Sunteți pe pagina 1din 18

5

Fracture of Metals

5.10 In an 'ideal' metal, that is one containing no flaws or defects, fracture will occur when atomic bonding is overcome across an atomic plane which is perpendicular to the tensile force; but in practice metals fail at much lower stresses than the theoretical (4.14). This is due partly to the presence of impurities and other discontinuities in the structure (Fig. 3.16), and partly to the polycrystalline nature of metals which in itself leads to the 'pile-up' of dislocations at or near crystal boundaries. Thus fracture is a common cause of failure in metals and whilst it will always occur when a metal is stressed beyond its tensile strength, under certain conditions it can also occur at stresses even below the elastic limit. Thus, some metals may creep during long periods of time, particularly at high temperatures, until failure occurs; whilst fracture which is the result of repeated cyclic stress is termed fatigue failure. Some metals fail by a form of brittle fracture which is influenced by low temperature conditions. 5.11 The nature of fracture differs from one metal to another. The type of fracture which follows large amounts of plastic deformation is generally referred to as ductile fracture whilst that which occurs after little or no plastic deformation is called brittle fracture (Fig. 5.1).

(tup cone.

FORCE

EXTENSION

FORCE EXTENSION

Fig. 5.1 Types of fracture (i) ductileshowing the well-known cup and cone; (ii) brittle with a strongly 'crystalline' appearance.

It was stated in the previous chapter that during plastic deformation slip takes place along some crystallographic planes more readily than along others. Similarly, metallic fractureor cleavagetends to follow preferred crystallographic planes. Thus FCC metals are likely to suffer cleavage along the {111} planes (3.15), whilst BCC metals cleave along the {100} planes. In CPH metals cleavage occurs along the basal planes of the hexagon (designated the {0001} planes).

Brittle Fracture
5.20 Whilst plastic deformation takes place in a ductile metal by the 'rippling' of dislocations along slip planes (4.16), brittle fracture occurs as a result of complete and sudden separation of atoms as indicated in Fig. 5.2. The theoretically-calculated stress necessary to achieve such separation is considerable, yet in practice, the actual stress required to cause brittle fracture is often relatively small. Moreover, the results obtained for tensile strength measurements on a large number of apparently identical test pieces of a single brittle metal are often very variable (Fig. 5.3(i)) compared with similar determinations made for a ductile metal (Fig. 5.3(ii)). Such results suggest that in brittle materials some factor is present which gives rise to the variability of results. 5.21 In 1920, A. A. Griffith postulated that fractures in brittle solids were propagated from minute flaws in the material. He demonstrated that the strength of freshly-drawn glass fibres often approached the 'theoretical'

Fig. 5.2

Cleavage along crystallographic planes.

TENSILE STRENGTH

NUMBER OF TESTS

TENSILE STRENGTH

NUMBER

OF TESTS

Fig. 5.3 The variability of results during the tensile testing of (i) a brittle material; (ii) a ductile material.

value, but if these fibres were allowed to come into contact with any other substance, including the atmosphere, even for short periods then the strength was considerably reduced. This suggests that the strength of glass fibre was very dependent upon surface perfection and that anything likely to initiate even minute surface irregularities would weaken it. In certain respects these principles may also be applied to brittle metals. Engineers will already be familiar with the concept of 'stress raisers' and stress concentrations associated with the presence of sharp in-cut corners and the necessity of eliminating these in engineering design whenever possible. Thus in iron castings in-cut corners are rounded by using leather 'fillets' on the wooden pattern. 5.22 Griffith's Crack Theory. Whilst the presence of a small fissure (Fig. 5.4) will obviously reduce the effective cross-section of the material, the reduction in breaking stress is very much greater than can be accounted for by this reduction in cross-sectional area. This is because an applied stress, S, generates stress concentrations at the tip of the fissure. Griffith concluded that the concentrated stress, Sc, is related to the applied stress, S, the width of the crack, c, and the radius of curvature of the tip of the crack, r, by:

Fig. 5.4.

If we assume that the tip radius of such a crack is of the order of 10~10 m (roughly one atomic radius) and that such cracks are about 10~4 m in width, then the value SJS (which we can call the factor of stress concentration) is of the order of 103. This indicates that the actual stress, 5C, operating at the tip of the crack is some thousand times greater than the applied stress, S. In this case fracture is taking place at a localised stress nearer to the theoretical value. When a crack progresses through a brittle material under the action of a constant applied stress, S, the concentrated stress, Sc, at the tip increases since an increase in c gives rise to an increase in the term dr. The speed of crack propagation therefore increases and failure is certain. 5.23 So far we have considered only the effects of an applied tensile stress on a brittle material. If however the applied stress is compressive this stress will be transmitted across existing micro-cracks without causing any stress concentrationsthat is, a compressive stress will tend to 'close up' existing fissures. For this reason many brittle metallic materials such as cast iron are relatively weak in tension but strong in compression. Failure in compression will ultimately take place when compressive forces are so high that they produce tensile components of sufficient magnitude along crystallographic planes in the region of a fissure tip. 5.24 lntercrystalline Brittle Fracture In the foregoing paragraphs we have been dealing with brittle fracture in terms of cleavage along transcvystalline planes. Brittle fracture also occurs by the propagation of cracks along grain boundariesthat is, by mtercrystalline fracture. In some cases this type of fracture is due to the presence of grain-boundary films of a hard, brittle second phase. Such films may be formed by segregation during solidification (3.41) so that fracture of this type is more commonly found in cast materials. Brittle films of bismuth segregate at the crystal boundaries of copper in this way so that very small quantities of bismuth (less than 0.01%) may cause excessive brittleness in copper. Impurities present in solid solution may also segregate at grain boundaries during the normal process of coring (8.23). A high concentration of the solute atoms in the grain-boundary region may give rise to brittleness. This is probably due to the pegging of dislocation movements and the consequent initiation of micro-cracks at the grain boundaries. 5.25 Temper Brittleness This occurs in some low-alloy steels when tempered in the range 250-4000C (13.42), and is probably due to the precipitation of films or particles of carbides at grain boundaries. Although tensile strength, and even ductility, are not seriously reduced a very large reduction in impact value is experienced and fracture is intercrystalline.

Ductile Fracture
5.30 As was indicated in Fig. 5.3 the stress at which a ductile metal is likely to fail is much more predictable than that stress in a brittle metal and this fact alone makes ductile metals more suitable for fail-safe engineering design. Moreover, as plastic deformation begins, the tip radius of any

micro-crack present is likely to increase and this will automatically reduce the concentrated stress, 5C, as the value of clr (in the Griffith equation) decreases. As a result of this stress relaxation propagation of the crack may cease. The fracture which ultimately occurs in ductile materials is due partly to strain hardening which progressively reduces the ductility which is necessary for further stress relaxation. 5.31 It is difficult to differentiate implicitly between brittle and ductile fracture since some brittle metals undergo some ductile deformation before fracture, whilst many ductile metals exhibit final brittle-type cleavage. However, whilst a brittle fracture is one in which the progress of the crack involves very little plastic deformation of the neighbouring metal, a ductile fracture is one which proceeds as a result of high localised plastic deformation of the metal near the tip of the crack. Thus, whilst the two extremes of these methods of failure are easily recognised, there can be no sharp division between brittle and ductile fracture. A fracture which is almost completely of the brittle variety will exhibit a mass of minute facets which reflect light strongly and reveal the crystalline nature of the metal. An almost completely ductile fracture, on the other hand, shows a rough, dull dirty-grey surface because much of that surface has been deformed along the general plane of fracture. When viewed under the scanning electron microscope (10.33) however the surface of a ductile fracture has a 'dimpled' appearance, caused by the presence of a multitude of minute cavities accompanying the rupture process. Closer examination reveals that the formation of these minute cavities is due to the presence of very small particles of impurity and other phases. This confirms the observed fact that as the purity of a metal increases so does its ductility. 5.32 Most ductile polycrystalline metals fail with the well-known cupand-cone fracture. Fracture of this type follows the formation of a neck in a tensile test piece. Crack formation begins at the centre of the neck on a plane that is roughly normal to the applied stress axis. Small cavities form near the centre of this cross section (Fig. 5.5(i)) and these coalesce to form a visible fissure (Fig. 5.5(iii)). As deformation proceeds the crack spreads outwards towards the edges of the test piece. The final fracture then occurs rapidly along a surface which makes an angle of approximately 45 with the stress axis. This leaves a circular lip on one half of the test piece and a corresponding bevel on the surface of the other half, leading to the formation of a cup on one half and a cone with a flattened apex on the other half (Fig. 5.1(i)).

Factors Leading to Crack Formation


5.40 Very pure metals are much more ductile than those of slightly lower purity and will often draw down to a 'chisel point' where the cross-section at failure is approaching zero. It is therefore reasonable to suppose that even minute inclusions in the microstructure play an important part in nucleating cracks. When slip in a metal takes place dislocations will tend to pile up at the metal/inclusion interface. Assuming that the inclusion

crack nucleus

Fig. 5.5 The nucleation and development of a crack in a ductile material.

itself is strong and does not shear, a minute fissure will develop at the interface (Fig. 5.6). Of course, the development of the fissure will also depend upon the degree of adhesion between the surface of the inclusion and that of the metal. Thus, if there is little adhesion as, for example, at the interface between copper and cuprous oxide particles, then fissures will form easily, but if adhesion is high then the particles may indeed have a strengthening effect. Thus sintered aluminium powder (SAP), which contains a high proportion of aluminium oxide particles, is stronger than pure aluminium because of the high adhesion at the Al/Al2O3 interface. 5.41 Other barriers to the movement of dislocations can also initiate micro-cracks in a similar manner. Thus Fig. 5.7 suggests how a grain bounSTRESS O"

EMBRYO CRACK

OBSTACLE (SEPARATE PHASE)

STRESS

Fig. 5.6 A pile-up of dislocations at some obstacle leading to the formation of a crack which will be propagated if the stress, a, is increased.

EDGE DISLOCATIONS SLIP PLANE

MICRO-CRACK

Fig. 5.7

The nucleation of a grain-boundary fissure.

dary can act as a barrier to the movement of dislocations so that a pile-up occurs and nucleates a micro-crack. Such a crack continues to grow as further dislocations, possibly generated from the same Frank-Read source, move to join it. Dislocations will not cross grain boundaries since lack of alignment and some disorder will always be present there. Crack initiation can also be caused by movement of dislocations along close-packed or other planes within a crystal (Fig. 5.8(i)). These dislocations then pile up at the intersections of the slip planes and so form a crack (Fig. 5.8(ii)). 5.42 It was shown earlier (5.22) that, arising from Griffith's Crack Theory, at the tip of the crack: / Crack length Stress concentration oc J v -: Crack-tip radius Thus the spread of a crack in a plate glass window is arrested by drilling a hole in front of the tip of the advancing crack. Similarly, catastrophic brittle fracture in welded-steel ships (5.50) was prevented by including dummy rivet holes to arrest the progress of cracks already initiated at structural faults in the steel.

plane plane cleavage plane micro-crack

plane

Fig. 5.8 The initiation of a micro-crack by a running together of dislocations (after A. H. Cottrell).

Ductile-Brittle Transition in Steels


5.50 There have been many instances in the past of failure of metals by unexpected brittleness at low temperatures. That is, metals which suffered normal ductile fracture at ambient temperatures would fail at low temperatures by sudden cleavage fracture and at comparatively low stresses. The failure of the motor sledges during the very early stages of the British South Pole Expedition of 1912-13 may well have been due to fracture of this type and contributed towards the final disaster which overtook the Polar party. In more recent years similar failure was experienced in the welded 'liberty ships' manufactured during the Second World War for carrying supplies from America to Europe and was unexpected and dangerous. Under normal conditions the stress required to cause cleavage is higher than that necessary to cause slip, but if, by some circumstances, slip is suppressed, brittle fracture will occur when the internal tensile stress increases to the value necessary to cause failure. This situation can arise under the action of bi-axial or tri-axial stresses within the material. Such stresses may be residual from some previous treatment, and the presence of points of stress concentration may aggravate the situation. The liberty ships mentioned above were fabricated by welding plates together to form a continuous body. Cracks usually started at sharp corners or arc-weld spots and propagated right round the hull, so that, finally, the ship broke in half. Had the hull been riveted, the crack would have been arrested at the first rivet hole it encountered. Some riveted joints are now in fact incorporated in such structures to act as crack arresters. 5.51 The relationship between mechanical properties and the method of stress application has already been mentioned (2.52) with particular reference to the impact test. Plastic flow depends upon the movement of dislocations and this occurs in some finite time. If the load is applied very rapidly it is possible for stress to increase so quickly that it cannot be relieved by slip. A momentary increase of stress to a value above the yield stress will produce fracture. 5.52 As temperature decreases, the movement of dislocations becomes more difficult and this increases the possibility of internal stress exceeding the yield stress at some instant. Brittle fracture is therefore more common at low temperatures. This is supported by the fact that the liberty ships were in service in the cold North Atlantic. 5.53 Those metals with a FCC structure maintain ductility at low temperatures, whilst some metals with structures other than FCC tend to exhibit brittleness. BCC ferrite is particularly susceptible to brittle fracture, which follows a transcrystalline path along the (100) planes (see 4.12). This occurs at low temperatures as indicated in Fig. 5.9 and the temperature at which brittleness suddenly increases is known as the transition temperature. Other things being equal, as the carbon content of a steel increases so does the transition temperature, making steel more liable to brittle fracture near ambient temperatures. Phosphorus has an even stronger effect in

AUSTENITIC (F.C.C.) STEEL

IMPACT VALUE (j)

FERRITIC (B(ZC.) STEEL

DUCTILE FRACTURE

BRITTLE FRACTURE 8 O % BRITTLE WITH DUCTILE FRACTURE AT RIM

TEMPERATURE ( 0 C ) Fig. 5.9 steels. The relationship between brittle fracture and temperature for ferritic and austenitic

raising the transition temperature in steel and this is one reason why phosphorus is one of the least desirable impurities in ordinary carbon steels. Some elements, notably manganese and nickel, have the reverse effect in that they depress the transition temperature. Thus, for applications involving atmospheric temperatures the transition temperature can be reduced to safe limits by increasing the manganese-carbon ratio of the steel, whilst at the same time controlling the grain size by small additions of aluminium. A suitable steel contains 0.14% carbon and 1.3% manganese. Where lower temperatures are involved it is necessary to use low-nickel steels.

Fatigue
5.60 Engineers have long been aware that either 'live' loads or alternating stresses of relatively small magnitude can cause fracture in a metallic structure which could carry a much greater static or 'dead' load. Under the
action of repeated or fluctuating stresses a metal may become fatigued.

Some of the earliest quantitative research into metal fatigue was carried out in 1861 by Sir William Fairbairn. He found that raising and lowering a 3-tonne mass onto a wrought iron girder some 3 x 106 times would cause the girder to break, yet a static load of 12 tonnes was necessary to cause failure of a similar girder. From further investigations he concluded that there was some load below 3 tonnes which could be raised and lowered an infinite number of times without causing failure. Some ten years later Wohler did further work in this direction and developed the fatigue-testing machine which bears his name. Many engineering metals are subjected to fluctuating stresses during service. Thus the connecting rods in a piston engine are successively pushed

and pulled; contact springs in electrical switch gear are bent to and fro; rotating axles suffer a reversal of stress direction through every 180 of rotation; and aircraft wings are continually bent back and forth as they pass through turbulent air. Often a structure will vibrate in sympathy with some external vibration emanating from equipment such as an air compressor. This too may initiate fatigue failure so that fracture by fatigue is probably the most common cause of engineering failure in metals. 5.61 The yield strength of a material is a measure of the static stress it can withstand without permanent deformation, and is applicable only to components which operate under static loading. Metals subjected to fluctuating or repeated forces fail at lower stresses than do similar metals under the action of 'dead' or steady loads. A typical relationship between the number of cycles of stress (N) and the stress range (S) for a steel is shown in the S-N curve (Fig. 5.10 (H)). This indicates that if the stress in the steel is reduced then it will endure a greater number of stress cycles. The curve eventually becomes almost horizontal indicating that, for the corresponding stress, the member will endure an infinite number of cycles. The stress born under these conditions is called the fatigue limit, 5 D . For many steels the fatigue limit is approximately one half of the tensile strength as measured in a 'static' test. 5.62 Most non-ferrous metals and alloys and also some steels operating under conditions of corrosion, give S-N curves of the type shown in Fig. 5.10(iii). Here there is no fatigue limit as such and a member will fail ultimately if subjected to the appropriate number of stress reversals even at extremely small stresses. With such materials which show no fatigue

cantilever test-piece

No. of Stress Cycles (N)

Stress (S)

Stress (S) fatigue limit(%) Na of Stress Cycles (N) endurance limit i\) No. of Stress Cycles (N)

Fig. 5.10 (i) Represents a test piece suffering alternating stress of range S about a mean value of zero; (ii) a typical S-N curve for steel; (iii) an S-N curve typical of many non-ferrous alloys and some steels, particularly those operating under conditions promoting corrosion.

limit an endurance limit, SN, is used instead. This is the maximum stress which can be sustained for a stated number, N, of cycles of stress. Components made from materials of this type must therefore be designed with some specific life (in terms of stress cycles) in mind and then 'junked' as our American friends put itafter an appropriate working life, that is, before the number of cycles (N) for the corresponding stress (S) has been reached. It should be noted that many authorities now use the terms 'fatigue limit' and 'endurance limit' to mean the same, but the above distinction still seems valid in differentiating between the two classes of S-N curve obtained for different materials. 5.63 Fatigue-testing machines vary in design but many are based on the original Wohler concept (Fig. 5.11). Here the test piece is in the form of a cantilever which rotates in a chuck, the load, W, being applied at the 'free' end. For every rotation of 180 the stress will change from W in one direction to W in the opposite direction. Thus the stress-range will be 2W with a mean value of zero. To determine the fatigue limit (or endurance limit) a number of test pieces are tested in this way, each at a different value of W, until failure occurs, or alternatively, until an infinite number of stress reversals have been enduredsince it is somewhat impracticable to apply an infinite number of reversals, a large number (say 20 x 106) is used instead.

LOADING SYSTEM

CHUCK TEST PIECE REVOLUTION COUNTER BALL RACE

EQUIVALENT CANTILEVER EFFECT

FINAL TEAR (CRYSTALLINE)

STRESs(S)

S/N CURVE

FATIGUE LIMIT NUMBER OF REVERSALS OF STRESS ( N )

FATIGUE CRACK (BURNISHED)

Fig. 5.11 = 2W.

The principles of a simple fatigue-testing machine in which the stress range, S

5.64 The Mechanism of Fatigue Failure Fatigue failure begins quite early in the service life of the member by the formation of a small crack, generally at some point on the external surface. This crack then develops slowly into the material in a direction roughly perpendicular to the main tensile axis. Ultimately the cross-sectional area of the member will have been so reduced that it can no longer withstand the applied load and ordinary tensile fracture will result. A fatigue crack 'front' advances a very small amount during each stress cycle and each increment of advance is shown on the fracture surface as a minute ripple line. These ripple lines radiate out from the origin of fracture as a series of approximately concentric arcs. The individual ripples are far too small to be visible on the fractured surface except by using very high-powered metallographic methods, but under practical conditions a few ripples much larger than the rest, probably corresponding to peak stress conditions, are produced and these are visible on the fractured surface showing the general path which the crack has followed (Fig. 5.12).

burnished crystalline

NUCLEATION

CRACK GROWTH

FRACTURE

Fig. 5.12 The progress of fatigue failure. A fatigue fracture is often easy to identify because the crack-growth region is burnished by the mating surfaces rubbing together as stresses alternate. The ultimate fracture is strongly 'crystalline' in appearance.

A fatigue fracture thus develops in three stagesnucleation, crack growth and final catastrophic failure. Since the crack propagates slowly from the source, the fractured surfaces rub together due to the pulsating nature of the stress and so the surfaces become burnished whilst still exhibiting the conchoidal markings representing the large ripples. Final fracture, when the residual cross section of the member is no longer able to carry the load, is typically crystalline in appearance. Fatigue failures in metals are therefore generally very easy to identify. Fatigue cracks are not the result of brittle fracture but of plastic slip. During cyclic stressing, at stresses above the fatigue limit, plastic deformation is produced continuously and alternately positive and negative. It is this continuous to-and-fro plastic deformation in localised regions which ultimately propagates and spreads a fatigue crack. The continual plastic oscillation of metal layers along slip planes in and out of the surface causes

some of the metal to produce ridges as work hardening sets in to resist 'back slip'. These ridges are forced up at the surface and are termed extrusions (Fig. 5.13). Narrow fissures or intrusions are formed in a similar manner. Several of these intrusions may then interconnect to initiate the start of a fatigue crack.

slip plane

EXTRUSION

INTRUSION slip planes

Fig. 5.13 (i) Stages in the formation of an extrusion by movements of blocks of atoms along slip planes; (ii) the nature of extrusions and intrusions.

5.65 Some Causes of Fatigue Failure Since the fatigue characteristics of metals can be measured accurately, it seems reasonable to suppose that fatigue failure due to a lack of appropriate allowances in design should not occur. This is of course true, yet fatigue failure does continue to happen even in the most sophisticated items of engineering equipment. The catastrophic fatigue failure in Comet airliners some years ago is a case in point. In many such instances a member may be designed to carry a static load (well above 5Dor S N ), yet it may be suffering undetected vibrations which give rise to reversal of stress at a value above SD (or S N ). Such vibrations are often sympathetic. Other design faults include the presence of such stress raisers as a sharpcornered key way in a shaft or an unduly sharp radius in an in-cut corner. Poor workmanship in the shape of surface roughness, scratches or careless toolmarks can also cause stress concentrations. Minute quench cracks in heat-treated steels are also a source of fatigue failure as are the presence of small cavities, inclusions or other discontinuities just below the surface of the material. A surface weakened by decarburisation or other types of deterioration which have led to the softening and weakening of the surface layers also favours the initiation of micro-cracks via the formation of intrusions. Corrosive conditions in the environment can also provide stress-raisers in the form of etch-pits and other surface intrusions such as grain boundary attack. Ordinary surface oxidation may have a similar effect and consequently nucleate a fatigue crack. 5.66 Methods of Improving Fatigue Strength In order to maintain fatigue strength it follows from the above that surface finish should always

be good. Engineer craftsmen of the past were not wasting their time when producing a high surface polish on many of their items of equipment. Since the fatigue strength of a metal is approximately proportional to its tensile strength, any method used to increase the tensile strength of the material will correspondingly improve the fatigue strength. As fatigue failure nearly always commences at the surface, methods of surface hardening will be most effective in limiting the initiation of fatigue cracks. Thus, work-hardening of the surface by shot-peening is beneficial, whilst the carburising and nitriding (19.10) of steels will improve fatigue strength. A case-hardened axle provides a good all-round combination of core toughness, surface wear-resistance and fatigue resistance.

Creep
5.70 So far in this description of plastic deformation and ultimate failure of metals little mention has been made of the part played by time. Nevertheless, metals in service are often required to withstand steady stresses over long periods of time and it has been shown that under these conditions gradual deformation of the metal may occur at all temperatures and stresses. Metals can therefore fail in this way at a stress well below the tensile strength at that particular temperature. This phenomenon of continuous gradual extension under a steady force is known as creep. The effects of creep in most engineering metals are not serious at ambient temperatures though some metals of low melting point (Tm) such as lead exhibit noticeable creep under these conditions. Lead sheeting which may have been protecting a church roof for centuries is sometimes found to be slightly thicker near the eaves than at the ridge. The lead has 'crept' under its own weight over the centuries. Creep becomes very serious with many metals at temperatures above 0.4r m where Tm, as noted previously (4.42) is measured on the Kelvin (or 'absolute') scale. Thus Tm for lead is equivalent to 327 + 273 or 600 K. Hence 0.4 Tm (for lead) is 240 K or -33C. This explains why lead sheeting on a church roof will be likely to creep appreciably at ambient temperatures between, say, 100C and 300C. Similarly the effects of creep must be very seriously considered in the design of gas and steam turbines, steam and chemical plant and furnace equipment. Creep is thus associated in practical terms with both time and temperature and is basically due to slip by the movement of dislocations. 5.71 Fig. 5.14 shows the type of creep curve obtained when a metal is suitably stressed. This curve indicates that, following the initial elastic strain, the plastic strain associated with creep occurs in three stages: (i) Primary, or transient creep, EP, beginning at a fairly rapid rate which then decreases with time because work-hardening sets in. (ii) Secondary, or steady-rate creep, PS, in which the rate of strain is uniform and at its lowest value. (iii) Tertiary creep, SX, in which the rate of strain increases rapidly

" creep

primary

secondary creep

tertiary
creep

STRAIN initial elastic strain TIME Fig. 5.14 A typical creep curve, showing three stages of creep during a long-time, high temperature creep test.

until fracture occurs at X. This stage coincides with necking of the test piece. The form of relationship which exists between stress, temperature and the resultant creep rate is shown in Fig. 5.15. At a low stress and/or a low temperature (curve A) some primary creep occurs but this falls to a negligible value as strain-hardening prevents further slip from taking place. With increased stress and/or temperature (curves B and C) the rate of secondary creep also increases leading to tertiary creep and ultimate failure.

high temperature and/or high stress

STRAIN

low temperature and/or low stress

TIME

Fig. 5.15 Variations of creep rate with stress and temperature. In curve A the creep rate soon becomes negligible as work-hardening sets in. In curve C the creep rate is higher than in curve B because of the use of either higher stress or higher temperature.

5.72 In creep testing, the specimen, in form similar to a tensile test piece, is enclosed in a thermostatically controlled electric tube furnace which can be maintained with accuracy over long periods at any given temperature up to 10000C or more. A simple lever system is often used to load the test piece, and some form of delicate extensometer or strain-gauge system employed to measure the resultant extension at suitable time intervals. The extensometer is sometimes of the optical type, using mirrors, scales and telescopes. Creep values were originally assessed in terms of the limiting creep stress. This was defined as that stress at any given temperature, below which no measurable creep takes place. Obviously this value depends upon the sensitivity of the measuring equipment and in any case it is now known that some creep occurs at all combinations of stress and temperature. Moreover, since formal creep tests involve long periods of time sometimes running into many months, other related tensile values are now used instead of the old concept of 'limiting creep stress'. Most tests involve measuring the stress which will give rise to some predetermined rate of creep during the secondary stage of uniform deformation. The creep rate will therefore be equivalent to the slope of the curve during the secondary stage of creep. The Hatfield 'time-yield value' is derived from such a short-term creep test. It determines the stress at a given temperature which will produce a strain of 0.5% of the gauge length in the first twenty-four hours and a further strain of not more than one part per million of the gauge length in the next forty-eight hours. Creep tests of this type do not attempt to derive a 'creep limit' but are practical values upon which engineering design can be based. 5.73 The Mechanism of Creep Creep is a deformation process in which three main features appear to be involved: (i) the normal movement of dislocations along slip planes; (ii) a process known as 'dislocation climb' which is responsible for rapid creep at temperatures above 0.5 Tm\ (iii) slipping at grain boundaries. In the primary stages of creep, dislocations move quickly at first but soon become piled-up at various barriers. Nevertheless, thermal activation enables them to surmount some barriers, though at a decreasing rate so that the creep rate is reduced. At temperatures in excess of 0.5 Tm, thermal activation is sufficient to promote a process known as 'dislocation climb' (Fig. 5.16). This would bring into use new slip planes and so reduce the rate of work hardening. Hence creep is a process in which work hardening is balanced by thermal softening which allows slip to continue. At low temperatures recovery does not take place due to lack of thermal activation (Curve AFig. 5.15) and so unrelieved work-hardening leads to a reduction in the creep rate almost to zero. In addition to plastic deformation by dislocation movement, deformation by a form of slip at the grain boundaries also occurs during the secondary stage of creep. These movements possibly lead to the formation of 'vacant sites', that is lattice positions from which atoms are missing, and this in turn makes possible 'dislocation climb' (Fig. 5.16). The relationship

(a)

vacant site

(b)

Fig. 5.16 Two examples illustrating 'dislocation climb'. In each case this is made possible by the presence nearby of a 'vacant site' (or 'vacancy'), ie a lattice position not occupied by an atom. In both (a) and (b) the dislocation has 'climbed' on to a new slip plane.

between grain boundaries and creep is indicated by the fact that at high temperatures fine-grained metals creep more than coarse-grained metals of the same compositions, presumably because the fine-grained metals contain a higher proportion of grain-boundary region per unit volume of metal. At low temperatures, where the grain-boundary material is more 'viscous', fine-grained metals are more creep-resistant and generally tougher. In the tertiary stage of creep micro-cracks are initiated at grain boundaries due largely to the movement of dislocations, but in some cases to the migration of vacant sites there. Necking and consequent rapid failure follow. 5.74 Creep Resistance This can be increased by impeding the movement of dislocations in a metal and also by inhibiting the formation of new ones. Thus the presence of solute atoms (8.21) which do not diffuse rapidly will 'pin down' dislocations effectively; whilst the presence of small dispersed particles of a hard, strong constituent will have a 'particlehardening' effect (8.62) by acting as barriers to the movement of a dislocation front. The high-temperature 'Nimonic' series of alloys (18.32) rely on this type of strengthening mechanism.

Exercises
1. Distinguish, with the aid of sketches, between ductile and brittle fracture. (5.20 and 5.30) 2. Show how Griffith's Crack Theory explains the considerable effects which the presence of microstructural faults have as stress raisers in a metal in tension. Why is cast iron weak in tension but strong in compression? (5.22) 3. What are the principal causes of intercrystalline brittle fracture? (5.24)

4. What are the main factors leading to crack formation in metals under the influence of mechanical stress? (5.40) 5. Discuss the relationship between impact value and service temperature for a 0.2% carbon steel. How may the properties of such a steel be improved for service at sub-zero temperatures? (5.50) 6. Write a short account of the importance of a consideration of fatigue in engineering design. (5.60) 7. Define the term fatigue limit and show how the character of this value varies as between ferrous and non-ferrous materials. Describe a method used to determine the fatigue limit of a given steel. (5.61 and 5.63) 8. Why is the concept of limiting creep stress no longer used as a criterion of creep phenomena? What other values are used to quantify creep and how are they derived? (5.72) 9. Creep is a high-temperature phenomenon. Show by what stages and mechanisms it proceeds. (5.71 and 5.73)

Bibliography
American Society for Metals, Case Histories in Failure Analysis. Biggs, W. D., The Brittle Fracture of Steel, Macdonald and Evans, 1960. Chell, G. G., Developments in Fracture Mechanics, Applied Science, 1979. Colangelo, V. J. and Heiser, F. A., Analysis of Metallurgical Failures, John Wiley, 1974. Duggan, T. V. and Byrne, J., Fatigue as a Design Criterion, Macmillan, 1977. Fuchs, H. O. and Stephens, R. I., Metal Fatigue in Engineering, John Wiley, 1980. Kennedy, A. J., Processes of Fatigue and Creep in Metals, Oliver and Boyd. Osborne, C. J., Fracture, Butterworths, 1979. Smallman, R. E., Modern Physical Metallurgy, Butterworths, 1985. Sully, A. H., Creep, Butterworths. BS 3518: 1984 Methods of Fatigue Testing. BS 3500: 1987 Methods for Creep and Rupture Testing of Metals.

S-ar putea să vă placă și