Anda di halaman 1dari 12

Contact mechanics and friction

This appendix summarizes, in the context of metal machining, understanding of the stressesthat occur at the contacts between sliding bodies. These stresses,with materials , responses to them, are responsible for materials' friction (and wear). All engineering components -for example slideways, gears, bearings, and cutting tools -have rough surfaces, characteristic of how they are made. When such surfaces are loaded together, they touch first at their high spots. Figure A3.l is a schematic view of two rough surfaces placed in contact under a load W, the top one sliding to the right under the action of a friction force F. FigureA3.l(a) shows a contact, the material properties and roughness of which are such that the surfaces have deformed to bring the direction of sliding into the planes of the real areas of contact Ar. Resistance to sliding then comes from the surface shear stresses s. Friction that arises from shear stresses is called adhesive friction. If the real areas of contact on average support a normal contact stressPr' the adhesive coefficient of friction .Uais given by
s .Ua= ~ (A3.1)

Figure A3.l(b) shows surfaces for which the real areas of contact are inclined to the sliding direction. Each contact is divided into two parts, ahead of (leading) and behind

Fig. A3.1 Friction caused (a) by shear stresses 5 and (b) by direct stresses p

364 Appendix 3 (trailing) the real contact mean normal n. Even in the absence of surface shear stresses, a resistance to sliding occurs if the normal forces on the leading and trailing portions of the contacts differ from one another. Friction arising from contact normal stresses is called deformation friction. If, on average, the normal stress p1on the leading part of a contact of sub-area A, is inclined at 0, to the direction of the load W, and on the trailing part of the contact the equivalent variables are p,, A, and e, force resolution in the directions of Wand , F give the deformation friction coefficient pd as
pd

plAlsinQ1- p&in0, plAlcosOl +p,+os0,

(A3.2a)

Special cases occur. If the contact is symmetrical (PI = pt;A, =A,; 6l, = et), equation (A3.2a) simplifies to p d = 0: this is the case of perfectly elastic deformation. At the other extreme, when the indenting surface plastically scratches (abrades) the other, there may be no trailing portion contact: A, = 0. Then, equation (A3.2a) becomes pd = tan el (A3.2b)

This type of deformation friction (abrasion of metals) is of most relevance to this book. (There is a third situation, of visco-elastic contact, intermediate between perfectly elastic and totally plastic contact, when pd may be shown to depend on both tan 0, and tan 6, the loss factor for the contact deformation cycle.) Equation (A3.1) shows that adhesive friction depends mainly on material properties s and p , although, as will become clear, p , also depends on surface contact geometry. By contrast, equation (A3.2b) shows that abrasive deformation friction depends mainly on surface geometry, insofar as the angle 6, is the same as the slope of the leading part of the contact, but this could be modified by material properties if, for example, the real pressure distribution over A, is not uniform. The main focus of this appendix is to review how the friction coefficient varies with material properties and contact geometry, in adhesive and deformation friction conditions, and when both act together. Two further points can usefully be introduced before proceeding with this review. The real contact stress p , in equation (A3.1) is the natural quantity to be part of a friction law, but in practice it is the nominal stress, the load divided by the apparent, or nominal, contact area A,, which is set in any given application. In Chapter 2, this stress has been written on. The first point is that, from load force equilibrium, the ratio of onto p , is the same as the ratio of the real to apparent contact area (AJA,): (A3.3a) The second point is that, in Chapter 2, onis normalized with respect to some shear flow stress k of the work or chip material. The dimensionless ratios p)k and slk can be introduced into equation (A3.1) and further p)k eliminated in favour of o,lk by means of equation (A3.3a): (A3.3b)

A single asperity on an elastic foundation

365

In the following sections, a view of how sliding friction depends on material properties, contact geometry and intensity of loading is developed, by concentrating on how p/k and A/An vary in adhesive and deformation friction conditions. A more detailed account of much of the contact mechanics is in the standard text by Johnson (1985). Reference will be made to this work in the abbreviated form (KLJ Ch.x).

As a first step in building up a view of asperity contact, consider the normal loading of a single asperity against a flat counterface. At the lightest loading, the deformation may be elastic. At some heavier load, plastic deformation may set in. The purpose of this section is to establish how transition from an elastic to a plastic state varies with material properties and asperity shape; and what real contact pressuresPr are set up.

A3.2.1 Elastic contact


Figure A3.2 shows asperities idealized as a sphere or cylinder of radius R, or as a blunt cone or wedge of slope /3, pressed on to a flat. The dashed lines show the asperity and flat penetrating each other to a depth 0, as if the other was not there. The solid lines show the deformation required to eliminate the penetration. How Pr varies with the contact width 2a, or with 0; and with R or /3; and with Young's modulus El and E2 and Poisson's ratio VI and V2of the asperity and counterface respectively, is developed here. The contact of an elastic sphere or cylinder on a flat in the absenceof interface shear is the well-known Hertzian contact problem. A dimensional approach gives insight into the contact conditions more simply than does a full Hertzian analysis. In the left-hand part of Figure A3.2, the asperity is shown flattened by a depth 01, and the flat by a depth O2,in accommodating the total overlap and creating a contact width 2a. From the geometry of overlap, supposing 2a to be a fixed fraction of the chordal length 2ac' and when ac R, = 01 + O2=
a2 c

a2
oc

(A3.4)
2R

The surface deformations in the asperity and flat cause sub-surface strains. In the asperity, these are in proportion to the dimensionless ratio 81/a and in the flat to 82/a. When the

Fig. A3.2 Models of elastic asperity deformation

366 Appendix 3
Table A3.1 Elastic contact parameters, from Johnson (1985, Chs 4 and 5)

Spherical Cylindrical Conical Wedge-like

0.42 0.39 0.50 0.50

2.6 2.2 1.6 1 .o

6.2 5.6 3.2 2.0

asperity and flat obey Hooke's law, the mean contact stress p , will increase in proportion to the product of Young's modulus and strain in each: from the asperity's point of view, from the flat's point of view, Combining equations (A3.4) and (A3.5) gives
p , = cE*(a/R)

p,

0~

E,(6,/a)

(A3.5)

Pr Dc E 2 @ 2 / 4

(A3.6)

and the constant of proportionality c requires the full Hertz where 1/E* = (l/E,+l/E2) analysis for its derivation. The full analysis in fact shows that the proper definition of E* involves Poisson's ratio:
--

1-v:
El

1-v;
E2

(A3.7)

E"

and c depends on whether the circular profile of radius R represents a spherically or a cylindrically capped asperity (Table A3.1). Similarly, the pressing together of two spherical or two cylindrical asperities with parallel axes, of radii R , and R,, creates a normal contact stress p,:
p, = cE*(a/R*) where

1/R* = 1/R,

1/R,

(A3.8)

The elastic contact of a wedge or cone on a flat (right-hand part of Figure A3.2(a)) generates a contact pressure pr (KLJ Ch. 5):
p, = cE" tan p

(A3.9)

where c is also given in Table A3.1. The quantities (a/R*) and tan /3 can be regarded as representative contact strains. Their interpretation as mean contact slopes will be returned to later. As they increase, so does p,.

A3.2.2 Fully plastic contact


Figure A3.3 shows a wedge-shaped asperity loaded plastically against a softer (left) and a harder (right) counterface, so that it indents or is flattened. The dependence of p , on asperity slope p and shear flow stress k of the softer material is considered here, by means of slip-line field theory (Appendix 1.2). In each case, the region ADE is a uniform stress region and the free surface condition along AE requires that p, = k. Region ABC is also uniformly stressed. Normal force equilibrium across AC gives

A single asperity on an elastic foundation 367

P,=P~+'
Slip-line EDBC is an a-line, so

(A3.10)

p2=p1 2ky/sk(l+2y/) +

(A3.11)

y The angle t is chosen to conserve the volume of the flow: material displaced from the overlap between the flat and the asperity must re-appear in the shoulders of the flow, but for small values of /3, y / = d 2 . This, with equations (A3.11) and (A3. IO), gives
pr = 2k(l

+ d 2 ) = 5k

(A3.12)

Fig. A3.3 Plastic indenting by, and flattening of, wedge-shaped asperities

A3.2.3 The transition from elastic to plastic contact


The elastic and plastic views of the previous sub-sections are brought together by nondimensionalizing the contact pressures p , by k. In Figure A3.4(a), the elastic and plastic model predictions are the dashed lines. The solid line is the actual behaviour. Departure from elastic behaviour first occurs in the range 1 < p i k < 2.6, at values of (E*/k)(a/R*or tanp) from 2 to 6.2. The values depend on the asperity shape: they are the last two columns in Table A3.1. The fully plastic state is developed for (E*/k)(a/R* or tanp) greater than about 50. p,/k continues to increase at larger deformations than this due to strain hardening.

Fig. A3.4 (a) Real contact pressure variation with asperity deformation severity; (b) the dependence of degree of contact on intensity of loading, in the absence of sliding

368 Appendix 3

A real contact area Ar is associated with a surrounding nominal contact area An. Figure A3.4(a) can be used with equation (A3.3a), to map how A/An increases with O"n!k. The result is shown in Figure A3.4(b) for different values of (E*/k)(a/R* or tan{3).

In the previous section, loading a single asperity was considered. When all the contacts between two surfaces have the same half-width a, Figure A3.4(b) can be used directly to predict the degree of contact from (jn/k. However, on real surfaces, asperities have random heights and are not loaded equally. The effect of this on the use of Figure A3.4(b) is the first point considered in this section. In Figure A3.4(b), predictions are only drawn for A/An < 0.5: the second point considered in this section is wha,t happens at higher degrees of contact, when asperity stress fields start to interact. A3.3.1 Loading of random rough surfaces

Figure A3.5 shows the loading of two rough flat surfaces against a smooth flat counterface. In case (a) all the asperities on the rough surface are identical, imagined as spherical caps of radius R, and are shown in contact with the counterface. In case (b ), the same asperities have been shifted in a random manner normal to the surface, so that the peaks have a random distribution O's heights about their mean height. This situation is the most simple of that can be pictured, to make the point that an increase of load in case (a) causes the load per contact, the half-width a and the stress severity to increase. However, in case (b), the number of contacts can also increase, so that the load per contact and the severity of stress will increase less slowly with load. The situation of Figure A3.5(b) was considered by Greenwood and Williamson (1966), supposing the contact stressesto be elastic, and is reproduced in (KLJ Ch. 13). Provided that the number of asperities in contact is a small fraction of the total available (this means in practice that A/An $ 0.5), the number of contacts grows almost in proportion to the load, so on average the load per contact is almost independent of load. The average real contact pressure jjr is Pr= (0.3 to 0.4)E*V aiR (A3.13a)

and is a function only of E* and the rough surface finish. Compared with equations (A3.8) and (A3.9), v'";;:)R is seen as the measure of mean asperity strain or equivalent slope.

Fig. A3.5 (a) A regular and (b) a random model rough surface loaded on to a flat

Asperities

with traction,

on an elastic foundation

369

I~'

(a)

onlk

(b)

Fig. A3.6 (a)Thedependence A/An on CTJk differentdegrees roughness an elastic of for of on foundation;(b) rigid plasticcompression ridge-shaped of asperities

Indeed, later analyses of rough surface elastic contact have replaced ~ RMS slope of the rough surface. In non-dimensional form: pr k E* A k
q

by A , the q (A3.l3b)

=c-

The severity index (E*/k) Llq. more commonly called the plasticity index \{I, may be used with Figure A3.4(b) to determine the degree of contact of a rough loaded surface. A3.3.2 Loading -~ at high degrees of contact

As AlAn increases above 0.5, ev~n for a randomly rough surface, the availability of new contacts becomes exhausted. A load increase will no longer cause a proportional increase in the number of contacts, but will cause increased deformation of existing contacts. AlAn will no longer increase in direct proportion to O"n/k. Figure A3.6(a) extends Figure A3.4(b) to higher values of O"nlk and AlAn: note the rescaling of O"nlk a log base. At one extreme to ('I' = 2), the displacement of material as the surfaces are brought together is taken up by elastic compression. In this example, full contact is reached at O"nlk= 2. At the other extreme of fully plastic flow ('I' = 50 or more), material displaced from high spots reappears in the valleys between contacts. Figure A3.6(b) represents a model situation of the plastic crushing of an array of wedge-shaped asperities. The material displaced from the crests of the array by the counterface is extruded into the ever-diminishing gap between the contacts. Slip-line field modelling suggeststhat, by the stage that the degree of contact has risen to 0.8, the hydrostatic stress beneath the contacts has risen from z 4k (for well separatedcontacts) to z 9k: then O"nlk 8 (Childs, 1973). z

Section A3.3 considered real contacts' ability to support load in the absence of sliding. When shear stressesdue to sliding are added to the stressesdue to loading, contacts that under load alone are elastic may become plastic; contacts that are already plastic will be overstressed and collapse. These are the sub-topics of this section.

370 Appendix 3

Fig. A3.7 Asperity state of stress dependence on surface shear and plasticity index

A3.4.1 Contact stress regimes under sliding conditions


The stressing of elastic spheres and cylinders loaded against flats, without and with sliding, is reviewed in detail in (KLJ Chs. 4 and 6). Without sliding, the largest shear stress zmaxoccurs from 0 . 4 8 ~ 0 . 7 8 ~ to below the centre of the contact. With sliding, if pa < 0.25 to 0.3, rmax not changed in size by sliding, but the position where it occurs moves is towards the surface. For pa> 0.25 to 0.3, rmax occurs at the surface and its size rises proportionally to pa.The constant of proportionality depends on whether a sphere or a cylinder is being loaded:

rmax (1.27 to 1.5)papr =

(A3.14)

These observations may be applied to the contact of random rough surfaces. Figure A3.7 shows the state of stress (elastic, elastic-plastic or fully plastic) to be expected for different combinations of plasticity index Y and slk. When slk = 0, the transition from elastic to elastic-plastic flow occurs for 4' = 5 to 6; fully plastic flow commences for Y = 50 ' to 60. These values are the same as the transition values of (E*/k)(a/R*)shown in Figure A3.4(a). As slk increases, the elastic boundary is not altered until slk reaches 0.67 to 0.78. For larger values, a purely elastic state does not exist. Thus, in Figure A3.7, the elastic region is capped at these values. (They are derived from equation (A3.14), by noting that pa = slp, and that plastic flow occurs once zmax k.) = How the elastic-plastic/fully plastic boundary is influenced by s/k is not well established theoretically. The boundary drawn in Figure A3.7 is a little speculative.

A3.4.2 Junction growth of plastic contacts


A real area of contact A,, loaded in the absence of sliding, has rmax k within it if it is in = a plastic state. The addition of a sliding force F to the contact, creating an extra shear stress FIA,, will, if A, does not increase, result in an increased rmax. fact, A, grows to prevent In rmax exceeding k . An alternative view of the cause of this junction growth is that, in the absence of sliding, the material surrounding a plastic contact helps to support the load by

Bulk

yielding

371

Fig. A3.8 Thedependence A/An and ./1on aJk and s/k, for hardridgesof slope{3= 5" slidingagainst soft flat of a

imposing a hydrostatic pressure on the deviatoric stress field. Its size, from slip line field modelling, P2 in equation (A3.11) with 2'1' = 1Z", about 4k. The addition of surface shear is on the contact reduces the surrounding's ability to support the load; in other words, the hydrostatic pressure component supporting the load reduces. How the slip-Iine fields of Figure A3.3 become modified by sliding have been studied by Johnson, for the case of a soft asperity on a hard flat (KLJ Ch. 7), and by Oxley (1984) for hard wedges ploughing over a soft flat. The conclusion of both, stemming from the connection of the plastic flow field beneath the contact to the free surface where p = k, is that, for s/k close to I, sliding causesPr to fall from around 5k to lk. For a constant load, this causes a fivefold increase in real contact, at least while asperities are sufficiently far apart not to interact with one another. Figure A3.8 shows Oxley's prediction of how A/An and .u depend on O"n/k, sliding for hard wedge-shaped asperities, of slope fJ = 5", over a soft flat, at different levels of surface shear s/k (the dependenceof A/An on O"n/k the absenceof sliding is shown by the dashed in line). It has been chosen becausethe situation of hard ridges sliding on a soft flat may more realistically represent the condition of a rake face of a cutting tool sliding against a chip than soft asperities sliding on a smooth hard flat. While A/An < 0.5, .u is independent of O"n/k, as A/An approaches I, .u reduces with increasing O"n/k. but Although Figure A3.8 is only one example, it illustrates three general points. (1) The hydrostatic pressure within a sliding contact is predicted by slip-Iine field modelling to be less the larger is s/k, but while it is controlled by the free surface boundary condition, it never becomes less than k. As a result, .u never becomes greater than 1. (2) The reduction in hydrostatic pressure, and hence the junction growth and .u, is very sensitive to s/k when, as in Figure A3.8, s/k is large (close to 1). (3) Once A/An = 1, .u is no longer independent of, but becomes inversely proportional to, load.

When an asperity is supported on an elastic bulk, the only way to accommodate its plastic distortions is by flow to the free surface. It is this, in the previous section, which ensured that Pr never became less than k. When the bulk is plastic, asperity plastic distortion can be

372 Appendix 3

Fig. A3.9 (a) Combined asperity and bulk plastic stressing and (b) derived degrees of contact ACA,, depending on the apparent contact stresses and the bulk field hydrostatic stress level P,lk, for s/k= 1

accommodated by flow into the bulk. Asperity hydrostatic stress andp, then depend on the state of the bulk flow field: it is possible for p, to be less than k. If this happens, the degree of contact becomes greater than considered previously. How it depends on the nominal contact stresses onand z, and on the hydrostatic stress in the bulk field, will now be developed, still for conditions of plane strain and a non-hardening plastic material to which slipline field theory can be applied. Figure A3.9(a) is adapted from Sutcliffe (1988). It shows a combined asperity and bulk slip-line field. The bulk field is described by the hydrostatic pressure pE and slip-lines inclined at Cbulk to the counterface. The asperity fields ADBC, ADBC around the real contacts AC, AC are connected to the bulk by DEC, DEC. It is supposed that p2, beneath the asperities, has been reduced so much by the influence of pE that the stresses on AD, AD are no longer sufficient to cause a plastic state to extend to the free surface: region ADEFEC has become rigid. For this to be the case, at least for high values of real surface shear stress s (slk 3 l), the results of the previous section suggest p/k < 1. How p/k - and hence (with o,/k) A/A, - depends on pE and Cbulk can be determined from the slip-line field and its limits of validity. However, it is simpler to consider the overall force balances between the bulk field, the nominal contact stresses onand z, and the real contact stresses p, and s. Force equilibrium between pa k, Cbulk and onand z creates the relations:
z = k cos 2CbU1,

on= pE+ k sin 2cbu,,


By elimination of

Cbulk,a relation is formed between pEand onand z:

(A3.15a)

Pk E =k 5

-Il -(j2)
%

(A3.15b)

In Figure A3.9(b) the dashed lines show combinations of (dk) and (on/k) consistent with asperities existing on a bulk plastic flow in whichpE/k = 0 or 0.5. The region marked elastic bulk is that for which (dk) and (o,/k) are associated with an elastic bulk unless pE/k < 0. That marked plastic bulk is plastic if pE/k > 0.5.

Friction coefficients greater than unity

373

Equilibrium between the forces of the nominal and real contact stressesgives
l' k Ar An s

(A3.16a)
k

O"n k

Ar

Pr

An

(A3.16b)

According to equation (A3.16a), A/An ~ T/k as s/k ~ I. From equation (A3.16b), if A/An < (O"n/k), p/k > I. However, the slip-Iine field is not valid if p/k > I: flow will break through to the free surface and p/k will be lirDited to I. Thus, for a plastic asperity on a plastic foundation, when s/k is close to I, A/An will equal (r/k) when (T/k) > (O"n/k), and (O"n/k) when (O"n/k) (T/k), up to its maximum possible value of I. Contours of A/An = I, > 0.9,0.8, satisfying this, are added to Figure A3.9(b). Figure A3.9(b) shows, firstly, the levels of dimensionless hydrostatic stress, PE/k, needed for a combination of (T/k) and (O"n/k) be associated with a bulk plastic flow. If to there is bulk plasticity, it then shows how degrees of contact much larger than when the bulk remains elastic (Figure A3.8a) can be generated at values of (O"n/k) I. In these condi< tions the ratio of friction to normal stress (the friction coefficient) becomes greater than I.

In metal machining, and elsewhere, friction coefficients > 1 have been measured in conditions in which asperities have been plastic but (T/k) and (O"n/k) have been too low for bulk plastic flow to be a possibility. What could account for this, that has not been considered in the previous sections? Work hardening offers two possibilities. First, in the same way as it changes the hydrostatic pressure distribution along the primary shear plane in metal machining (Figures 2.11 and 6.9(b, it can modify the pressure within a deforming asperity to reduce the mean value of Pr to a value less than k. However, there is likely only to be a small effect with the rake face asperities in machining, already work hardened by previous deformations. A second possibility imagines a little work hardening and high adhesion conditions, leading to the interface becoming stronger than the body of the asperity. Unstable asperity flow, with contact area growth larger than expected for non-hardening materials, has been observed by Bay and Wanheim (1976). There is a second type of possibility. In the previous sections it has been assumed that an asperity is loaded by an amount W by contact with a counterface and that W does not change as sliding starts. For example, in Section 3.4.2 on junction growth of plastic contacts, it is written that the addition of a sliding force F to a real contact area creates an extra shear stress F/Ar which, if Ar does not increase, will cause Tmax increase. This to assumesthat the stress W/Ar does not decrease. If W is constant, the extra force F causes the two sliding surfaces to come closer to one another: it is this that enables Ar to grow. Green (1955) pointed out that, in a steady state of sliding (between two flat surfaces), the surfaces must be displaced parallel to one another. In that case anyone junction must go through a load cycle. Figure A3.10 is based on Green's work. With increasing tangential displacement, asperities make contact, deform

374 Appendix 3

Tensile Fig. A3.10 Qualitative junction load history for zero normal displacement

and break. The load rises, passesthrough a maximum and falls, but the friction force rises and stays constant until failure. If, at anyone time, there are many contacts in place, each at a random point in its life cycle, an average friction coefficient will be observed that is obtained from the areas under the curves of Figure A3.10, up to the point of failure. Green argued that when conditions were such that junctions failed when the load dropped to zero, the friction coefficient would be unity. Higher coefficients require junctions to be able to withstand tensile forces, as shown. The exact value of the friction coefficient win depend on the exact specification of how the surfaces come together and move apart; and on the junctions' tensile failure laws. Quantitative predictions do not exist.

Bay, N. and Wanheim, T. (1976) Real area of contact and friction stress at high pressure sliding contact. Wear 38, 201-209. Childs, T. H. C. (1973) The persistence of asperities in indentation experiments. Wear 25,3-16. Green, A. P. (1955) Friction between unlubricated metals: a theoretical analysis of the junction model. Proc. Roy. Soc. Lond. A228, 191-204. Greenwood, J. A. and Williamson, J. B. P. (1966) Contact of nominally flat surfaces. Proc. Roy. Soc. Lond. A295, 300-319. Johnson, K. L. (1985) Contact Mechanics. Cambridge: Cambridge University Press. Oxley, P. L. B. (1984) A slip line field analysis of the transition from local asperity contact to full contact in metallic sliding friction. Wear 100,171-193. Sutcliffe, M. P. (1988) Surface asperity deformation in metal forming processes. Int. J. Mech. Sci. 30. 847-868.

Anda mungkin juga menyukai