a feature of the subject, must be understood as wide variations in the wear coefficient K, The foregoing theory was first stated in this form by Archard [11] and was based on preceding less precise analysis of the subject by Holm [12] and by Burwell and Strang [13]. It is now generally accepted as a suitable framework within which the Quantitative aspects of the subject can be discussed. Refinements of the theory [9] are essentially concerned with specific interpretations which can be given to the wear coefficient K. Equation (15) is a statement of wear theory in fundamental terms, but is not in a form particularly suitable for the designer who is interested in the depth of wear, d. If we divide both sides of Eq. (15) by the apparent area of contact, zy, we obtain theresult t-@G2)- Ge cc) where P is the mean or nominal pressure on the bearing (the load divided by the apparent area). If we now express the sliding distance L as the product of the sliding velocity, V, and the time, r, Eq. (16) becomes, after rearrangement, an Equation (17) expresses the life (#) of a bearing in terms of an acceptable depth of wear and a materials selection factor (H/K). The life is then inversely proportional to the PV factor. The use of this factor in the selection of dry bearings is described 0 (14a) the tact ab) ‘ance ovide Eqs. lisa worn amp- ould third as, tudes bya real re of, ased and the y 9) wear ina nd. stain a9) 1 the ding ay thof onal ‘ibed WEAR THEORY AND MECHANISMS. later in Section 13. It will also be shown how it will require modification, for ‘example, in the light of deviations from the simple wear law of Eq. (15) caused by frictional heating. Itis appropriate to consider at this stage the material factors of importance in the reduction of adhesive wear. In particular, itis appropriate to consider ways in which adhesive wear in its worst form (Le., severe metallic wear—see mechanism 1 of Table 2) can be avoided. As we have indicated, avoidance of such severe wear or failure is the first stage in the tribological selection of materials. We list below a series of general principles, together with some comments upon them. a, Material Hardness. As Eqs. (15) and (17) imply, a high value of the hardness is usually helpful. However, as Eqs. (11), (12) and the associated discussion suggests, the probable factor of significance is not hardness but the ratio (H/E) which represents the yield or proof strain of the material. However, with metals and alloys, ‘Hcommonly varies more widely than E. . Surface Hardness. What matters is the hardness of a layer which encompasses. the major contact stresses. As a consequence, a whole variety of surface hardening techniques have been developed from the early well-tried examples of case- carburizing and nitriding to more recent methods for the deposition of hard wear~ resistant layers too numerous to be listed here. [En passant, it is worth while noting other factors worth considering in such layers: the surface finish achieved and the possible avoidance of abrasive wear, the depth of the layer compared with the ex- pected depth of wear in the lifetime of the component, the bond between the layer ‘and the substrate and its behaviour under stress. c. Avoidance of Welding. For this reason the use of same material for both rubbing members should be avoided where possible. By the same token ‘metallurgically similar materials should also, in general, be avoided; this is usually termed compatability 4d. Compatability. High mutual solubility between pure metals clearly favours welding and makes the combination incompatible. It has been argued (see the discussion of Ref. [9a]) that it is the ratio of the interfacial energy of adhesion W to the surface energies, 7, and 7%, of two contacting metals which is significant. For the two close-packed metallic structures, welding occurs more readily with FCC than HCP structures; perhaps the more ready availability of planes of easy shear with the FCC structure (compared with the HCP) is the relevant factor here. All these, and other factors have been supported by simple wear experiments often at low speeds of sliding. However, correlations exist between some of these suggested factors and which of them is of crucial importance is difficult to determine. Moreover, it is sometimes difficult to assign values of some of these factors (e.g., surface energy), particularly when we take into account the complexity of the surface conditions attained under equilibrium rubbing conditions. , Surface treatments for the reduction of welding. Other surface treatments involve the deposition or production of soft layers; better known examples of such methods include compounds of sulphur and phosphorous. These layers are relatively thin. They are most widely used in the presence of a lubricant and their ‘major function is to prevent scuffing or severe failure of otherwise vulnerable components. It is possible that their main contribution is to enable the surfaces to survive during the earlier stages of running-in and thus to ensure the achievement of ‘an acceptable state of hydrodynamic lubrication. The softness of the layers and their 6 WEAR CONTROL HANDBOOK FIG.11 PLASTIC DEFORMATION UNDER A SLIDING INDENTOR Indentor Is Sliding ina Direction Normal othe Plane ofthe Figure. consequential low shear stress reduces friction at asperity contacts which might otherwise contribute to a catastrophic and cumulative increase in damage, The essential point of the theory of adhesive wear (and indeed, of theories ap- plicable to other mechanisms) is that the wear rate, expressed as volume removed per unit sliding distance, is proportional to the load and independent of the sliding velocity and the apparent area of contact. Thus it mirrors the adhesion theory of friction presented in Section 4, Wear rate is proportional to the load because wear rate is proportional to area of true contact and area of true contact is proportional to the load. Into this theory has been introduced a wear coefficient, K. Therefore, before proceeding to a discussion of other mechanisms of wear, it is appropriate to Pause for a discussion of wear theory and, in particular, to consider the significance Of the wear coefficient, K. 8 DISCUSSION OF WEAR THEORIES {tis appropirate to start afresh by considering the problem of wear by means of an approach familiar to engineers: i.e., nondimensional analysis. Simplify the problem by assuming that we are dealing with an isothermal system and confine attention, at the outset, to a single combination of rubbing materials. It then seems reasonable to assume that the volume of wear, V, is a function of the sliding distance, L, the hardness of the softer material, H, and the sliding velocity, V; i.e, (L, W, H.%) as) This approach has been adopted by Shaw [14], who goes on to point out that if the usual methods of nondimensional analysis are used, the sliding velocity, ‘V, drops ‘out and one obtains the result [22] = come = 7 iw which is the familiar wear Eq. (15) in a rearranged form. Shaw goes on to argue that the K factor is a nondimensional parameter of central significance. It therefore requires a physical interpretation. In fluid mechanics, Reynolds’ number is a similar concept; its physical interpretation is that it is Proportional to the ratio of inertial to viscous forces. A physical interpretation of K. a ofan olem nat leto the ag) f the ops a9 atral it is of K WEAR THEORY AND MECHANISMS, was then obtained by considering an indentor (Fig. 11) loaded, in the fully plastic range, under a load, W, into the material of hardness, 1, and then slid through distance L. Consider, in this operation, the total volume of material which is sub- jected to plastic deformation. This volume will be LA’, where A’ is the cross- sectional area of the plasticity deformed region beneath the indentor (see Fig. 11). The plastically deformed volume of material, V,, ina sliding distance L is therefore V=A'L 0 Now by dimensional similarity the area A’ is proportional to the true area of contact, A, formed by the indentor under the load, W, i.c., as before A=WiH 7) and therefore WL i vat en ‘Thus, Eq. (19) may be rewritten in the physically meaningful form elas (98) Shaw thus argues that the wear coefficient X is proportional to the ratio of the worn volume V to the volume of material V, which would be plastically deformed under an indentor loaded, under fully plastic conditions, travelling the same sliding distance, L, under the same load, W. Examination of the above derivation and those of Sections 7 above and 9 below show that they are not too different in their central concept. In essence, for example, Section 7 derives K as the ratio of the worn volume to that which would have been produced if every single event produced a worn particle. In all of these derivations, Kis the ratio of the worn volume to another volume which is rhe worst which we could conceive of under the same conditions (Le,, the same values of W, L, H), In the light of this interpretation we can draw some wry comfort. We find that when we take common materials and rub them in air without undue precautions the values of K, as we have seen, ate very low. In- deed, this is very fortunate for us and makes tribological engineering a viable proposition, ‘A reexamination of the derivation of the theory of Section 7 shows that if every event (asperity contact) during rubbing resulted in the production of a worn particle, then the value of K (or more specifically K,) would be unity. Therefore, the most common interpretation of the meaning of K is that it simply represents the proportion of all the asperity contacts which result in the production of a worn particle. The consequences of this interpretation of K are important and revealing. It will be seen that K values are low, ranging down to 10-* — 10-* even for unlubricated rubbing. Assume a value of 10-*, which is unacceptably high for most engineering practice. This means that only one in a thousand events results in a worn particle whilst 999 events occur without damage. All events contribute to the fric- tion, which is therefore dominated by the events which occur without damage. Wear, on the other hand, is dominated by the K factor which can vary widely. We can see, at once, why wear varies more widely than friction and why, in general, they are not correlated (see Table 1). a WEAR CONTROL HANDBOOK A second important consequence of this interpretation of K has already been drawn at the start of Section 5; this is that in continuous rubbing, each part of the surface is rubbed many times and therefore the majority of the contacts will be elastic, This should not be taken to upset the central principles of tribology as outlined in Fig. 1. The contact stresses remain high. Indeed, if the initial defor- ‘mation of asperities is plastic, then one would expect them to run-in so that the subsequent contacts operate close to the elastic limit; according to the principles of Fig. 5 the contact stresses should then be of the order of 0.4. One of the few at- tempts to deduce directly the value of the contact stresses was made many years ago by Holm [15]. His results are rarely quoted by reviewers yet in the light of Section 5, and the above comments, it is remarkable to find that he deduced contact pressures of the order of 0.5H. ‘The principles involved in the discussion of the last paragraph also provide the response to a possible objection to the adhesive wear theory of Section 7. As we have seen, comparison of this theory with measured wear rates leads to the conclusion that much asperity deformation is elastic. Yet, paradoxically, the theory itself is based upon the assumption of plastic deformation; indeed, the appearance of H in the RHS of Eq. (15) arises essentially from this assumption about the magnitude of the contact pressure (as the central portion of the equation shows). The fact that the pressures are of the same order as, (but slightly lower than) the hardness, H, simply introduces a small correction factor which is of minor importance compared with the very small magnitudes of the K values (and the very wide range of their variation). This is of the same minor significance as the factor or one-third which ‘was discused below Eq. (15) above. Theories of wear, such as those of Section 7, forecast that the wear rate is proportional to the load and independent of the sliding speed and the apparent area of contact. These simple laws do not hold universally even when the mechanism of wear remains unchanged and no major transition occurs. The most common cause of disturbance of this simple law is frictional heating. It has two effects: microscopic and macroscopic. In Fig, 9 the T; transition arose from the microscopic form of behaviour, the very high temperatures (so-called ‘flash temperatures”) generated at the small area of true contact by the dissipation of frictional energy. In the specific example of the T; transition it was calculated that the magnitude of these tem- peratures was in excess of 750 C, the temperature necesary to austenitise the material in the contact region and to harden it by the consequential rapid cooling. The duration of these effects are typically of the order of a millisecond or less. This, example illustrates the intense, transient, and very local nature of flash temperature effects. Macroscopic effects of frictional heating are probably of greater consequence in engineering practice. Such effects may be calculated by assuming generation of heat over the whole of the rubbing surfaces and its transfer to the environment by the full range of heat transfer mechanisms. Whether macroscopic or microscopic effects are of greater influence in oxidation and other reactions with the environment is a matter of some controversy. However, it is certainly true that these macroscopic effects are a major limiting influence in many types of bearing; for example, later in this handbook it will be shown how the upper limiting conditions of load and speed of some types of dry bearings may be determined by these considerations. We conclude this discussion of wear theories by returning to the theme of the 64 WEAR THEORY AND MECHANISMS sy been TABLES WEAR CORFFICIENTS (K) AND ANTIWEAR NUMBERS (411) FOR A RANGE tof the [OF SITUATIONS (FROM FEIN, REF. 16) will be 7 logy as BEARING MATERIALS ENVIRONMENT 1 defor- — ——— 7 tole Wear Surface OpposingSurace Atmosphere Lbican k_aWy siples of 52100 Steel s2100Steel Dry Argon ‘None ox? 20° ae Mil Sct ia in None 23x10 26 a (60/40 Leaded Brass ‘Tool Steel Air None 20x 103 27 FS ag0 S2100Stet s2leosiel Dry Ae Nene roxio? 30 sation 5, sielite Tool Ste! ‘ir None taxis 47 ressures 2100 Ses s2ioosesl Air, = Gycohesane «RA IDF SL PTFE Tool Stel air None £3x 10-8 Su idee, s2100 tet suioosel Air, == Parafinicool «3.2 10-7 6S ie the | Polyethylene “Tool Steal, Air None 43x18 73 wehave Auminiom Bronze CrburisedSteel Air Gear Lubricant «2510-16 aelusion CarbureaStec. —CarburiedStesl Ae Gear Oi Lex 10-7 78 itself is 1 2100 Ste s200Ste = Air_—Paraflne OWV/TCP 3.3.x 10-7 8S ofHin COSicel___52100Stel Nic EngineO <2.0 10°! 91 tude of that the ‘, simply central significance of the K factor as a statement of the value of the wear “ed with behaviour, a given combination of materials and environment. Fein [16] has pointed of their out that as a figure of merit the K factor has two disadvantages. First, as we have which already indicated the range of K values is very large indeed. Second, its values are inverted, excellent wear behaviour corresponding to low values of K. He therefore © rate is ‘suggests the adoption of an Anti Wear Number (A WN) defined by ent area anism of AWN= —l0810K 2) on cause ‘The results of Fein’s approach is shown in Table 3 taken from his publication. It will roscopic be seen that itis appropriate to express the AVIV in a number of only two digits form of since this is justified by the pr. arated at specific 9. ABRASIVE WEAR T ese tem- ‘material \ If the harder surface is rough or if hard abrasive particles are present, abrasive ing, The ‘wear may be the dominant mechanism. In the theory of abrasive wear, its assumed ae that each abrading clement is so sharp (rather than rounded) so as to indent the 7 opposing surface. We will assume here that the abrasive has a conical shape (Fig. 12) perature ° of semicone angle 6. Under a load 81 the abrasive particle indents the opposing surface to a depth z. This load is supported only on the leading half of the contact as uence in it moves forward. Then from the geometric assumptions (Fig. 12) 1 of heat , iam aw ire Tanto 230) Yeots are Also, from the geometric assumptions of Fig. 12, the volume of material 6” vent is a , removed in sliding a distance 8L is fouls av=(2 Tan 9)8L 3b) + Tater in nd speed. Then combining Eqs. (23a, b) one obtains ay _¢2cotey aw te of the ay ae 230) 65 WEAR CONTROL HANDBOOK FIG. 2. TYPICAL GEOMETRIC ASSUMPTIONS OF ABRASIVE WEAR THEORY As before, aking the wear rate forall abrasive paris, assuming cotinty of load support and introducing constant of proportion K, ore tx v 2Cot0) W =«, eo where Cot@ is an average value for all abrasive particles. This can be rewritten in the teneralised form of wear equation alrendydaevsed, te as) Comparison of Eqs. (24 and 15) shows that in abrasive wear the wear coefficient Kis the product of a geometric factor [2Cot6/z] and a constant of proportionality K,, which may again be interpreted as the proportion of all events resulting in the production of worn particles. The geometric factor is simply a function of the ‘assumed shape of the abrasive indentor (Fig. 12) and can vary over a narrow range depending on differing possible assumptions (e.g., pyramidal indentors in different orientations), Values of this geometric factor are of the order of 10-'; Cot8 also rep- resents the mean slope of the resultant abraded surface, which is of this same order (see Section 2 above), Abrasive wear theory, as formulated here, has said little or nothing about the way in which the material is removed. The assumption implicit in the theory (as shown in Fig. 12) is simply that material corresponding to the area of the indentor projected in the direction of motion is removed, leaving a groove. The mechanics of abrasion is certainly more complex than this. Scratch tests with simple indentors show that material is often not removed but is ploughed, leaving a furrow with displaced ‘material piled up on either side of the groove (scanning electron micrographs of abraded surfaces show similar details). Analysis of abrasion shows that it is a much less efficient process than might be expected; typical deduced values of K, Eq. (24) are of the order of 10-'. Samuels and coworkers have made a careful examination 66 ly of a) nthe as, cient tality athe f the ange arent rep order way vain edin that aaced 1s of uch (24) ation WEAR THEORY AND MECHANISMS of the abrasion mechanism using abrasive papers [9]. They conclude that this value of K; arises from the fact that only 10 percent of the abrasive grits have the fa- vourable geometry or “attack angle”” (favourable rake angle in metal cutting ter- minology) which is necessary for the removal of material. The rest of the grits, itis suggested, plough without material removal, It remains to be shown that these broad conclusions are the only features affecting the efficiency of abrasion and apply to!a wider field of abrasion studies. ‘Most fundamental studies of abrasive wear have been made using materials or conditions (e.g., abrasive papers) which are such that abrasion is deliberately fostered. Although the undesirable occurrence of abrasive wear in engineering practice has been less widely studied, the conditions under which it can occur are fairly well understood, First, one has the combination of a soft material rubbing against a rough hard counterface. In this situation, the most important factor in reducing abrasive wear is the reduction in the roughness of the hard counterface. However, its roughness, expressed in the conventional way as an rms or R, value, is not necessarily an adequate guide, What matters, as studies of abrasion and the theory of abrasive wear show, is the sharpness of the asperities on the abrading surface; this will determine whether they indent and abrade the opposing surface. (Un this respect, an analysis of surface roughness of the type given in Section $ is probably appropriate.) It might be noted here that methods of producing a finer finish in common usage result in not only lower values of rms or R,, but also provide surfaces with less sharp asperities (ie., lower slopes). Perhaps the most common form of abrasive wear experienced in engineering practice is three-body abrasion, i.e., abrasion caused by loose abrasive material between the rubbing surfaces. Such abrasive material is of two main types. The first is adventitious material, conveniently described as “‘dirt.”” Some tribologists con- cerned with the analysis of failures claim that this is the single most common cause of major tribological failures. The remedies are simple and obvious: the installation and maintenance of efficient sealing systems (also filtration systems when lubricants fare present) and the emphasis upon the appropriate meticulous standards of cleanliness in installation. ‘The second form of three body abrasion is sometimes less easy to anticipate and arises from abrasive material generated by the rubbing surface themselves. The example of Al,0, has already been cited. Clearly an exploration of the properties of the oxides from the rubbing materials is appropriate (other products of chemical reaction are relevant in other environments). Although conventional hardness values (ay, DPN) of oxides are rarely available, their Mhos hardness values often are and these can be interpreted to indicate their ability to scratch and abrade other materials ul 10 FATIGUE WEAR THEORY Fatigue wear exists in two forms, macroscopic and microscopic. The macroscopic form can occur in nonconforming machine elements in the form of pitting or rolling contact fatigue. The major forms of this type of wear are so severe as to constitute failure. In general, it follows the principles which one would expect from the discussion of contact mechanics given in Section 3. It occurs when the contact stresses approach the elastic limit (W, of Fig. 5). As in other forms of fatigue, the a WEAR CONTROL HANDBOOK number of stress cycles necessary to cause failure decreases with increasing stress. (Typically in a Hertzian contact the number of cycles of failure varies as a power of the load of the order of ~3.) A typical failure consists of pits of material removed from the surface, The depth of material removed usually closely corresponds to the Position of the maximum shear stress as explained in the discussion between Eqs. (4) and (5) of Section 3. Fatigue wear on a microscopic scale is similar to that just described except that it is associated with individual asperity contacts rather than with the single large region of the nonconforming machine element as in the example just described, We can, therefore, apply the same principles adopted in the adhesive wear theory of Section 7. Once again, the derivation follows the same general path, the volume 6 removed at an individual event is proportional to a where a is the radius of the individual area of contact; similarly, the sliding distance 8. is proportional to a giving as before v Sr alas (14a) and summing for all contacts (14b) Y akA=K 15) 7 aKa! c As in the earlier discussion of adhesive wear theory, we have ignored the factor of one-third in Eq. (14b) as being negligible compared with the very small values of K (and their very wide range of variation). Again we have incorporated the hardness, A, into Eq. (15) as an approximate statement of the contact pressure despite the fact that the true pressures will be slightly lower. Once again, as in the discussion of adhesive wear theory, it is convenient to in- erpret K as the proportion of all contacts which contribute worn particles However, we can now invert this statement and give it a specific meaning. Thus, 1 k This interpretation immediately reveals the attractive nature of fatigue wear theory as an explanation of K factors. For contacts repeatedly loaded to stresses close to the clastic limit, the number of stress cycles lies in the range 10) to 10° or more. This brings wear phenomena immediately into line with another main-stream aspect of the mechanical behaviour of materials. From this theory, methods of reducing fatigue wear follow logically. The fatigue behaviour of materials is of central significance (i.., cycles to failure versus stress characteristics). However, some care should be made in making such deductions too literal. Fatigue behaviour, on a macroscopic scale, is markedly influenced by surface effects such as defects. Therefore fatigue wear, being fatigue on a microscopic scale, will be similarly influenced by such factors as the presence of an oxide film. Clearly surface topography can influence contact stresses and its role will be important, particularly the topography of a hard surface rubbing against a softer counterface. Therefore the role of the factors discussed in Section $ are probably important, For example, the example of the wear of polymers on hard counterfaces number of stress cycles to failure es, 6 stress. wer of noved to the 4s. (4) that it region ecan, ection noved vidual ng as (14a) (ase) as) tor of sof kK ness, te fact to in ticles. es) heory tothe . This ect of atigue stress astoo face scale, ill be softer bably “faces WEAR THEORY AND MECHANISMS discussed below Eqs. (12) above may well be an example of the role of surface topography in fatigue wear. Finally, the role of the environment in fatigue should not be ignored. It has been shown that the presence of boundary additives has an influence upon normal fatigue (presumably through surface effects); it therefore seems likely that they would similarly influence fatigue wear. 11 CORROSION WEAR THEORY ‘The principles of corrosive wear have already been discussed in Section 6. It has been indicated that an important example of corrosive wear occurs in those situations where additives, or other materials, are deliberately introduced to prevent severe metallic adhesive wear. Protective films are formed by reactions between these additives and the rubbing materials and they are removed by rubbing (mechanism 2 of Table 2). Corrosive wear also constitutes a significant element of the mild wear of metals; in this case the worn debris consists largely of oxides which have formed upon the surfaces and have been removed by rubbing, Theories of corrosive wear can be quite complex (see Ref. [9a] and its discussion). However, to illustrate the most important principles involved, we shall confine our treatment to a very simple analysis similar to the adhesive wear theory of Section 7. We shall assume the following: a. A rubbing situation consisting of asperity contacts such as that used in Section ». A reactive environment which produces a slow growth of protective films upon the surfaces c. These films remain protective and undamaged until they reach a critica thickness, A, when they are liable to be removed by rubbing. 4. The growth of the protective films is a conventional thermally activated rate process. Proceeding as in the theory of adhesive wear (Section 7) the following equations replace Eqs. (13a, b, c) of that section for one individual event. The maximum area of contact is fama? aH (26a) ‘The volume of the worn particle (a layer of thickness Nis bV=ra?d 260) and, as before, the sliding distance is SL=20 (260) ‘Then the contribution to the wear rate from this event is ra) and the total wear rate is 7) Tr where, as before, we introduce a factor K, representing the proportion of all events contributing to the worn volume. Now, if as explained in Section 5, an increase in o WEAR CONTROL HANDBOOK the load results in a proportional increase in the number of contacts (n), their mean size (a) remaining constant, Acnre?, naw Then the total wear rate becomes ae a cn Dae Lea (ana 9 which is essentially the wear law of Eq, (15). We note that the coefficient of wear is given by KA Kak (28a) Which is @ constant for our assumed conditions. In the assumptions of this simplified theory, the factor in Eq. (28a) most liable to variation is K, , the proportion of events Which produce a worn particle. Now, in the above assumptions, a worn particle will be produced when the protective film reaches its critical value, \. Factors which will increase K, (and thus both K and the wear rate) will therefore include the following: @. An increase in the chemical reactivity of the environment. ». An increase in the chemical reactivity of the rubbing materials to the given environment. & An increase in the temperature. This effect will be most marked because for a thermally activated process there will be an exponential increase with the tem. perature. Although grossly simplified, this theory shows the essential elements of corrosive ‘wear. Corrosive wear is usually present to ensure protection against severe metallic Wear (cither by design or, in the case of atmospheric oxygen, almost unwittingly). But, as we have seen, an increase in reactivity either through a change in the chemical compositions of the environment (e.g., additive) or by relatively small increases in the temperature may cause unacceptably large increases in the rate of corrosive wear, Some of the principles and pitfalls ofthis mechanism of wear can be illustrated by SSlerence to a common lubricant additive, zine dialkyl-tricresyl phosphate (ZDTP), Only after its initial use as an oxidation inhibitor was it found to have E.P. Properties suitable for hard steel-steel combinations such as the hypoid rear axle Asain, later stil, it was found that the E.P. efficiency of the material depended critically upon the particular molecular architecture chosen from the variety of alternatives available for this class of material Inthe classical example of the hypoid ear axle, the probable mode of action of such an E.P. additive is to protect the surfaces against damage during running-in and during subsequent overloading or nisalignment and thus to allow a large element of hydrodynamic (or, strictly, ¢lastohydrodynamic) load support. At a still later stage, it was found that addition of this additive to fluids caused severe wear in the bronze-steel rubbing combination of pumps. The E.P. action and running-in which occurred in the hypoid stecl_stee! combination did not occur under the milder pump conditions; as a consequence, the steel surfaces of the pumps remained rough and abrasive wear persisted. 12, LUBRICATED WEAR Before discussing wear under lubricated conditions, we recall the main ” ‘cts (n), their mean 8) officient of wear is 28a) 3 of this simplified portion of events worn particle will tors which will de the following: ls to the given sd because for @ ve with the tem- ents of corrosive i severe metalic ost unwittingly), 2 change in the relatively small €s in the rate of beillustrated by sphate (ZDTP), {to have EP, ‘oid rear axle. terial depended the variety of leofthe hypoid to protect the overloading or € (or, strictly, Uthat addition sg combination >oid steel-stee! asequence, the all the main WEAR THEORY AND MECHANISMS mechanisms of lubrication which were briefly discussed in Section 1 and summarised in Figs. 1 and 2. Broadly speaking, it will be recalled that lubricated systems can be divided into nominally conforming machine elements (Figs. 2a, b, c) to which classical theories of lubrication apply, and nonconforming machine elements (Figs. 2d, ¢) to which theories of ehd lubrication apply. In all of these situations it is possible to calculate a value of the lubricant film thickness which separates the surfaces and methods of making such calculations are given elsewhere in this handbook. From the point of view of wear under lubricated conditions, the first important criterion is the value of this film thickness compared with the combined roughness of the rubbing surfaces. This is expressed in two ways, the specific film thickness, A, ‘or the D ratio; one is the reciprocal of the other. Thus: thickness Le ee 9a) Combined surface roughness = or pu Combined surface roughness sw) Thubricant film thickness In these expressions the combined surface roughness is obtained from the rms or R, values (sometimes peak to valley values have been used). However, whatever the ‘method of measuring the roughness, the essential concept isthe same, For values of ‘A large compared with unity (small values of D), the probability of some solid contact between the surfaces is small and therefore there should be no problems of ‘wear. For values of A small compared with unity (large values of D), some solid contact through the lubricant film is expected and wear or even severe damage may be possible. Most nominally conforming machine elements work successfully with a generous value of lubricant film thickness and a corresponding high value of A. However, there are a number of possible situations where problems can occur and three of these are worthy of brief comment 4. In journal bearings (Fig. 2b) the shaft runs off centre and thus provides a converging-diverging gap. The positive pressure distribution occurs mainly in the converging gap; in the diverging gap a liquid lubricant film breaks up to form air ‘bubbles (cavitation). This can cause cavitation damage to soft bearing alloys. ». Even with high values of A, adventitious material (.e., “dirt”) can bridge the sap and cause damage. Some examples of severe damage have been recorded [17] ‘The remedies are similar to those listed in the discussion of abrasive wear. Moreover, the selection of compatible materials is not entirely irrelevant even in these examples of designs with high values of A. ©. Some conforming machine elements (¢.g., some face scales) operate without any designed tapered film, which is the basis of self-acting lubrication. Some clement of hydrodynamic action does arise in these situations, possibly arising from waviness of the surfaces. However, in these poorly lubricated systems the com- patability of the rubbing surfaces becomes of still greater importance. The factors discussed in Section 7, above, provide a useful first guide. Nonconforming machine elements generally operate with lower values of the lubricant film thickness (lower values of A) and with more concentrated loads. In order to understand the susceptibility of such mechanisms to wear and failure, itis n WEAR CONTROL HANDBOOK (al Ib) id FIG.13 KINEMATICS AND HYDRODYNAMIC LUBRICATION OF NONCONFORMING SURFACES, Situations (@, (6 and (©) Occur at Light Loads andthe Essetial Kinematics Conditions of Thee Lubrication are Identical Situation, (€) Represents Situation (a) at a Heavier Load When EHD Lubrication Occurs, vital first to explore the role of kinematics of mechanisms in the generation of Iubricant pressures by hydrodynamic action. We shall set out the important essential principles by consideration of the simplified example of two cylinders in rolling/sliding contact or the similar example of a cylinder and a flat surface. Consider, first, loads so light that the generated pressures are insufficient either to deform the surfaces or to cause significant increases in the lubricant viscosity. These conditions are those of classical hydrodynamic lubrication; they occur rarely in nonconforming engineering mechanisms (vane-pumps, perhaps), but the following discussion will illustrate certain vital principles of more general applicabilit In Fig. 13(@) two discs each of radius 2R are loaded together in the presence of a lubricant and rotate with surface velocities u, and 13. As shown, the directions of 1, and w, have been chosen such that they assist each other in dragging lubricant into the convergent gap between the discs. Thus the effective entraining velocity for hydrodynamic action is, (uy tu) For reasons which will emerge shortly, this velocity is sometimes replaced by ¥ (1; + u;), as this is the effective rolling velocity. Also, in this system the sliding velocity, which determines the main dissipation of viscous energy, (a This sliding velocity is obviously important in the wear and damage of the surfaces, since it represents the relative motion of the surfaces Now consider the system shown in Fig. 13(b) in which a cylinder of radius R, having a surface velocity of u, slides and rolls upon a flat surface which is moving with a velocity of w;. Close examination shows that the kinematics and hydrodynamic action of this sytem is, essentially, the same as that of the two cylinders of Fig. 13(a). The shape of the oil-filled gap between the surfaces is the ‘same (those parts of the surfaces close to the minimum film thickness are the only regions of importance since effective pressure generation is confined to parts of the n of tial Fto ese ‘ing fa fu nto for (uy ling sR ving and two, the only the WEAR THEORY AND MECHANISMS oil film whose thickness is comparable with the minimum value). The motions of the ‘wo surfaces are the same in the two examples. Now consider the example shown in Fig. 13(c) in which a cylinder of radius R both rotates with a surface velocity u, and, at the same time, moves over the stationary flat surface with a velocity u, which is opposite in direction to that of the flat in Fig, 13(b). Once again, consideration will show that this situation of Fig. 13(c) is, essentially, similar to those of Figs. 10(a), (b). Here the important criterion is the ‘motion of the surfaces relative to the load which in this last example moves over the flat surface. In ai these examples the hydrodynamic action is determined by (i, + 1), the sliding velocity is (u, ~ u,), and the load transverses the (wo surfaces with velocities u, and u;. We also note that in all these examples the special case u, = 1, corresponds to pure rolling. The only minor difference between these examples which could be of significance in, say, rolling contact fatigue, is that there are differences in the periods between the repeated application of the load to a given point on the surface. Fig. 13(d) shows the situation of Fig. 13(a) at a heavier load under conditions of elastohydrodynamic (ehd) lubrication, The arguments advanced above about the relevance of the kinematics still apply and the arrangements of Figs. 13(b), (¢) under a heavier load would be essentially similar to that of Fig. 13(d). As Fig. 13(d) shows, ‘one important difference introduced by the imposition of a heavier load is that the surfaces are deformed; their shape is almost the same as that which would occur under the same load in the absence of a lubricant. Thus, in the conditions of ehd lubrication, two important regions can be observed. a. A convergent entry region in which the mechanisms of classical hydrodynamic lubrication are largely operative. At the end of this convergent region the generated hydrodynamic pressures have reached such a high value that there is a significant rise in the lubricant viscosity; this rise in viscosity continues, in a cumulative fashion, as the lubricant passes through the rising pressures of the subsequent parallel region. The criterion that the lubricant viscosity must reach a high value at this boundary between the two regions means that itis in the convergent entry region Where the value of the overall film thickness is determined. b, The convergent region is followed by an almost parallel region where the pressures are close to the Hertzian values obtained in dry contact; the lubricant viscosities here are greatly increased (typically by factors of 10? to 10°) as a result of the high pressures. The film thickness is maintained at an almost constant value by these large changes in viscosity. At the same time, this region supports almost all of the applied load, the convergent entry region making a negligible contribution to load support, The high values of viscosity in this region also result in relatively high values of the coefficient of friction when sliding occurs (~0.05 compared with typical values of, say, 10-? under conditions of classical hydrodynamic lubrication). ‘The divergent outlet region is of no major practical significance; the film cavitates close to the outlet edge of the parallel region and there is no full fluid film to con- tribute to load support, It will be obvious that the principles of the kinematics of machine elements discussed in connection with Figs. 13(a), (b), (c) also apply under the ehd conditions of Fig, 13(4). Thus, for a steel mechanism and a hydrocarbon lubricant, the film thickness will, as expected, be decided in the convergent entry region and will be dependent on the factors which are listed in Table 4. This also shows the role of 2 WEAR CONTROL HANDBOOK TABLES DESIG Ty ehabricant fim thickness i) 1.1 fnereases strongly with enraining velocity, (uy + u2) 1.2 Increases strongly with controlling viscosity, np (viscosity of lubricant a and at ambient temperature ofthe device). 1.3 Increases strongly with pressure dependence of viscosity 14 Increases with relative radius of curvature. 1.5. Decreases only slighty with increasing load and decreasing elastic modulus FACTORS IN LUBRICATED NONCONFORMING MACHINE ELEMENTS" tmospherie pressure 2, Combined surface roughness 2.1 Increased roughnes ncreases chances of solid contact and consequential wear of filure. 2.2. propriate parameters ato of roughness to film thickness (Eqs. (298, b)]. 3, Loud 3.1 Increased load increases streses, both main Hertzian stress and asperity stress if contact occurs 53.2 Thus if solid contact oovurs (se 2.2 above) increased load ean cause increased chances of damage 3.3 Increased stresses increase chances of fatigue wear both on a macroscopic and on a microscopic seale ce Section 10) 34 Increased load can increase dissipation and thus raises ambient temperature of device; hence a fection in yp and film thickness (se 1.2, above). 3.5 Increased load increases the value ofthe transient lash temperatures. 4, Sliding velocity (4 ~ 43) 4.1 Increase in (i ~ my) increases dissipation and thus raises ambient temperature of dev ‘reduction in yp and film thicknes (se 1.2 above). 4.2 Apuet from 4.1, (u, ~ u:) has no influence per se upon flim thickness. 43 solid contacts occur (ee 2.2 above) increased value of (uy ~ ua) increases chances of wear oF failare arising from them, 4.4 Increased value of (u, ~ ui) increases he val J of the transient ash temperatures. 5. Lubricant additives 5.1 Boundary lubrication properties of lubricant and of additives (including E.P, action) reduces ‘chances of wer or damage arising from solid contacts (See 2.2 above). 5.2 Additives increase chanets of corrosive wear (se Section 11) 5.3 Additives need choosing with care bearing in mind al the rubbing combinations involved (see Section 1). 6. Material properties of solid surfaces 6.1 Fatigue properties influence fatigue wear (Setion 10) £62 Compatibility influences lability to welding and damage (Se (See main text for explanation and definitions) other factors in the performance of nonconforming machine elements which will be discussed elsewhere. For the sake of completeness it is important to state briefly that a form of ehd lubrication occurs with materials of low elastic modulus, i.e., in rubber seals, when the pressures are so low that there is no significant increase in viscosity. In these situations, the convergent entry region again generates a vital pressure at the edge of the nearly parallel load-bearing region which now ensures that this region can act as a tilted pad with a very small tilt. The film thickness again depends strongly on (wu, + 1) and jp, but the strong dependence on the pressure-viscosity coefficient is replaced by a strong dependence upon the elastic modulus of the solid members. The film thickness is, again, almost independent of the load. ” of hd en se of as is he WEAR THEORY AND MECHANISMS, ‘A more complete explanation of these mechanisms of lubrication will be given elsewhere in this handbook and this will include methods for the calculation of the film thickness. The broad knowledge of the mechanism of chd lubrication has existed for the past 20 years or so, and its impact upon our understanding of the behaviour of gears and rolling clement bearings has been well established. Its impact in some other areas has been less well marked. For example, it is surprising to find that a generalised analysis of kinematics of some automotive cams and followers has appeared only in the very recent past [18]. The kinematics of cams is, of course, rather more complex than that of other mechanisms, but itis of interest to find that this analysis shows that the poor wear behaviour of some cam mechanisms at certain points of their cycle of operation corresponds fairly closely to the predictions of the chd film thickness: i.e., high wear corresponds to low values of (u, + 3) in an analysis of the type associated with Fig. 13 and Table 4. (On the other hand, most rolling element bearings operate with an adequate value of the lubricant film thickness and their kinematics are well understood. The fun- damental limitations upon the range of their performance under ideal conditions is usually rolling contact fatigue; as indicated in Section 10, the cycles to failure of such bearings varies as the load to a power of the order of —3. The full range of wear and failure mechanisms for rolling element bearings is well documented in the literature [19] Similarly, the kinematics of gears and the relationship of ehd lubrication to their successful behaviour is fairly well understood. Major failures of gearing are usually of two main types. Particularly with softer materials, macroscopic rolling contact fatigue of the type discussed briefly in Section 10 is a major limitation on their range of performance. In the content of this discussion, it is relevant to note that in rolling contact fatigue the number of cycles to failure is (along with other factors such as stress) a function of the A ratio or D ratio (Eqs. (29a, b)] ‘The other major form of failure of gears is scuffing or scoring which arises from severe and persistent metallic contact. The explanation of scuffing is a matter of some controversy (20, 21] which lies outside the range of this chapter. Briefly, one explanation postulates a critical value of the total contact temperature [20], i.e., ambient temperature plus transient flash temperature (see Table 4). The second, and more recent, explanation [21] attempts a theory of ehd lubrication which takes into account the role of surface roughness. ‘A Jess well documented cause of distress in gears (and some other forms of nonconforming machine elements) is excessive wear and the conditions under which this occurs seem less well defined. One example of this which is easily understood in terms of the fundamentals of this chapter is the use of a hard-soft gear material combination in which the roughness of the harder gear exceeds the chd film thickness. Under these conditions, the hard gear takes an excessively long time to run-in and thus severe solid contact through the lubricant film is continued well beyond the normal running-in period. Indeed, the surface of the softer gear may be roughened and damaged to such an extent that running-in is never completed before ‘wear has exceeded acceptable limits. 13 SUMMARY AND CONCLUSIONS ‘One major purpose of this chapter has been to present existing fundamental mB WEAR CONTROL HANDBOOK Prbssure mat i ig aaa Fe Ln te a cil Getta ‘ah SHOWING THE PERFORMANCE RANGE OF DIFFERENT. DUCED FROM FUNDAMENTAL CONSIDERATIONS. (@) Dry Rubbing Beavngs (6) Rolling Element Bearings (o Self Acting Lubricated Besrings knowledge of the subject as an introduction to the chapters which follow. The in~ tention is that this information can later be clothed in more detail so as to be of greater direct value to practising engineers and designers. Both to summarise some of the important information discussed earlier and to illustrate its use in practice, we will now conduct a simplified exercise; it is intended to show how these fun- damentals determine the range of conditions (essentially load and speed) over which three major types of bearings can be used, Plain journal-type bearings for a rotating shaft will be considered and we shall discuss only the following types of bearing; dry rubbing bearings, rolling bearings and self-acting hydrodynamically lubricated journal bearings. Of course, these are not the only types of bearing which are possible, but the choice will be limited to these three types in order to simplify the treatment. Thus, this chapter on fundamentals will end at the starting point of bearing design by suggesting the selection of bearing type as the first stage of that process (22, 23) DRY RUBBING DESIGN The equation of wear for dry rubbing rearranged in its form suitable for the designer is, from Section 7, Lele en ‘This shows that for a given material combination [represented by (H/K)] and a given acceptible depth of wear (d), the life of the bearing, (0 is inversely proportional to the (P) factor, ‘V being the rubbing speed and P the nominal pressure on the bearing surface. ‘Thus, for a dry bearing of given size and material combination, the acceptable range of operating conditions can be represented on a plot of load versus speed, When plotted using logarithmic scales, the result is as shown schematically in Fie, 14(a), the safe operating range being represented by the area beneath the line of slope —1. At the two extremes of load and speed, this simple picture is modified. As the speed is decreased, the acceptable load rises, but eventually would reach an tunaceeptably high value determined by strength and other considerations. In practice, most dry bearing materials with acceptable PV (or H/K) factors contain a Significant proportion of a polymer and the maximum acceptable load is determined of oh ry ad re he of at he ” to he he wus of AS In aa ved WEAR THEORY AND MECHANISMS by extrusion of the polymer. Similarly, at low loads and high speeds the heating effects increase and eventually the polymer would soften and be extruded, These two additional limiting factors lead to the typical limiting load/speed characteristics for ‘dry rubbing bearing shown in Fig. 14(a). ROLLING BEARINGS It has been explained in Section 11 that the major fundamental limitation in the operating conditons of rolling bearings is rolling contact fatigue and that the fatigue life, in cycles, decreases with load to a power of the order of 3. Thus, for @ given acceptable life, in running time, the acceptable load falls as a power of the speed of the order of —1/3. This type of basic behaviour is represented in Fig. 14(b). Again there are additional limitations at the two extremes. At the lowest speed, the ac ceptable loads may reach values which are well beyond the elastic limit (as shown by W, in Fig. 5) and thus there exists a maximum possible load beyond which unac- ceptable indentation (or Brinelling) of the rolling surfaces would occur. At the hhighest speeds, the centrifugal forces on the rolling elements become comparable ‘with the imposed forces and this consideration provides a basic upper speed limit for a bearing of a given size. Thus, in simple form, the limiting load/speed characteristic of a typical rolling bearing assumes the form shown in Fig. 14(b). SELF-ACTING HYDRODYNAMIC PLAIN JOURNAL BEARING In the classical theory of hydrodynamic lubrication, for a journal bearing of given goometry (i.e., given diameter, clearance and eccentricity, all of which specifies @ ‘minimum oil film thickness) the pressures increase linearly with the speed and are also proportional to the viscosity of the lubricant. At moderately low speeds the pressures, and thus the load (for a given acceptable minimum film thickness), in- crease linearly with speed as shown in Fig. 14(c). However, as the speed is increased still further, this increase in load is modified by increases in the temperature which cause reductions in the viscosity. The increase of acceptable load with speed is reduced, and eventually the load falls with further increases in speed. Thus one obtains the typical limiting load/speed characteristic of the form shown in Fig 14(c). The upper limit of shaft speed will be determined by bursting under cen- trifugal forces ‘When numerical data for typical bearings of each type is used, and the results for the three types of bearings are superposed, one obtains a simple design chart of the type shown in Fig. 15. Data for three sizes of shaft are included, i.e., for diameters of 6.3 mm (% in.), 50 mm (2 in.) and 250 mm (10 in.). This chart should not be taken as precisely definitive, but simply as a broad guide to the range of conditions under which each type of beating is appropriate. Clearly the boundaries of the operating range of any type of bearing can be extended by a more precise design study. In any event, such a study must represent the next stage of the design process. Moreover, other factors will also enter into the considerations at this stage. For example, for the relatively light loads and low speeds used in the bearings of domestic machines, economic considerations will be of crucial importance. The costs and operating characteristics of a range of dry bearing materials will need to be considered. The operating characteristics of such material will commonly be ex- pressed in the form of Eq, (17) or Fig. 14(a. WEAR CONTROL HANDBOOK Load in Shaft frequency in reun FIG. 15 A GUIDE TO THESELECTION OF BEARING TYPE ‘Adapted from References [22] and [23] and Deduced Using Fig. 14. The Shaft Diameters (Three ‘Sizes) Are Shown at the Left Hand Side === Dry Rubbing Bearings Rolling Element Beasinss ~ aoa. Self-Acting Lubricated Bearings To conclude, we will summarise briefly the most important topics which have been discussed in this chapter. 1. Rubbing wear arises from the intimate contact of the interacting surfaces at small areas of true contact. 2. ‘The mechanisms of wear are many and diverse. 3. In any one regime of wear, the worn volume normally increases linearly with the load and with the distance of sliding. For a given distance of sliding, the worn volume is commonly independent of the sliding speed. 4. Inthe formulation of these laws of wear, the most influential factor is the wear coefficient K; its most common interpret that it represents the portion of all rubbing events (asperity contacts), which result in the production of a worn particle. 5. Wear rates, even under unlubricated conditions, vary over a very wide range, typically by a million or more. This is primarily the consequence of a very wide range of values of K which therefore reflects the merit of the given combination of materials in resisting wear. 6. Deviations from these laws of wear can occur; the most common cause of such deviations are the effects of frictional heating. 7. Changes in the regime of wear can occur with changes in the operating con- ditions. Such changes can result in massive and unacceptably large changes in the rate of wear. sat with vorn wear of all icle inge, wide mot such con- athe WEAR THEORY AND MECHANISMS 8. With metals and alloys severe wear involves extensive metal-to-metal contact and a catastrophic increase in wear. 9, Severe wear is unacceptable in engineering terms and should be regarded as a failure. Its avoidance is the first objective of tribological design. 10. A number of different mechanisms of lubrication exist. The most efficacious is the generation of a full fuid film. 11, Full fluid film lubrication can occur through the motion of the surfaces (self- acting hydrodynamic lubrication) or by the injection of fluid under pressure (ex- ternally pressurised lubrication). For most of such systems, methods exist for the calculations of the minimum lubricant film thickness. This can then be compared ‘with the roughness of the surfaces. 12. To avoid failures and to reduce wear the character of the total system should be considered. This includes the materials of the two interacting surfaces, the lubricant, if present, and the environment (including dirt). 13. The role of dirt should not be ignored. It has caused catastrophic failures in apparently well designed lubricated systems of the type designated in [11] above. 14, The most important element in any total system is the compatibility of the materials of the two interacting surfaces. 15, Compatibility can be improved by surface treatments and by the use of lubricant additives. 16. When additives are used to protect the most vulnerable part of an engineering system, their influences on other parts of the device must have careful consideration. 17. The results of wear tests under simplified laboratory conditions constitute a valuable guide to the relative performance of different combinations of materials; however, their application in quantitative terms to engineering systems should always be viewed with caution. 18. Where possible tests on the engineering device itself should be made; where this is not possible, laboratory tests should simulate, as closely as possible, the conditions of the engineering device. REFERENCES 11 Tabor, D, The Hardness of Metts, Clarendon Press, Oxford 191 (2) There docs not seem to be a single simplified version ofthe fll range ofthe theory of contact mechanics as presented here. Aspects ofthe Hertz east theory sto be found in Timoshenko, S. P. and Goodie, 1.N., Theory of lastiiy, McGraw-Hil, New York, 1970, pp. 409-820. Other detils are 0 be found in Ret.) [5] and [6 [31 Holm, R, Electric Contacts Handbook, Springer-Verlag, Berlin, 1988, pp. 199-207, (4) (@) Merchant, M. E., “The Mechanism of Static Friction," Journal of Applied Physics, Vol. 11, 1940p. 230, () Ernst, H. and Merchant, M. E, “Surface Finish of Clean Metal—a Basic Factor inthe Metal Cutting Process," Friction and Surface Finish, MIT Report No. 15, 1940 (issue with a new introduction and Bibliogeaphy, MIT Press, Cambridge, Mass 1969), pp. 76-101. (1 Bowden, F.P., and Tabor, D, The Friction and Lubriation of Solids, Part I, Clarendon Press, Oxford, 1954, pp 90-10 (6) Bowden, FP, end Tabor, D., The Frietion and Lubrication of Solids, Part I, Clarendon Press, Oxford, 1964, pp. 52-86, (7) Reviews ofthis subject andthe developments which follow are to be found inthe folowing: (a) Archard, J. F, "Surface Topography and Tribology,” Tribology Intemational, Vo. 9, 1978, pp. 213-220. (©) Archard, 1 F,, Hunt, R. T., and Onions, R. A., “Stylus Profilometry and the Contact of 9 1 8) 9) an na 03] ual us) us, 07 us us} (20) eu ry pa} WEAR CONTROL HANDBOOK Surfaces,” The Mechanics of the Contact bet ween Deformable Bodies, (.U.T.A.M.) Symposium, Bd. A.S-de Pater and J.J, Kalker) Delt, Unversity Pres, 1975, pp. 282-303. (© Greeawood, J. A., and Williamson, J. B. P., "Contact of Nominally Fiat Surfaces," Proc. Roy. Soe (Lon Series A, Vol. 298, 1966, pp. 300-319 ‘Burwell, J.T, "Survey of Possible Wear Mechanisms," Wear, Vol 1 1957-58, 9p. 19-181 ‘A summary of this and other work on mechanisms of weat with references to original papers i t0 be found in the Following (@) Archard, J. F,, "Wear," Interdscptinary Approach to Frietion and Wear, Ed. P.M. Ki [NASA sp- 181, Washington, D.C., 1968, pp. 267-333 (including discussion). () Peterson, M. By, "Mechanisms of Wear," Boundary Lubricaions, An Appraisal of World Literature, BF. P’ Ling, E.E, Klaus and R. 8. Fei, Americen Society of Mechanical Engineers, 1969, pp. 19-37 Fein, Rr S., discussion of Reference [9 ‘Archard, J-F., "Contact and Rubbing of Fist Surfaces," Journal of Applied Physics, Vol. 24, 1983, pp. 981-88, Holm, R. Reference [3], pp. 242-254, Burwell, J.T, and Strang, C.D.,""On the Empirical Law of Adhesive Wear,” Journal of Applied Physics, Vol.33, 1983, pp. 18-30. ‘Shaw M. C., "Dimensional Analysis for Wear System Holm, R., Reference [3], pp. 59-63 Fein, R.'S., “AWN—A Proposed Qualitative Measure of Wear Protection,” Lubrication Engineering, Vol. 31,1978, pp. $81-S82. (@ Davson, P. H. and Fidler, F,, “The Behaviour of Chromium Stels in Large High Speed Bearings,” AEI Enginering, VO. 2 (No.2), 1962, pp. 54-62. () Dawson, P. Hand Fidler, F., “Wire Wool Type Bearing Failures; the Formation of Wire Wool,” Proceedings of the Instation of Mechanical Enginers, Vol, 18D, Part I, 1965-66, PP 513-8. ( Fowie, 7. ,, “Problems in the Lubrication Systems of Turbo Machinery,” Proceeding ofthe Institution of Mechanical Enginears, Vol. 16, 1972, pp. 705-716. Dyson, A., "'Elastohydrodynamie Lubrication and Wear of Cams Bearing Against Cylindrical Tappets,” Incerational Automotive Engineing Congress, Society of Automotive Engineers Warrendale, Pa., 1977, Paper No. 770018. (To be published in The Society of Automotive Engineers Transactions.) Tallian, T. E., "Roling Contact Failure Contol through Lubreation,” Proceedings of the I stitution of Mechanical Engineers, Vol. 182, Part 3A, pp 208-236. {@) Blok, H., “The Postulate about the Constancy of Scoring Temperature,” Inerisciplinary “Approach 10 the Lubrication of Concentrated Contacts, Ed. P. M. Ku, NASA SP-237, ‘Washington, D,C., 1970 pp. 153-248 (including discussion). () Cameron, A. and Gentle, C. R., "Mechanics and Thermodynamics in Lubrication,” fn terdiseiplinary Approach to Liquid Lubricant Technology, Ed. P, M. Ku, NASA SP-318. Washington, D.C., 1973, pp. 315-364 (including discussion) (@) Dyson, A., “The Failure of EHL Lubrication of Citcumferemially Ground Diss,” Proceedings ofthe Institution of Mechanical Engineers, Vol 190, Part , 1976, pp. 699-711 (©) Cheng, H. S. and Dyson, ., “EHD Lubrication of Ciscumfecentially Ground Dis “American Soe, of Lubrication Eng. Transactions, Vol. 21, 1976, pp. 25-40. Neale, M.J., “The Design of Plain Bearings,” Proceedings of the Institution of Mechanical Engineers, Vol 19, Part 3D, 1964-65, pp. 1-1 ‘General Guide to the Choice of Bearing Type, Engineering Sciences Data Unit, London, 1965 Wear, Vol 3, 1977, pp. 263-266, Engineering Sciences Dato, Wem, No. 65007, 80

You might also like