Vous êtes sur la page 1sur 33

TMM4195

Dimensjonering mot utmatting Fatigue Design

Introduction to Fatigue

NTNU 2007
Per Jahn Haagensen

Norges teknisk-naturvitenskapelige universitet Institutt for produktutvikling og materialer

CONTENTS
1 INTRODUCTION ............................................................................................................. 1 2 CHARACTERISTICS OF FATIGUE FRACTURE SURFACES ............................ 1 3 NATURE OF THE FATIGUE PROCESS ................................................................... 2 4 FATIGUE LOADING....................................................................................................... 4 5 FATIGUE LIFE DATA.................................................................................................... 5 5.1 Fatigue endurance curves .................................................................................................... 5 5.2 Fatigue testing ...................................................................................................................... 6 5.3 Fatigue test data ................................................................................................................... 7 6 MAIN FACTORS THAT INFLUENCE FATIGUE LIFE ......................................... 7 6.1 Material effects .................................................................................................................... 8 6.2 Mean stress effects............................................................................................................... 9 6.3 Notch effects ...................................................................................................................... 10 6.4 Size effects ......................................................................................................................... 12 6.5 Effects of surface finish ..................................................................................................... 13 6.6 Residual stress effects........................................................................................................ 14 6.7 Effects of Corrosion........................................................................................................... 15 7 CLOSURE ........................................................................................................................ 16 8 REFERENCES ..................................................................................................................... 16

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 0

NTNU 2006 1 INTRODUCTION Fatigue is generally referred to as a process in which damage is accumulated in a material undergoing fluctuating loading, eventually resulting in failure even if the maximum load is well below the elastic limit of the material. Fatigue is a process of local strength reduction that occurs in engineering materials such as metallic alloys, polymers and composites, e.g. concrete and fiber reinforced plastics. Although the phenomenological details of the process may differ from one material to another, the following definition given by ASTM [1] encompasses fatigue failures in all materials:
Fatigue - the process of progressive localized permanent structural change occurring in a material subjected to conditions that produce fluctuating stresses and strains at some point or points and that may culminate in cracks or complete fracture after a sufficient number of fluctuations.

The important features of the process relevant to fatigue in metallic materials are indicated by the underlined words in the definition above. Fatigue is a progressive process in which the damage develops slowly in the early stages and accelerates very quickly towards the end. A simple measure of fatigue damage can be crack length, but this is easily measurable only at a relatively late stage in the life of a component. Thus the first stage consists of a crack initiation phase, which for smooth and mildly notched parts that are subjected to small loads cycles may occupy more than 95percent of the life. In most case cases the initiation process is confined to a small area, usually of high local stress, where the damage accumulates during stressing. In adjacent parts of the components, with only slightly lower stresses, no fatigue damage may occur and these parts thus have an infinite fatigue life. The initiation process usually results in a number of micro-cracks that may grow more or less independently until one crack become dominant through a coalescence process as the microcracks start to interact. Under steady fatigue loading this crack grows slowly, but starts to accelerate when the reduction of the cross section increases the local stress field near the crack front. Final failure occur as an unstable fracture when the remaining area is too small to support the load. These stages in the fatigue process can in many cases be related to distinctive features of the fracture surface of components that have failed under fluctuating loads, the presence of these features can therefore be used to identify fatigue as the probable cause of failure.

2 CHARACTERISTICS OF FATIGUE FRACTURE SURFACES Typical fracture surfaces in mechanical components that were subjected to fatigue loads are shown in Fig. 1. A characteristic feature of the surface morphology which is evident in both macrographs is the flat, smooth region of the surface exhibiting beach marks (also called clamshell marks). This part represents the portion of the fracture surface over which the crack grew in a stable, slow mode. The rougher regions, showing evidence of large plastic deformation, is the final fracture area through which the crack progressed in an unstable mode. The beach marks may form concentric rings that point toward the areas of initiation. The origin of the fatigue crack may be more or less distinct. In some cases a defect may be identified as the origin of the crack, in other cases there is no apparent reason why the crack should start at a particular point in a fracture surface. If the critical section is at a high stress concentration fatigue initiation may occur at many points, in contrast to the case of unnotched parts where the crack usually grows from one point only, see Fig. 2.

P J Haagensen

Introduction to fatigue

Page 1

NTNU 2006 While the presence of any defects at the origin may indicate the cause of the fatigue failure, the crack propagation area may yield some information regarding the magnitude of the fatigue loads and also about the variations in the loading pattern. Firstly, the relative magnitude of the areas of slow-growth and final fracture regions give an indication of the maximum stresses and the fracture toughness of the material. Thus, a large final fracture area for a given material indicates a high maximum load, whereas a small area indicates that the load was lower at fracture. Similarly, for a fixed maximum stress, the relative area corresponding to slow crack growth increases with the fracture toughness of the material (or with the tensile strength if the final fracture is a fully ductile overload fracture). Beach marks are formed when the crack grows intermittently and at different rates during random variations in the loading pattern under the influence of a changing corrosive environment. Beach marks are therefore not observed in the surfaces of fatigue specimens tested under constant amplitude loading conditions without any start-stop periods. The average crack growth is of the order of a few millimeters per million cycles in high cycle fatigue, and it is clear that the distance between bands in the beach marks are not a measure of the rate of crack advance per load cycle. However, examination by electron microscope at magnifications between 1000 and 30 000 may reveal characteristic surface ripples called fatigue striations, see Fig. 3. Although somewhat similar in appearance, these lines are not the beach marks described above as one beach mark may contain thousands of striations. During constant amplitude fatigue loading at relatively high growth rates in ductile material such as stainless steels and aluminum alloys the striation spacing represents the crack advancement per load cycle. However, in low stress, high cycle fatigue where the striation spacing is less than one atomic spacing (~ 2.510-8 m) per cycle. Under these conditions the crack does not advance simultaneously along the crack front, growth occurring instead only along some portions during a few cycles, then arrests while growth occurs along other segments. Striations as shown in Fig. 3 are not seen if the crack grows by other mechanisms such as microvoid coalescence or, in brittle materials, microclevage. In structural steels the crack can propagate by all three mechanism, and striations may be difficult to observe. Fig. 4 shows an example of beach marks and striations in the fracture originating at a large defect in a welded C-Mn steel with a yield strength of about 360 MPa.

3 NATURE OF THE FATIGUE PROCESS From the description of the characteristics of fatigue fracture surfaces, three stages in the fatigue process may be identified: Stage I: Crack initiation Stage II: Propagation of one dominant crack Stage III: Final fracture Fatigue cracking in metals is always associated with the accumulation of irreversible plastic strain. The crack initiation process which is discussed in the following applies to smooth specimens made of ductile materials. In high cycle fatigue the maximum stress in cyclic loads that eventually cause fatigue failure may be well below the elastic limit of the material, and large scale plastic deformation does not occur. However, at a free surface plastic strains may accumulate as a result of dislocation movements. Dislocations are line defects in the lattice structure which

P J Haagensen

Introduction to fatigue

Page 2

NTNU 2006 can move and multiply under the action of shear stresses, leaving a permanent deformation. Dislocation mobility and hence the amount of deformation (or slip) is greater at a free surface than in the interior of crystalline materials due to lack of constraint from grain boundaries. Grains in polycrystalline structural metals are individually oriented in a random manner. Each grain, however, has an ordered atomic structure giving rise to directional properties. Deformation for example, takes place on crystallographic planes of easy slip along which dislocations can move more easily than other planes. Since slip is controlled primarily by shear stress, slip deformation takes place along crystallographic planes that are oriented close to 45 to the tensile stress direction. The results of such deformation is atomic planes sliding relative to each other, resulting in a roughening of the surface in slip bands. During further cycling slip band deformation is intensified at the surface and extending into the interior of the grain, resulting in so-called persistent slip bands, (PSB's). The name originated from the observation in early studies of fatigue that slip band would reappear -"persist"- at the same location after a thin surface layer was removed by electropolishing. The accumulation of local plastic flow result in surface ridges and troughs called extrusions and intrusions, respectively, Fig. 5. The cohesion between the layers in slip band is weakened by oxidation of fresh surfaces and hardening of the strained material. At some point in this process small cracks develop in the intrusions. These microcracks grow along slip planes, i.e. a shear stress driven process. Growth in the shear mode, called stage I crack growth extends over a few grains. During continued cycling the microcracks in different grains coalesce resulting in one or a few dominating cracks. The stress field associated with the dominating crack cause further growth under the primary action of maximum principal stress; this is called stage II growth. The crack path is now essentially perpendicular to the tensile stress axis. Crack advancement is, however, still influenced by the crystallographic orientation of the grains and the crack grows in a zigzag path along slip planes and cleavage planes from grain to grain, see Fig. 6. Most fatigue cracks advance across grain boundaries as indicated in Fig. 6, i.e. in a transcrystalline mode. However, at high temperatures or in a corrosive environment, grain boundaries may become weaker than the grain matrix, resulting in intercrystalline crack growth. The fracture surface created by stage II crack growth are in ductile metals characterized by striations whose density and width can be related to the applied stress level. Since crack nucleation is related to the magnitude of stress, any stress concentration in the form of external or internal surface flaws can marked reduce fatigue life, in particular when the initiation phase occupies a significant portion of the total life. Thus a part with a smooth, polished surface generally has a higher fatigue strength than one with a rough surface. Crack initiation can also be facilitated by inclusions, which act as internal stress raisers. In ductile materials slip band deformations at inclusions are higher than elsewhere and fatigue cracks may initiate here unless other stress raisers dominate. In high strength materials, notably steels and aluminum alloys, a different initiation mechanism is often observed. In such materials, which are highly resistant to slip deformation, the interface between the matrix and inclusion may be relatively weak, and cracks will start here if decohesion occurs at the inclusion surface, aided by the increased stress/strain field around the inclusion. Fig. 7 shows small fatigue cracks originating at inclusions in a high strength steel. Alternatively, a hard brittle inclusion may break and a fatigue crack may initiate at the edges of the cleavage fracture. From the discussion above it is evidently not possible to make a clear distinction between crack nucleation and stage I growth. "Crack initiation" is thus a rather imprecise

P J Haagensen

Introduction to fatigue

Page 3

NTNU 2006 term used to describe a series of events leading to a stage II crack. Although the initiation stage includes some crack growth, the small scale of the crack compared with microstructural dimensions such as grain size invalidates a fracture mechanics based analysis of this growth phase. Instead, local stresses and strains are commonly related to material constants in prediction models used to estimate the length of stage I. The material constants are normally obtained from tests on smooth specimens subjected to stress or strain controlled cycling.

4.

FATIGUE LOADING

The simplest form of stress spectrum to which a structural element may be subjected is a sinusoidal or constant amplitude stress-time history with a constant mean load, as illustrated in Fig. 8. Since this is a loading pattern which is easily defined and simple to reproduce in the laboratory it forms the basis for most fatigue tests. The following six parameters are used to define a constant amplitude stress cycle: Smax Smin Sa S= Sm = R Sm = = maximum stress in the cycle = minimum stress in the cycle = stress amplitude = 1 ( S max - S min ) 2 stress range = S max - S min = 2 S a mean stress in the cycle = 1 ( S max + S min ) 2 = stress ratio = S min / S max S max (1+ R) / 2 = S a (1+ R) / (1 - R)

The stress cycle is uniquely defined by any two of these quantities, except combinations of stress range and stress amplitude. Various stress patterns are shown in Fig. 9, with definitions in accordance with ISO [2] terminology. The stress range is the primary parameter influencing fatigue life, with mean stress as a secondary parameter. The stress ratio is often used as an indication of the influence of mean loads, but the effect of a constant mean load is not the same as for a constant mean stress. The difference between S-N curves with constant mean stress or constant R-ratio is discussed in the section on fatigue testing. The test frequency is needed to define a stress history, but in the fatigue of metallic materials the frequency is not an important parameter, except at high temperatures when creep interacts with fatigue, or when corrosion influences fatigue life. In both cases a lower test frequency results in a shorter life. Typical stress-time histories obtained from real structures are one shown in Fig. 10. The sequence in Fig. 10a has a constant mean stress, individual stress cycles are easily identifiable, and it is necessary to evaluate this stress history in terms of stress range only. The more "random" stress variations in Fig. 10b is called a broad band process because the power density function (a plot of energy vs. frequency) spans a wide frequency range, in contrast to the one in Fig. 10a which contains essentially one frequency. The difference is illustrated in Fig. 11. The load history in Fig. 11 can be interpreted as a variation of the main load with superimposed smaller excursions that could be caused by e.g. second order vibrations or by electronic noise in the load acquisition system. In case of true mean load variations not only the range but also the mean of each load cycle needs to be recorded in order to estimate the influence of mean load on the damage accumulation. In both cases it is necessary to eliminate the smaller cycles since they may be below the fatigue limit and

P J Haagensen

Introduction to fatigue

Page 4

NTNU 2006 therefore cause no fatigue damage, or because they do not represent real load cycles. Thus a more complicated evaluation procedure is required for identifying and counting individual major stress cycles and their associated mean stresses. Counting methods such as the range pair, rainflow and the reservoir methods are designed to achieve this.

FATIGUE LIFE DATA

The total fatigue life in terms of cycles to failure can be expressed as (1) Nt = Ni + N p where Ni and Np are number of cycles spent in the initiation and propagation stages, respectively. As noted, the two stages are distinctly different in nature and different material parameters control their length. The life of unnotched components, for example, is dominated by crack initiation. In sharply notched parts, however, or in parts containing crack-like defects, e.g. welded joints, the crack growth stage dominates and crack propagation data may be used in an assessment of fatigue life using fracture mechanics analysis. Therefore different test methods are necessary to assess the fatigue properties of these types of components.

5.1

Fatigue endurance curves

Fatigue data for components whose lives consist of an initiation phase followed by crack propagation are usually presented in the form of S-N curves, where applied stress S is plotted against total cycles to failure, N (= Nt). As the stress decreases, the life in cycles to failure increases, as illustrated in Fig. 12. The S-N curves for ferrous and titanium alloys exhibit a limiting stress below which failure does not occur; this is called the fatigue or endurance limit. The branch point or "knee" of the curve lies normally in the 105 to 107 cycle range. In aluminum and other nonferrous alloys there is no stress asymptote and a finite fatigue life exists at any stress level. All materials, however, exhibit a relatively flat curve in the high-cycle region, i.e. at lives longer than about 105 cycles. A characteristic feature of fatigue tests is the large scatter in endurance data, this is particularly evident when a number of specimens are tested at the same stress level, as illustrated in Fig. 13. Plotting the data for a given stress level along a logarithmic endurance axis gives a distribution which can be approximated by the Gaussian (or normal) distribution, hence endurance data are said to have a log normal distribution. Alternatively the Weibull distribution may be used, but the choice is not important since about 200 specimens, tested at the same stress level, are required to make a statistically significant distinction between the two distributions. This number is about one order of magnitude larger than the quantity of specimens that typically is available for fatigue testing at one stress level. Assuming the life distribution to be log normal, the associated mean life curve and the standard deviation can be used define a design S-N curve for any desired probability of failure. When the crack propagation stage dominates fatigue life, design data may be obtained from crack growth curves, an example of which is shown schematically in Fig. 14. The stress intensity factor K uniquely describes the stress field near the crack tip, and is therefore used in the design against unstable fracture. Likewise, the range of the stress intensity factor, K, may be expected to govern fatigue crack growth. The validity of this

P J Haagensen

Introduction to fatigue

Page 5

NTNU 2006 assumption was first proved by Paris [3], and later verified by many other researchers. The crack growth curve, which has a sigmoidal shape, spans three regions as indicated schematically in Fig. 14. In Region I the crack growth rate drops off asymptotically as K is reduced towards a limit or threshold, Kth, below which no crack growth takes place. Like fatigue endurance data, crack growth data exhibit considerable scatter and test results must be evaluated by statistical methods in order to derive useful design data.

5.2 Fatigue testing The basis for any design methodology aimed at preventing fatigue failures is data characterizing the fatigue strength of components and structures. Fatigue testing is therefore essential for the fatigue design process. The ideal fatigue test may be defined as a test in which an actual structure is subjected to the service load spectrum of that structure. However, life estimates are required before the design is finalized or details of the loading history are known. Additionally, each structure will experience a particular load history that is unique for that structure, so many simplifications and assumptions need to be made regarding the test stress sequence which is going to represent the many types of service histories that can occur in practice. Fatigue testing is therefore performed in several ways, depending of the stage the design or production of the structure has reached or the intended use of the data. The following five main types of tests can be identified: 1. 2. 3. 4. 5. Stress-life testing of small specimens Strain-life testing of small specimens Crack growth testing S-N tests of components Prototype testing for design validation

The first three tests are idealized tests that produce information on the material response. The use of the results from these tests in life prediction of components and structures requires additional knowledge of influencing factors related to the geometry, size, surface condition and corrosive environment. S-N tests of components are also normally standardized tests that make life predictions more accurate compared with the three other tests because the uncertainties regarding the influence of notches and surface conditions are reduced. Service loading or variable amplitude testing normally requires a knowledge of the response of the actual structure to the loading environment, and is therefore normally used only for prototype or component testing at a late stage in the production process. In accordance with the many different needs for fatigue testing a large variety of laboratory testing machines have been developed. A classification of commonly used machines might be as follows: 1. Rotating bending machines a. Constant bending moment b. Cantilever bending Reciprocating bending machines Axial stress machines a. Resonant type, mechanical (unbalanced mass) or electromagnetic b. Direct force type

2. 3.

P J Haagensen

Introduction to fatigue

Page 6

NTNU 2006 4. 5. 6. 7. 8. Torsion machines Multiaxial machines Computer controlled electro-hydraulic machines Special component testing machines Full scale or prototype testing systems

Rotating bending machines were used in the past to generate large amounts of test data in a relatively inexpensive way. Two types are shown schematically in Fig. 15. The computer-controlled closed loop testing machines are widely used in all modern fatigue testing laboratories. Most are equipped with hydraulic grips that facilitate insertion and removal of specimens. A schematic diagram of such a testing machine is shown in Fig. 16. These machines are capable of a precise control of almost any type of stress-time, strain-time or load-time pattern and are therefore replacing other types of testing machines. 5.3 Fatigue test data

Among the first systematic fatigue investigations reported in the literature are those set up and conducted by the German railway engineer, August Whler, between 1852 and 1870. He performed tests on full scale railway axles and also small scale bending, axial and torsion tests on several types of materials. Typical examples of Whler's original data are shown in Fig. 17. These data are presented in what is now known as Whler or S-N diagrams. Such diagrams are still commonly used in the presentation of fatigue data, although the stress axis is often on a logarithmic scale in contrast to Whler's linear stress axis. Basquin's equation is often fitted to test data: (2) S a N b = constant where Sa is the stress amplitude, and b is the slope. When both axes have logarithmic scales, Basquin's equation becomes a straight line. Other types of diagrams are used, for instance to demonstrate the influence of mean stress; examples are the Smith or Haigh diagrams which are shown in Fig. 18. Low cycle fatigue data are almost universally plotted in strain vs. life diagrams since strain is a more meaningful and more easily measurable parameter than stress when the stress exceeds the elastic limit.

6 MAIN FACTORS THAT INFLUENCE FATIGUE LIFE The difference in fatigue behavior of full scale machine or structural components as compared with small laboratory specimens of the same material is sometimes striking. In the majority of cases the real life component exhibits a considerably poorer fatigue performance than the laboratory specimen although the computed stresses are the same. This difference in fatigue response can be examined in a systematic manner by evaluating the various factors that influence fatigue strength. Qualitative and quantitative assessments of these effects are presented in the following paragraphs.

P J Haagensen

Introduction to fatigue

Page 7

NTNU 2006 6.1 Material effects

Effect of static strength on basic S-N data For small unnotched, polished specimens tested in rotating bending or fully reversed axial loading there is a strong correlation between the high-cycle fatigue strength at 106 to 107 cycles (or fatigue limit) So, and the ultimate tensile strength Su. For many steel materials the fatigue limit (amplitude) is approximately 50% of the tensile strength, i.e. So = 0.5 Su. The ratio of the alternating fatigue strength So to the ultimate tensile strength Su is called the fatigue ratio. The relationship between the fatigue limit and the ultimate tensile strength is shown in Fig. 19 for carbon and alloy steels. The majority of data are grouped between the lines corresponding to fatigue ratios of 0.6 and 0.35. This holds true also for aluminium alloys. Another feature is that the fatigue strength does not increase significantly for Su > 1400 MPa . Other relationships between fatigue strength and static strength properties based on statistical analysis of test data are listed in Table 1. Table 1. Relationship between fatigue limit at zero mean stress So and tensile test properties for unnotched, small polished steel specimens Authors Heywood Heywood Stribeck Houdremont and Mailnder Jnger Relationship (stress in MPa) So = 0.5 Su So = 0.43 Sy' + 104 So = (0.285 20%)(Sy + Su) So = 0.25(Sy + Su) + 50 So = 0.20(Sy + Su + Z)

In the table Su' is the 0.2% proof stress and Z is the elongation at failure. For real life components, the effects of notches, surface roughness and corrosion reduce the fatigue strength, the effects being strongest for the higher strength materials. The variation in fatigue strength with the tensile strength is illustrated in Fig. 20. The data in Fig. 20 are consistent with the fact that cracks are quickly initiated in components that are sharply notched or subjected to severe corrosion. The fatigue life then consists almost entirely of crack growth. Crack growth is very little influenced by the static strength of the material, as illustrated in Fig. 21, and the fatigue lives of sharply notched parts are therefore almost independent of the tensile strength. An important example is welded joints which always contain small crack-like defects from which crack start growing after a very short initiation period. Consequently the fatigue design stresses in current design rules for welded joints are independent of the ultimate tensile strength. Crack growth data Fatigue crack growth rates seem to be much less dependent on static strength properties than crack initiation, at least within a given alloy system. In a comparison of crack growth data for many different types of steel, with yield strengths from 250 to about 2000 MPa levels of steel, Barsom [4] found that grouping the steels according to microstructure would minimize scatter. His data for ferritic-pearlitic, martensitic and austenitic are shown in Fig. 22. Also shown in the same diagram is a common scatter band which indicates a relatively small difference in crack growth behavior between the three classes of steel. While data for aluminum alloys show a larger scatter than for steels, it is still possible to define a common

P J Haagensen

Introduction to fatigue

Page 8

NTNU 2006 scatter band. Recognizing that different alloy systems seem to have their characteristic crack growth curves, attempts have been made to correlate crack growth data on the basis of the following expression m da K = C (3) E dN An implication of Eq. 3 is that at equal crack grow rates, a crack in a steel plate can sustain three times higher stress than the same crack in an aluminum plate. Thus, a rough assessment of the fatigue strength of an aluminum component whose life is dominated by crack growth can be obtained by dividing the fatigue strength of a similarly shaped steel component by three.

6.2

Mean stress effects

In 1870 Whler identified the stress amplitude as the primary loading variable in fatigue testing; however, the static or mean stress also affects fatigue life as shown schematically in Fig. 22. In general, a tensile mean stress reduces fatigue life while a compressive mean stress increases life. Mean stress effects are presented either by the mean stress itself as a parameter or the stress ratio, R. Although the two are interrelated through
Sm = Sa 1+ R 1- R

(4)

the effects on life are not the same, i.e. testing with a constant value of R does not have the same effect on life as a constant value of Sm, the difference is shown schematically in Fig. 24. As indicated in Fig. 23a), testing at a constant R value means that the mean stress decreases when the stress range is reduced, therefore testing at R = constant gives a better S-N curve than the Sm = constant curve, as indicated in Fig. 23b). It should also be noted that when the same data set is plotted in an S-N diagram with R = constant or with Sm = constant, the two S-N curves appear to be different, as shown in Fig. 25. The effect of mean stress on the fatigue strength is commonly presented in Haigh diagrams as shown in Fig. 26, where Sa/So is plotted against Sm/Su. So is the fatigue strength at a given life under fully reversed (Sm = 0, R = -1) conditions. Su is the ultimate tensile strength. The data points thus represent combinations of Sa and Sm giving that life. The results were obtained for small unnotched specimens, tested at various tensile mean stresses. The straight lines are the modified Goodman and the Soderberg lines, and the curved line is the Gerber parabola. These are empirical relationships that are represented by the following equations: Modified Goodman Sa Sm + = 1 So Su
Sa S m + Su So
2

(5)

Gerber

= 1

(6)

P J Haagensen

Introduction to fatigue

Page 9

NTNU 2006

Soderberg

Sa Sm + = 1 So Sy

(7)

The Gerber curve gives a reasonably good fit to the data, but some points fall below the line, i.e. on the unsafe side. The Goodman line represents a lower limit of the data, while the Soderberg line is a relatively conservative lower bound that is sometimes used in design. Another, less conservative approach to design than the Soderburg line is indicated by the dashed line in Fig. 26. which represents the yield line, i.e. yielding takes place in the region to the right of this line. The Gerber curve is used at the left of the intersection point while the dashed line is used at higher mean stresses to avoid the nominal stress exceeding the yield limit. The three expressions should be used with care in design of actual components since the effects of notches, surface condition, size and environment are not accounted for. Also stress interaction effect due to mean load variation during spectrum loading might modify the mean stress effects given in the three equations.

6.3

Notch effects

Fatigue is a weakest link process which depends on the local stress in a small area. While the higher strain at a notch makes no significant contribution to the overall deformation, cracks may start growing here and eventually result in fracture of the part. It is therefore necessary to calculate the local stress and relate this to the fatigue behavior of the notched component. A first approximation is to use the S-N curve for unnotched specimens and reduce the stress by the Kt factor. An example of this approach is shown in Fig. 27 for a sharply notched steel specimen. The predicted curve fits reasonably well in the high cycle region, but at shorter lives the calculated curve is far too conservative. The tendency shown in Fig. 27 is in fact a general one, namely that the actual strength reduction in fatigue is less than that predicted by the stress concentration factor. Instead the fatigue notch factor Kf is used to evaluate the effect of notches in fatigue. Kf is defined as the unnotched to notched fatigue strength, obtained in fatigue tests:
Kf = fatigue strength of unnotched specimen fatigue strength of notched specimen

(8)

From Fig. 27 it is evident that Kf varies with fatigue life, however, Kf is commonly defined as the ratio between the fatigue limits. With this definition Kf is less than Kt , the stress increase due to the notch is therefore not fully effective in fatigue. The difference between Kf and Kt arise from several sources. Firstly, the material in the notch may be subject to cyclic softening during fatigue loading and the local stress is reduced. Secondly, the material in the small region at the bottom of the notch experiences a support effect caused by the constraint from the surrounding material so that the average strain in the critical region is less than that indicated by the elastic stress concentration factor. Finally, there is a statistical variability effect arising from the fact that the highly stressed region at the notch root is small, so there is a smaller probability of finding a weak spot. The notch sensitivity q is a measure of how the material in the notch responds to fatigue cycling, i.e. how Kf is related to Kt. q is defined as the ratio of effective stress increase in fatigue due to the notch, to the theoretical stress increase given by the elastic stress concentration factor. Thus, with reference to Fig. 28

P J Haagensen

Introduction to fatigue

Page 10

NTNU 2006

q =

maxeff - n K f n - n Kf - 1 = = max - n Kt n - n Kt - 1

(9)

where maxeff A is the effective maximum stress, see Fig. 28. This definition of Kf provides a scale for q that ranges from zero to unity. When q = 0, Kf = Kt = 1 and the material is fully insensitive to notches, i.e. a notch does not lower the fatigue strength. For extremely ductile, low strength materials such as annealed copper, q approaches 0. Also materials with large defects, e.g. grey cast iron with graphite flakes have values of q close to 0. Hard brittle materials have values of q close to unity. In general q is found to be a function of both material and the notch root radius. The concept of notch sensitivity therefore also incorporates a notch size effect. According to Neuber [5] the fatigue notch factor can be expressed as

Kf = 1 +

Kt - 1 1 + / r

(10)

where B is a characteristic size across which a large stress gradient cannot exist. Substituting the Neuber formula in the definition of notch sensitivity in Eq. 8 yields

q =

1 Kf - 1 = 1 + / r Kt - 1

(11)

Kuhn and Hardrath [6] found the relationship between C and the ultimate tensile strength given in Fig. 29. Peterson [7] proposed another, somewhat similar formula for the fatigue notch factor: 1 Kf - 1 q = = (12) 1 + a/r Kt - 1 where a also is a material parameter, which varies with type of material. Some examples are given as follows: Material: Parameter a: (in.) (mm) Normalized steel 0.01 0.25 Tempered steel 0.0025 0.064 Aluminum alloys 0.025 0.64

Later studies [8] of correlations of a with test data resulted in the following expression for steel: 1.8 300 (13) a = 10-3 = 28.8 S u-1.8 Su with a in inches, Su in ksi, or, in metric units: 1.8 2070 a = 0.025 = 2.32x 104 S -1.8 u Su a in millimeters, Su in MPa.

(14)

P J Haagensen

Introduction to fatigue

Page 11

NTNU 2006 The fatigue notch factor applies to the high cycle range, at shorter lives Kf approaches unity as the S-N curves for notched and unnotched specimens converge and coincide at N = 1/4 (tensile test). In experimental investigations involving ductile materials it was found that the fatigue notch factor need to be applied only to the alternating part of the stress cycle and not to the mean stress. For brittle materials, however, Kf should be applied to the mean stress as well. 6.4 Size effects

The size of a component affects the strength for various reasons. Although a size effect is implicit in the fatigue notch factor approach since Kf varies with r, a size reduction factor is normally employed when designing against fatigue. The need for this additional size correction arises from the fact that the notch size effect saturates at notch root radii larger than about 3-4 mm, i.e Kf Kt , while it is well known from tests on full scale components, also unnotched ones, that the fatigue strength continues to drop off with increasing size without any apparent limit. The size effect in fatigue is generally ascribed to the following sources, - A statistical size effect, which is an inherent feature of fatigue crack initiation which is a weakest link process where a crack nucleates when variables such as internal and external stresses, geometry, defect size and number, and material properties combine to give optimum conditions for crack nucleation and growth. Increasing size therefore produces a higher probability of a weak location. - A technological size effect, which is due to the different material processing route and different fabrication processes experienced by large and small parts. Different surface conditions and residual stresses are important aspects of this type of size effect. - A geometrical size effect, also called the stress gradient effect. This effect is due to the lower stress gradient present in a thick section compared with a thin one, see Fig. 30. If a defect, in the form of a surface scratch or a weld defect, has the same depth in the thin and thick parts, the defect in the thick part will experience a higher stress than the one in the thin part, due to the difference in stress gradient, as indicated in Fig. 30. - A stress increase effect, due to incomplete geometric scaling of the microgeometry of the notch. This takes place if the notch radius is not scaled up with other dimensions. Examples of components for which the latter effect is important are welded joints and threaded fasteners. The critical locations for crack initiation are the weld toe and the thread root, respectively. In both cases the local stress is a function of the ratio of thickness (diameter) to the notch radius. In welds the toe radius is determined by the welding process and is therefore essentially constant for different size joints. The t/r ratio therefore increases and also the local stress when the plate is made thicker, with r remaining constant. A similar situation exists for threaded fasteners, due to the fact that the thread root radius is scaled to the thread pitch, rather than the diameter for standard (e.g. ISO) threads. Since the pitch increases much slower than the diameter the result is an increase in the notch stress with bolt size. For bolts as well as welded joints this increased notch acuity effect comes in addition to the notch size effect discussed earlier, the result is that the experimentally determined size effects for these components are among the strongest P J Haagensen
Introduction to fatigue
Page 12

NTNU 2006 recorded. An example of size effects for welded joints is shown in Fig. 31. The solid line represents current design practice, according to e.g. Eurocode 3 and the UK Department of Energy Guidance Notes. The equation for this line is given by S t0 = t S0
n

(15)

The exponent n, the slope of the lines in Fig. 31a), is the size correction exponent. The experimental data points indicate that the thickness correction with n = 1/4 is on the unsafe side in some cases. As indicated in Fig. 31a) thickness correction exponent of n = 1/3 instead of the current value of 1/4 gives a better fit to the data in Fig. 31a). For unwelded plates and low stress concentration joints in Fig. 31b) a value of n = 1/5 seems appropriate [9]. Examples of size reduction factors to be used with baseline fatigue endurance data obtained from tests on small scale specimens are shown in Fig. 32. In Fig. 32) the same size correction is to be applied to both notched and unnotched parts. However, this may be an non-conservative approach for notched parts since the data in Fig. 32) indicates an increase in the size reduction factor with increasing stress concentration factor. This is in accord with other experimental findings that indicate a relationship between the stress gradient and the size effect. Based on an analysis of experimental data similar to those shown in Fig. 32b) the following size reduction factor has been proposed to account for the larger stress gradient found in notched specimens [10].
n = 0.10 + 0.15 log K t

(16)

6.5

Effects of surface finish

Almost all fatigue cracks nucleate at the surface since slip occurs easier here than in the interior. Additionally, simple fracture mechanics considerations show that surface defects and notches are much more damaging than internal defects of similar size. The physical condition and stress situation at the surface is therefore of prime importance for the fatigue performance. One of the important variables influencing the fatigue strength, the surface finish, commonly characterized by Ra, the average surface roughness which is the mean distance between peaks and troughs over a specified measuring distance. The effect of surface finish is determined by comparing the fatigue limit of specimens with a given surface finish with the fatigue limit of highly polished standard specimens. The surface reduction factor Cr is then defined as the ratio between the two fatigue limits. Since steels become increasingly more notch sensitive with higher strength, the surface factor Cr decreases with increasing tensile strength, Su. Typical examples of the relationships between the surface factor Cr and surface roughness Ra and Su are shown in Figs. 33 and 34. The data in Fig. 34 are given by the formula [11]

C r = 1 - 0.22 (log Ra )

0.64

log S u + 0.45 (log Ra )0.53

(17)

with Ra in m and Su in MPa. For aluminium alloys the effects are similar but few data exist on the influence of surface roughness.

P J Haagensen

Introduction to fatigue

Page 13

NTNU 2006

6.6

Residual stress effects

Residual stresses or internal stresses are produced when a region of a component is strained beyond the elastic limit while other regions are elastically deformed. When the force or deformation causing the deformation are removed, the elastically deformed material springs back and impose residual stresses in the plastically deformed material. Yielding can be caused by thermal expansion as well as by external force. The residual stresses are of the opposite sign to the initially applied stress. Therefore, if a notched member is loaded in tension until yielding occurs, the notch root will experience a compressive stress after unloading. Welding stresses which are set up when the weld metal contracts during cooling are an example of highly damaging stresses that cannot be avoided during fabrication. These stresses are of yield stress magnitude and tensile and compressive stresses must always balance each other, as indicated in Fig. 35. The high tensile welding stresses contribute to a large extent to the poor fatigue performance of welded joints. Stresses can be introduced by mechanical methods, for example by simply loading the part the same way service loading acts until local plastic deformation occurs. Local surface deformation a such as shot peening or rolling are other mechanical methods frequently used in industrial applications. Cold rolling is the preferred method to improve the fatigue strength of axi-symmetric parts such as axles and crankshafts. Bolt threads formed by rolling are much more resistant to fatigue loading than cut threads. Shot peening and hammer peening have been shown to be highly effective methods for increasing the fatigue strength of welded joints. Thermal processes produce a hardened surface layer with a high compressive stress, often of yield stress magnitude. The high hardness also produces a wear resistant surface; in many cases this may be the primary reason for performing the hardness treatment. Surface hardening can be accomplished by carburizing, nitriding or induction hardening. Since the magnitude of internal stresses is related to the yield stress their effect on fatigue performance is stronger the higher strength of the material. Improving the fatigue life of components or structures by introducing residual stresses is therefore normally only cost effective for higher strength materials. Residual stresses have a similar influence on fatigue life as externally imposed mean stresses, i.e. a tensile stress reduces fatigue life while a compressive stress increases life. There is, however, an important difference which relates to the stability of residual stresses. While an externally imposed mean stress, e.g. stress caused by dead weight always acts (as long as the load is present), residual stress may relax with time, especially if there are high peaks in the load spectrum that cause local yielding at stress concentrations.

6.7

Effects of corrosion

Corrosion in fresh or salt water can have a very detrimental effect on the fatigue strength of engineering materials. Even distilled water may reduce the high-cycle fatigue strength to less than two thirds of its value in dry air. Fig. 36 schematically shows typical S-N curves for the effect of corrosion on unnotched steel specimens. Precorrosion, prior to fatigue testing introduces notch-like pits

P J Haagensen

Introduction to fatigue

Page 14

NTNU 2006 that act as stress raisers. The synergistic nature of corrosion fatigue is illustrated in the same figure by the drastic lower fatigue strength which is obtained when corrosion and fatigue cycling act simultaneously. The strongest effect of corrosion is observed for unnotched specimens, the fatigue strength reduction is much less for notched specimens, as shown in Fig. 37. Protection against corrosion can successfully be achieved by surface coatings, either by paint systems or through the use of metal coatings. Metal coating are deposited either by galvanic or electrolytic deposition or by spraying. The preferred method for marine structures, however, is cathodic protection which is obtained by the use of sacrificial anodes or, more infrequently, by impressed current. The use of cathodic protection normally restores the high cycle fatigue strength of welded structural steels to its in-air value, while at higher stresses hydrogen embrittlement effects may reduce the fatigue life by a factor of 3 to 4 on life. 7 SUMMARY 1. Fatigue is a weakest link process of a statistical nature in which a crack initiates at a location where stress, local and global geometry, defects and material properties combine to give a worst case situation. The crack thus nucleates at a local weak spot, and may cause failure of the structure, even if the rest of the structure has a high fatigue resistance. Good fatigue design practice is therefore based on close attention to details that increase the stress locally and hence are potentially initiation sites for fatigue cracks. 2. The fatigue life consists of two stages, initiation and crack growth. In the initiation stage material strength, yield or ultimate strength has a strong influence on fatigue life, while the crack growth life is almost independent of static strength. 3. For mechanical components for which the crack initiation dominates there is a major influence of: Material strength Stress concentrations Externally applied mean stress Surface condition, particularly roughness and residual stress Size (thickness or diameter) Type of loading Environment (temperature and corrosion) The large number of factors influencing fatigue strength makes the combined effects of these factors difficult to predict. The safest way to obtain design data is therefore to perform fatigue test on prototype components under realistic conditions with respect to fabrication, loads and environment.

P J Haagensen

Introduction to fatigue

Page 15

NTNU 2006 8 REFERENCES Metals Handbook, American Soc. for Metals, 8th Ed., Vol. 9, 1974, Metals Park, Ohio. 2. ISO Standard 373-1964 3. P.C Paris, and F. Erdogan, "A Critical Analysis of Crack Propagation Laws," Trans. ASME, Vol.85,No.4, 1963. 4. J.M. Barsom, " Fatigue Crack Propagation ," Trans. ASME, Ser. B, No.4, 1971 5. H. Neuber, "Kerbspannungslehre," Springer 1958 6. P.Kuhn and H.F. Hardrath, "An Engineering Method for Estimating Notch-Size Effect in Fatigue Tests on Steel, NACA-TN-2805, 1952. 7. R.E. Peterson, "Stress Concentration Factors", John Wiley & Sons, 1974 8. R.E. Peterson, "Analytical Approach to Stress Concentration Effect in Fatigue of Aircraft Structures", Proc. Symp. on Fatigue of Aircraft Materials, WADC Technical Report 59-507, 1959. 9. O. rjaster et al., "Effect of Plate Thickness on the Fatigue Properties of a Low Carbon Micro-Alloyed Steel", Proc. 3rd Int. ECSC Conf. on Steel in Marine Structures (SIMS'87), Delft, 15-18 June 1987. 10. P.J. Haagensen, "Size Effects in Fatigue of Non-Welded Components", Proc. 9th Int. Conf. on Offshore Mechanics and Arctic Engineering, (OMAE), Houston, Texas, 1823 Feb. 1990. 11. "Leitfaden fr eine Betriebsfestigkeitsberechnung", Verein Deutcher Eisenhttenleute, 2. Auflage, 1985 1.

P J Haagensen

Introduction to fatigue

Page 16

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 17

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 18

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 19

NTNU 2006

E(f)

E(f)

Frequency, f

Fig. 12 S-N curves with and without fatigue limits, schematic.

P J Haagensen

Introduction to fatigue

Page 20

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 21

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 22

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 23

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 24

NTNU 2006

m = constant

R = constant

P J Haagensen

Introduction to fatigue

Page 25

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 26

NTNU 2006

Fig. 30. Stress gradient effect in welded joint, showing the stress at a defect to be higher in the thicker joint if the defect size is the same.

Fig. 31. Size effect in welded joints; a) as-welded joints, b) unwelded joints and welds improved by grinding.

P J Haagensen

Introduction to fatigue

Page 27

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 28

NTNU 2006

P J Haagensen

Introduction to fatigue

Page 29

Vous aimerez peut-être aussi