Sunteți pe pagina 1din 16

Fatigue & Fracture of Engineering Materials & Structures

doi: 10.1111/j.1460-2695.2010.01484.x

On the mechanism of fatigue crack propagation in ductile metallic materials


R . P I P P A N 1,2 , C . Z E L G E R 2,3 , E . G A C H 4 , C . B I C H L E R 1,5 a n d H . W E I N H A N D L 1
1 Erich

Schmid Institute of Materials Sciences, Austrian Academy of Sciences, Jahnstrasse 12, A-8700 Leoben, Austria, 2 Christian Doppler Labor fur die lokale Analyse von Verformung und Bruch, Jahnstrasse 12, A-8700 Leoben, Austria, 3 AMAG, A-5282 Ranshofen, Austria, 4 B ohler Edelstahl, Mariazeller Str 25, A-8605 Kapfenberg, Austria, 5 Plansee Metall GmbH u Plansee GmbH, A-6600 Muhl bei Reutte, Austria Received in final form 28 July 2009

A B S T R A C T An overview of our research performed during the last 15 years is presented to improve the

understanding of fatigue crack propagation mechanisms. The focus is devoted to ductile metals and the material separation process at low and intermedial crack propagation rates. The effect of environment, short cracks, small-scale yielding as well as large-scale yielding are considered. It will be shown that the dominant intrinsic propagation mechanism in ductile metallic materials is the formation of new surface due to blunting and the resharpening during unloading. This process is affected by the environment, however, not by the length of the crack and it is independent of large- or small-scale yielding. Keywords crack closure; fatigue crack propagation; fatigue mechanisms; long cracks; low cycle fatigue; short cracks.
NOMENCLATURE

a = crack length c = MansonCoffin exponent CTOD = crack tip opening displacement da/dN = fatigue crack propagation rate K max , K min = maximum and minimum stress intensity factor Nf = number of cycles to failure R = stress ratio K min /K max CTOD = cyclic crack tip opening displacement pl = plastic strain range J = cyclic J-integral J eff = effective cyclic J-integral K = stress intensity rang We = elastic component of the remote strain energy density range Wp = plastic component of the remote strain energy density range y = yield stress

INTRODUCTION

Figure 1 shows the fatigue crack propagation behaviour of austenitic steel. Such fatigue crack propagation rate, da/dN , versus stress intensity range, K , curves are typical for ductile metals. The fatigue crack propagation behaviour of such materials is characterized by: a steep drop of the crack propagation rate at the threshold of stress intensity range, K th , an extended Paris regime with a
Correspondence: R. Pippan. E-mail: reinhard.pippan@oeaw.ac.at Note: This manuscript was originally compiled for a Special Issue by Prof. C. Rodopoulos and Prof. Sp. Pantelakis based on the 1st International Conference of Engineering against Fracture.

relatively small Paris exponent between 2 and 4 and a pronounced increase of the propagation rate approaching the fracture toughness of the material. Despite the vast number of studies, which show similar results, there are a lot of very different explanations for this behaviour, some very similar, others very different. When considering the mechanisms controlling fatigue, it is very helpful to follow the classification into intrinsic and extrinsic fatigue mechanisms introduced by Ritchie.1 The large number of different proposed intrinsic fatigue mechanisms, or in other words the real material separation processes, can be divided into three groups: fracture due to accumulated damage in the volume in front of the crack, a pure

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

R . P I P P A N et al.

Fig. 1 Fatigue crack propagation rate as a function of stress intensity range of a cold-rolled austenitic steel at a stress ratio, R, of 0.1 and 0.8.

separation processes near the surface of a sample maybe somewhat different to the processes that determine the crack propagation along the rest of the crack front. In the last few years we, like other research groups, have performed several experiments and simulations to obtain a better understanding of the fatigue crack propagation mechanisms. A summary of these partly published results and some very recent findings will be presented in the following paper. First, we will consider the formation of striations, the most well-known features of fatigue. Then the reduction of the propagation rate near the threshold will be analysed by discrete dislocation mechanics. The effect of the environment is then investigated using experimental methods, which were used for the analyses of the formation of striations. The differences in the short crack regime are then considered, and lastly, a comparison of the processes in ductile and semi-brittle materials will be presented.

deformation induced crack propagation and a local fracture dominated fatigue crack propagation governed by cyclic damage or the applied strain. It is surprising that despite the extensive studies there are so many discrepancies in the explanation of, often, one and the same experimental result. If one is not familiar with the special details, one may believe that a decision is easy, whichever model describes the fatigue crack propagation in a certain loading region and in certain materials: one would only have to look what causes the separation of the material at the crack tip. If taking the length scales into account, which is essential for the processes at the crack tip, one can imagine that this really is a difficult task from both ways of looking at it in experiments and simulations. The crack length, a, the length at which the crack flanks can be partly in contact and the monotonic and cyclic plastic zone can vary from few nm to the size of the sample or component. The crack propagation rate and the cyclic plastic opening of the crack can vary from atomic distances to few 10 m. For experimental observations one has, furthermore, to take into account that the material

FORMATION OF STRIATIONS

Striations as depicted in Fig. 4 are the characteristic features on the fracture surface for fatigue in many ductile metals and alloys in the mid Paris regime.2 Different models for the formation of these characteristic markings are proposed,39 the two most well known are shown in Fig. 2. A summary of these early studies are published in an overview by Tomkins.9 The base of all these models are the fractographs obtained by scanning electron microscope, replica technique and thin foil transmission electron microscopy of the material just below the fracture surface or free surface observations, which in many cases are not representative for the propagation mechanism along the crack front. From the simple fractographic features alone one cannot distinguish which model gives the best description of the crack propagation process. Therefore, we developed a special technique to determine the details of the deformation of a crack tip and the crack propagation process in the midsection of a specimen. It is

Fig. 2 Schematic representation of different textbook explanations for the formation of striations: (a) Laird model, (b) Pelloux model.

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

MECHANISM OF FATIGUE CRACK PROPAGATION

load

Specimen X

time

(a)
load

Specimen Y
time

(b)
Fig. 3 Loading procedure to determine the shape of the crack tip in a multi-specimen experiment, examples are presented for K max and the case of 50% K max during loading.

based on a fractographic technique in combination with a stereophotogrammetric reconstruction of the fracture surface; for details see Refs [10, 11]. This technique is usually used to study processes during overload fracture.1012 For the determination of the fracture process during cyclic loading a special loading procedure was developed.13,14 Figure 3 shows, schematically, the loading procedure. It is a multi-specimen experiment, in order to characterize the shape of the crack tip at a certain load, one sample is necessary. For the determination of the results presented in Fig. 4, the procedure will be described shortly. The material in this case was a cold-rolled austenitic steel A 220 (Bohler steel, which corresponds about to a 316L) with a

yield stress of 890 MPa and an ultimate tensile strength of 1010 MPa. The fatigue crack propagation behaviour of this material is depicted in Fig. 1. The grain size was between 1 and 2 mm (more details about the material, see Refs [13, 15]). This material forms nice striations in a K regime between 20 and 120 MPa m. The stress intensity range K for the presented experiment was 70 MPa m at a stress ratio of 0.05, where the striation spacing is about 1 m. This specimen was fatigued with the chosen load amplitude to obtain a steady state. At a chosen load, in the presented example at K max , the test was interrupted. Then a constant small load amplitude sequence with K of about 10 MPa m was applied until the specimen failed. This small K loading is used to freeze the shape of the crack tip. The shape of the fracture surface of both specimen halves was then reconstructed. This was done by analysing stereoscopic scanning electron (SEM) images of exactly corresponding fracture surfaces with an automatic image processing system. Figure 4a and b shows an example of such corresponding regions taken from the midsection of the sample. The fracture surface generated at K = 70 MPa m and at small load amplitude is clearly visible. The automatic image processing of the stereo images leads to 3D models of the fracture surfaces. By fitting together the two 3D models in the post fatigue region, one obtains the shape of the crack tip at the corresponding load (in this case at K max ). An extraction of profiles is also possible, see Refs [13, 15]. The perfect fitting of the 3D models of the post fatigue fracture surfaces, the blunting region and the striations, which are formed in the previous cycles, are clearly visible. It is evident that during the loading in the blunting region (marked with A in Fig. 3c) a new fracture surface is generated. The

Fig. 4 Fracture surface of the sample interrupted at K max = 74 MPa m. (a) and (b) show exactly corresponding fracture surface at the transition from K = 70 MPa m to small K region. (c) Exhibits the reconstructed 3D shape of the crack tip in the midsection of the specimen. The new generated fracture surface in the blunting region is marked with A and the last formed striations with B. (d) shows the load versus CTOD measured 2 m behind the crack tip determined from the evaluated 3D shape of crack tip.

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

R . P I P P A N et al.

extension of this region is comparable to the spacing of the striations. In Fig. 3c, the width of the striation formed in the last cycle is marked with B. The same experiments were performed by interrupting the test at 50 and 75% of K max during loading and at 75, 50 and 25% of K max during unloading. Profiles of the crack tip are presented in Ref. [13]. The determined crack tip opening displacement (CTOD) from these profiles, 2 m behind the crack tip, are depicted in Fig. 4d. It is evident that the fracture surfaces at the crack tip are in contact below about 25% of K max and during loading the crack blunts as expected (similar as in finite element simulations, see e.g. Ref. [16]). During unloading the crack re-sharpens and closes near the crack tip between 20 and 25% of K max . As mentioned, several models have been proposed to explain the formation of striations. Most of these can be classified in two groups. The first is based on a plastic forming of new fracture surfaces,46 others involve brittle fracture at the crack tip in association with cyclic hardening, for example Refs [6, 17]. A further point of controversy is the formation of the fissure or trench. Laird3 assumes the formation during unloading at the tip of the blunted crack (the new formed striation), whereas Refs 6 and17 suppose that the trench is generated at the loading part. From our observations (for details of the crack tip shape at different K during the load cycle see Refs [13, 18]) we can draw a simplified model, which is sketched in Fig. 5. The model assumes a symmetric deformation at the crack tip, which is an idealization, but it should be noted that in the asymmetrical case the behaviour is similar, which is supported by experimental observations.13,18 The formation process can be separated in the following steps:
At the minimum of the load in a constant load amplitude experiment, the crack is sharpened and closed. At about 25% of K max at R = 0.05 the crack tip opens. Additional loading causes a blunting and forms a V-shaped micronotch. The plastic flow is mainly concentrated at the tip of this V-shaped micro-notch. The blunting process continues till K max is reached. The new crack surface was produced by the formation of this V-shaped micro-notch. Load reversal causes re-sharpening, i.e. shearing of the activated slip bands in opposite direction occurs. This deformation is again concentrated at the tip of the V-notch. This continues until the crack is fully closed. The fissure is formed between the last two striations. During loading, a corner is formed between the last striation and the new created fracture surface in the blunting region. During unloading the shearing is not strictly reversible, the material on the corner between the previous formed striation and the partly closed micro V-notch generates a fold, which results in a fissure.

Fig. 5 Schematic representation of the crack tip deformation and propagation model obtained from the different 3D shapes of the crack tip during cyclic loading.

THE NEAR THRESHOLD BEHAVIOUR

The experimental technique described above is not applicable to analyse the fatigue process near the threshold of stress intensity range. The CTOD and the crack extension per cycle is so small that it cannot be detected by SEM techniques. It may be that, in the future, in situ transmission electron microscope fatigue experiments will offer the possibility to study these phenomena; however, it is difficult to apply defined load to the crack tip and to avoid buckling of the necessary thin foil during fatigue unloading. Therefore, we simulated the propagation of cracks at such small K . Because the size of the cyclic plastic zone is in the order of the typical dislocation distance and dislocation substructure and, furthermore, the cyclic deformation is in the order of atomic distances, classical plasticity theory cannot be applied. Therefore, we used a discrete dislocation model, which is described in detail in Refs [1921]; similar approaches are used in Refs [2225]. Discrete dislocation modelling is performed in the framework of linear elasticity, where the non-linearity is taken into account by the movement of discrete dislocations. The crack propagation mechanism we used is schematically depicted in Fig. 6. We assumed that the crack tip is the dislocation source, which is a good approximation for small stress intensities.19,26,27 Dislocations are generated at the crack tip, when the local stress intensity factor

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

MECHANISM OF FATIGUE CRACK PROPAGATION

Fig. 7 Change of the cyclic CTOD as a function of crack extension determined by a discrete dislocation simulation. The used material parameters and the details of the simulations are described in detail in Ref. [29]. CTOD values and a are given in units of Burgers vector, b.

Fig. 6 Schematic representation of the fatigue crack propagation mechanism on the atomistic scale at small K .

at the crack tip is larger than a certain ke .1 In the case of mode I loading with symmetric arrangement of slip planes the generated dislocations form a similar V-shaped notch, which is what we have observed at larger K on a much larger scale. During unloading most of the generated dislocations returned to the crack tip. It is assumed that the crack flanks do not reweld; hence, the crack propagation rate is proportional to the new generated surface during loading and, as a consequence, it is proportional to the CTOD, which is under constant amplitude loading equal to CTOD. Figure 7 shows the calculated variations of CTOD as a function of crack extension for different constant K simulations. It is evident that CTOD decays with increasing crack extension and reaches saturation, or at very small K , CTOD vanishes and the crack stops the propagation in the simulation. This initial decrease in crack growth rate is mainly caused by the increase in the effect of crack closure. This is clearly evident
1

A similar local k one can obtain, if one assumes a single dislocation source very near the crack tip.

from Fig. 8, which shows the arrangement of the dislocation and the contour of the crack flanks at K min of a crack loaded with K = 2.5 ke after a crack extension of about 20 m. CTOD has decreased in this case from few 10 to about 1 Burgers vector. The crack is closed only over few micrometres behind the crack tip. It is important to note that a contact over such relatively short distance of the fracture surfaces causes a significant reduction in the cyclic plastic opening of the crack. Due to the assumed symmetry, no deflections of the crack takes place; hence, the developed contacts of the fracture surface can be denoted as plasticity induced crack closure. The simulations assume plane strain conditions; hence, this simulation which is a superposition of a larger number of exact linear elastic solutions shows clearly the significance of plasticity induced crack closure even in the near threshold region. This reduction of growth rate is well known from experiments; for example in the near threshold regime, see Ref. [28]. However, also at large K this effect is always visible.14,18 An example is presented in Fig. 9, showing a fracture surface, where a pre-fatigue crack is generated with a small K and then the crack propagation experiment was continued with a constant K . One can see that the striation spacing decreases and reaches a saturation value similar to that in the simulation in Fig. 7. The obtained CTODi , in the first few cycles and CTODs , in the steady state case, are depicted in Fig. 10. The monotonic CTOD as a function of load is given by the CTOD versus K curve. For K max smaller than ke no plastic deformation occurs, the crack tip is loaded elastically. For K max larger than ke dislocations are generated, a small

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

R . P I P P A N et al.

Fig. 8 Arrangement of dislocations and the crack contour at K min (b and d) and K max (a and c) after a crack extension of about 20 m. The material parameters used are the same as for calculations presented in Figs 7 and 10.

blunting takes place. The repulsive forces between the few generated dislocations are, however, too small that the dislocations return during unloading. For a cyclic plastic deformation, the load amplitude has to be increased furthermore. At a certain critical value which is about 1.4 ke , the first cyclic plastic deformation starts. However, with increasing crack extension, CTOD decreases as shown in Fig. 7 and the crack stops the propagation after a certain extension due to crack closure. For K larger

than about 2.5 ke , such vanishing of the cyclic plastic deformation does not take place. The steady state CTODs increases relatively quickly with K and then approaches a constant slope in a lg CTOD versus K plot. At larger K all curves CTOD versus K , CTODi versus K , and CTODs versus K show the expected K 2 dependence. The progressive decrease of the monotonic and cyclic crack tip deformation, and as a consequence, the progressive decrease of the fatigue crack

Fig. 9 Fractograph of a fatigue fracture surface in the same austenitic steel as described in section, formation of striations. The pre-crack was generated with a small K , then a constant load experiment was performed with K = 70 MPa m and R = 0.05. The initial decrease in the striation spacing and the approach of a steady state growth rate are illustrated.

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

MECHANISM OF FATIGUE CRACK PROPAGATION

THE EFFECT OF ENVIRONMENT

Fig. 10 CTOD in the first cycle, the cyclic CTOD in the first cycle, CTODi and the steady-state CTODs determined from discrete dislocation simulation of a mode I crack. The used material parameters are given in Ref. [29].

propagation rate, are not a mysterious change in the crack propagation mechanism, it is simply a consequence of the discrete nature of plasticity. A further interesting result of the discrete dislocation simulation is the formed small steps on the fracture surfaces. They are a direct consequence of the arrangement of dislocations in the wake of the crack tip in bends with spacing somewhat larger than 1000b. The bend-like arrangements of the dislocation are caused by the long range interaction force between dislocations.30 The steps formed by these remaining dislocations are responsible for the abnormal striation spacing, which is frequently observed in the lower Paris regime and near threshold. The spacing of these steps is much larger than the crack growth rate. Therefore, the occurrence of such abnormal striations does not indicate discontinuous fatigue crack propagation. These simulations show that a fatigue crack can propagate cycle by cycle by a blunting and re-sharpening process from the threshold to the upper Paris regime. The simulations show furthermore that plasticity induced crack closure is more pronounced in the near threshold regime than at larger K .2 Discrete dislocation simulations of short cracks23,24 have shown that propagation of these cracks can be described in a similar way. The discrete nature and the interaction of the dislocation with the grain boundary are responsible for the transition from stage I to stage II propagation and the characteristic zigzag growth of fatigue cracks at small load amplitudes.32

Except the case, when a transition from a plane strain to a plane stress dominated plastic deformation occurs in the Paris regime.29,31 In such thin samples plasticity induced crack closure is significantly larger than in the plane strain case.

It is well known that the environment can significantly affect the fatigue crack propagation rate (see, e.g. Refs [33, 34]). In order to investigate this effect, similar experiments are performed as those described in chapter 2. The experiments are performed in the austenitic steel A220 and the aluminium alloy 7020.35 Fatigue experiments are performed in 3.5 NaCl solution and in ultrahigh-vacuum (P < 107 Torr). The results were then compared with the behaviour in air. K was chosen in such a way to obtain a crack propagation rate in air of about 1 m/cycle. The austenitic steel was fatigued at K = 70 MPa m and the 7020 alloy at K = 20 MPa m. In both cases the stress ratio was 0.05. In order to examine the effect of the different environments, the shapes of the crack tips at K max in the midsection of the specimens were compared. Fig. 11 shows the fractograph of the 7020 alloy fatigued at first in air, then continued for 20 loading cycles in ultra-high vacuum interrupted at K max and then fatigued with a very small K with the same K max in air. The different regions can be clearly distinguished on the fracture surface. Fatigue in air at K = 20 MPa m induces clearly visible striations, in ultra-high vacuum they completely disappear, and the growth rate is significantly reduced. This effect is well known from the studies of Pelloux.37 The shape of the crack tip determined from the stereophotogrammetric analyses of the corresponding regions shows a rounded crack tip. A similar change was observed also in the austenitic steel in ultra-high vacuum.35 Figure 12 shows the fractographs and the determined crack shape of the austenitic steel fatigued in salt water at K max . The large K of 70 MPa m generates striations again, only the spacing in salt water is somewhat larger than in air. A comparison of the crack tip profile in Figs 4 and 12 indicates that the angle of the formed V-shaped micro-notch is somewhat smaller than in air. Table 1 and Fig. 13 summarize the data obtained from the crack tip profiles and fractographs. The CTOD is not affected, as expected, by the environment. The small differences are in the order of the scatter of such measurements. The crack tip blunting angle differs significantly. The determination of crack tip blunting angle of samples fatigued in ultra-high vacuum is not as defined as in air or salt water, because the crack tip is really blunted, whereas in the other cases the crack tip forms a pronounced sharp micro-notch. This indicates that the formed oxide layer during loading in air, and the corrosion layer induced by the aggressive environment change significantly the details of the blunting of the crack tip. It seems that the breaking of the oxide layer or corrosion layer induces a pronounced concentration of the deformation to the tip of the V-shaped micro-notch. The significant difference of the crack propagation in different environment is

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

R . P I P P A N et al.

air

UHV

y [mm]

0.020

crack propagation direction

0.010
side B

da/dN

0.000
side A 2 2 3 3

-0.010 0.015

x [mm]
0.045

0.025

0.035

Fig. 11 SEM-fractograph from two corresponding fracture surfaces of the 7020 alloy. The determined height profiles give the shape of the crack tip at K max fatigued in ultrahigh vacuum.

0.002 -0.001 -0.004

y [mm]

0.005

crack propagation direction

side B da/dN 2 2

3 side A

-0.007 -0.010

x [mm]
0 0.003 0.006 0.009 0.012 0.015 0.018 0.021

Fig. 12 SEM-fractograph of corresponding fatigue fracture surfaces of the austenitic steel, A220, fatigued in 3.5% NaCl solution. The two fitted height profiles show the shape of the crack tip at the maximum load.

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

MECHANISM OF FATIGUE CRACK PROPAGATION

Table 1 Summary of determined CTOD, measured 7 m behind the crack tip, crack tip blunting angle, CTBA, from the crack profiles and the crack propagation rate per cycle Austenitic steel CTOD m Vacuum Air Salt water 4.6 4.3 4.7 CTBA C 105 90 60 da/dN m/cycle 0.5 1 1.2 7020 Al alloy CTOD m 5.0 4.3 4.7 CTBA C 130 90 25 da/dN m/cycle 0.3 0.7 1.7

Fig. 13 Schematic representation of the shape of the crack tip at maximum load in air, ultrahigh vacuum and NaCl solution.

mainly caused by the difference in the opening angle of the V-shaped micro-notch.33,36 Vehoff 36 comes to similar conclusions from surface observations. Even nanofracture of the material induced by the embrittlement due to the environment may cause such change of the opening angle.33 As mentioned in ultra-high vacuum, the crack tip is extensively blunted, it seems that the deformation is more homogeneously distributed around the crack tip. The shape is similar to that obtained in detailed finite elementsimulations.38 In ultra-high vacuum it seems that most of the generated surface during loading shrinks during unloading, only a small part remains for the crack propagation.39,40 This shrinkage cannot be a simple reversible motion of dislocation, because the involved number of dislocations in the mid and upper Paris regime is too large. In the lower Paris regime and near the threshold this number is smaller; hence, a simple reversible motion, i.e. a return of dislocations and disappearance of the formed surface step, is more likely take place. The difference in the fatigue crack propagation rate in the near threshold regime is therefore more pronounced. This is shown in

Fig. 14. Here the fatigue crack propagation rate in the threshold regime in air and ultra high vacuum is compared. The threshold is not very significantly affected. However, the crack growth rate in vacuum at K values somewhat larger than K th is two or three orders of magnitude smaller than in air.3 The drop in the propagation rate at the threshold in high vacuum is not as pronounced as in air. As a consequence the threshold of stress intensity range in vacuum can depend significantly on the propagation rate which is used to define it.39,41 A more detailed discussion of the effect of irreversibility of the generated fracture surface, especially related to the modelling point of view is presented in Ref. [42]. With larger K the number of cyclic generated dislocations increases, the probability of a reversible motion of dislocation decreases, which results in a decreasing of the difference between air and vacuum. In the mid and upper Paris regime the difference in the growth rate approaches

This can have significant effect of the propagation of cracks from bulk defects and may cause some of the ultrahigh cycle fatigue phenomena.40

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

10

R . P I P P A N et al.

Fig. 14 Effect of ultrahigh vacuum on the fatigue crack propagation behaviour in technical pure Fe (ARMCO-Fe).39

a constant value, da/dN vacuum = 0.30.5 da/dN air , which was also reported in Ref. [4].

DO SHORT FATIGUE CRACKS PROPAGATE BY A DIFFERENT MECHANISM?

Plotting the propagation rate of short fatigue cracks in a standard da/dN versus K diagram shows often significant deviations from the long crack data.43,44 Reasons for the differences are the changes in crack tip shielding, the non-applicability of small-scale yielding parameters or the difference between the micromechanical response and assumed homogeneous material behaviour. Depending on the responsible mechanisms, one can distinguish, therefore, physically (or extrinsically), mechanically and microstructurally short cracks.45 A fundamental question remains usually open do short cracks propagate by a different intrinsic mechanism? In order to answer this

question for a ductile material, the austenitic steel A220, short crack experiments are performed in situ in a SEM46 and compared with the long crack behaviour described in section 2, formation of striations. The used samples are depicted in Fig. 15. A short notch is machined by a razor blade polishing technique at the corner of the sample. A cut with the focused ion beam at the root of the notch generates a short crack-like defect. Samples with notch depth between 30 and 140 m are investigated. The samples are loaded with constant displacement amplitude. Due to the hour glass contour of the sample, an accurate determination of the applied far field strain from the load versus cross head displacement is not possible. A determination of the applied far field strain amplitude from the SEM images taken at the maximum and the minimum load exhibits a plastic strain amplitude, pl , of about 6%, for the case depicted in Fig. 16. At different load cycles at the maximum load and for few selected cycles SEM micrographs at certain loads during loading and unloading are taken from the crack tip. An example is presented in Fig. 16. It shows that the crack at the minimum load is closed. During the loading the crack blunts and re-sharpens during unloading. The cyclic crack tip opening is therefore equal to the crack tip opening at maximum load, i.e. CTOD = CTOD as in the case of a steady state propagating long cracks. In Fig. 17 the measured CTOD are plotted as a function of the crack growth rate. The data obtained from Fig. 4, where the samples had a crack length of about 30 mm are also indicated. It is evident that the crack propagation rate is about one-fourth of CTOD independent of the crack length or notch depth. The long crack experiments were performed under well-defined small-scale yielding conditions, whereas the short crack experiments are in the extreme low cycle fatigue regime. Despite this difference the relation between CTOD and da/dN is not affected. On the contrary, in the standard da/dN versus K diagram the difference between the long and short crack data is huge (see Fig. 17a). The grain size of the austenitic steel A220, used for the short crack experiments in Figs 16 and 17 is about 10 m. Therefore, the investigated cracks are microstructurally long and mechanically and extrinsically short. Similar experiments are also performed in a very coarse-grained

Fig. 15 Representation of the sample geometry for in situ short crack examination and the SEM image of the notch.

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

MECHANISM OF FATIGUE CRACK PROPAGATION

11

Fig. 16 Load-displacement curves of a fatigue experiment and selected SEM micrographs of the crack tip during one loading cycle. The loads, where the micrographs are taken, are indicated in the load displacement curve.

10

-5

short crack R=-1

da/dN [m/cycle]

10

-7

pl=2-6%

10 10

-9

-11

long cracks

10

-13

R=0.1 R=0.8

10

100
1/2

(a)

K [MPam ]

(b)

Fig. 17 CTOD determined from the SEM micrographs at different crack propagation rate at samples with different depth of micro-notches and different crack lengths. For comparison, the long crack data determined in the same steel in a coarse-grained cold-rolled conditions are also indicated.

A220 steel with a grain size of few millimetres.46 The results are very similar, only the scatter was larger. One of the reasons for this large scatter is the uncertainty in the correct determination of CTOD and da/dN , which is induced from strong mixed mode loading, pronounced crack branching and crack deflections of the crack in the first grain. The determination of CTOD of a stage I like crack is more difficult than for the well-defined stage II case as shown in Fig. 17b. Furthermore it has to be noted that a plastic deformation of the crack flanks somewhat

behind the crack tip can cause an opening of the crack but no crack extension, as schematically depicted in Fig. 18. Hence, one has to distinguish between crack tip opening, which generates new fracture surface at the crack tip and a part of the CTOD, caused by the cyclic deformation of the crack flanks in the vicinity of the crack tip. For the determination of CTOD in Fig. 17a we used the definition of CTOD of Ref. [47], this is a good approximation for the deformation direct at the crack tip, which generates new fracture surfaces for a mode I crack.

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

12

R . P I P P A N et al.

Kmin Kmax

CTOD

CTODeff

Kmin
Fig. 18 Schematic representation of the deformation at and in the vicinity of the crack tip in order to illustrate the contribution of CTOD, which contributes to the fatigue crack growth.

The described experimental results indicate that short and long cracks in ductile metals propagate by the same mechanism. Hence, for life time prediction, one only has to find the appropriate driving force to describe CTOD. In the case of small-scale yielding, K or K eff and for large-scale yielding J or a corresponding effective cyclic J -integral, J eff are the standard fracture mechanics parameters. As a consequence, the low cycle fatigue life time, i.e. the MansonCoffin relation, Eq. (1) and Refs [48, 49]. pl /2 = const( N f )c (1)

should be describable in terms of crack propagation and should follow from an integration of the crack growth curve. For a semicircular surface crack J J 3.2 We a + 5 Wp a (2)

where We and Wp are the elastic and plastic components of the remote strain energy density ranges.50 J and CTOD for an open crack are proportional CTOD = d n J 2 y (3)

where dn is a parameter depending on strain hardening and the ratio of the yield stress, y , to the Youngs modulus. For low cycle fatigue experiments with very large pl the elastic component is small, for elastic ideal plastic material Wpl = 2 y pl and dn = 0.8, therefore CTOD 4 pl a . (4)

In such case the number of cycles to failure should be about inverse proportional to pl . However, it is well known that the MansonCoffin exponent is significantly smaller than 1 (typically between 0.5 and 0.7). One of the reasons for the deviation is that the change of the critical crack length at different plastic strain amplitudes are not taken in to account. A more accurate estimation gives that Nf should be proportional to ln (amax /a0 )/ pl , where a0 is the initial flow size and amax is the maximum length of the crack in fatigue. However, in most cases ln (amax /a0 ) is not strongly affected by pl . The main reason for the deviation from this simple estimation is that J or pl is applicable to describe CTOD for cracks only, which are open during the complete part of the load cycle. Otherwise one has to use J eff . Not the full amplitude J or pl only a smaller part J eff or pl eff , are responsible for the observed CTOD. This importance of crack closure is clearly visible if one considers the incremental strain maps determined during the unloading part of a load cycle in Fig. 19. In each of the load steps during unloading, the typical localization of the deformation in the vicinity of crack tip is visible; however, in the last load step (near the minimum load) the characteristic deformation pattern near the crack tip vanishes. This clearly indicates that the crack is closed during this part of the loading cycle. In the present example J eff is about 80% of J . An essential point to describe the low cycle fatigue behaviour in terms of fracture mechanics is, therefore, to predict the effective driving force as a function of pl and the loading history. In constant or pl experiments, the effective J or pl should decrease substantially as schematically depicted in Fig. 20 with decreasing pl . In the case of microstructurally short cracks, CTOD or CTOD values varies significantly between samples loaded with the same macro displacement and the same crack length.46 However, that is not surprising because for cracks smaller than the grain size the applied far-field J or J eff do not describe solely the local driving force in such cases. A micromechanic approach is necessary (see, e.g. Ref. [51]), where the local microstructure has to be taken into account for a correct description of the driving force for crack propagation.
FATIGUE CRACK PROPAGATION IN LESS DUCTILE MATERIALS

The number of cycles to failure is then given by the integration of the growth from an initial flaw size to a critical value. Taking into account that crack propagation rate is proportional to CTOD and using Eq. (4) leads to d a /d N proportional to pl a . (5)

The experiments described in section, the effect of environment, indicate that the crack tip blunting angle can vary, depending on material environment and may be on CTOD, which seems to be caused by an additional fracture process on the atomistic scale at the tip of the blunted crack. Such effect of local fracture on the nano-scale in the immediate vicinity of the crack tip is also proposed by

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

MECHANISM OF FATIGUE CRACK PROPAGATION

13

Fig. 19 Incremental strain maps determined during unloading. The load increments are marked in the load displacement curve. The change of the crack opening induces the apparent strain localization in the horizontal line in the midsection of the map. In front of the crack the typical strain distribution of crack is visible. The last strain map indicates that the contact of the fracture surfaces induces a disappearance of the characteristic crack tip deformation. For details of the determination of such maps see Refs [56, 57].

Fig. 20 Schematic illustration of the effective the plastic strain amplitude.

J eff as a function of

others (see, e.g. Ref. [33]). Near the fracture toughness in addition to the blunting a formation of voids and microcracks causes the pronounced acceleration of growth rate. In brittle materials, such phenomena may take place even in the Paris regime or near the threshold. From a single da/dN versus K it is usually not straight forward to decide, which crack propagation mechanism is the dominant one. Figure 21 shows such example. The long crack da/dN versus K curve of this designed fully lamellar Ti-46.5, Al-4 (Cr, Nb, Ta, B) alloy looks very similar to the crack propagation curve of a ductile metal.52 In order to determine the fatigue crack propagation mechanisms, fatigue

crack growth experiments with constant load amplitude on deep sharp notches containing very short pre-cracks are performed. The experiments were performed at K values significantly smaller than and larger than the long crack threshold. In the first cycle, the mean increase in crack length detected by the direct current potential drop technique was very large compared to the growth rate of long cracks. Even in the first few 10 cycles the crack propagates were significantly larger (more than one order of magnitudes) than the long crack data. In ductile materials in such experiments a decrease in the growth rate is also observed, see Figs 7 and 9. However the effect is much smaller, it is an effect of the development of crack closure. The decrease in the propagation rate in the Paris regime, at R = 0.1 is about a factor of 2. In this intermetallic alloy it is about 2 orders of magnitude, which is shown in Fig. 21. In this da/dN versus K diagram the crack propagation rate of a long crack (steady state propagation) the crack extension during the first load cycle and the propagation rate in the first few cycles are indicated. Blunting and re-sharpening of the crack tip takes place also in this alloy; however, it does not significantly contribute to the crack propagation per cycle. The dominant intrinsic process is the cleavage of the lamellar microstructure.53 The long crack fatigue propagation is dominated by the extrinsic processes crack bridging, crack branching and crack closure. This clearly becomes evident if one compares the

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

14

R . P I P P A N et al.

Fig. 21 Mean crack extension in the first cycle obtained in fatigue tests on short cracks in deep-sharp notched specimens versus K , the mean crack propagation rate in the first few cycles in these types of tests and the long crack da/dN versus K in a coarse-grained designed fully lamellar Ti-46.5 at% Al 4 at%(Cr, Nb, Ta, B) alloy.

Fig. 22 Fatigue crack propagation rate versus the stress intensity range at different stress ratios in a particle reinforced 359 cast aluminium alloy.

crack propagation rate in the first few cycles of the fatigue experiments in the samples, which are performed on short pre-cracked samples with a small contribution of crack tip shielding and the long crack data. This example indicates that crack propagation curves, which look similar to those of ductile metals, can be caused by different mechanisms. One may argue that this change in crack propagation mechanism in the intermetallic material is not surprising. However, sometimes such change to static fracture dominated fatigue cracked propagation over a relatively wide range may occur even in frequently used metallic materials. A relatively simple way to get an impression, whether static fracture over distances significantly larger than CTOD dominates the fatigue crack propagation

or not, is to compare da/dN versus K curves at different R ratios. If the effect of R in the Paris regime is relatively low, as in the case of Fig. 1, the crack propagation is governed by CTOD. If the increase in R induces a progressive increase of the propagation rate in the Paris regime, as presented in Fig. 22 and Ref. [54] static fracture at distances significantly larger than CTOD or CTOD gives an important contribution to the fatigue crack propagation rate.
CONCLUDING REMARKS

Fatigue crack propagation mechanisms are essential to understand fatigue properties of materials. From an en-

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

MECHANISM OF FATIGUE CRACK PROPAGATION

15

gineering point of view one may assume that these mechanisms are not important. What is important here is only the crack propagation rate at certain loading conditions. In standard fatigue crack propagation experiments such data are determined at constant load amplitude and usually under mode I loading. However, engineering components are usually loaded by variable amplitudes and in many cases complex multiaxial loading is dominant; small or large-scale yielding conditions and sometimes a combination of both loading cases can be present. In order to solve this problem, engineers and scientists have developed empirical and half empirical methodologies to estimate the life time of their components. However, one has to take into account that a methodology developed for a certain material for certain loading cases can be applied to other materials under similar loading condition, only when the fatigue crack propagation mechanisms are the same. This indicates why the fatigue crack propagation mechanisms are important for both material development and engineering design. In ductile metals the fatigue crack propagates by a formation of new fracture surface during loading by blunting at the crack tip, during unloading the crack tip re-sharpens and the crack flanks come into contact. The crack tip blunting angle and the part of loading during the crack opening, where the new surface is generated, governs the proportionality factor between da/dN and CTOD. The cyclic deformation of the crack flanks in the vicinity of the crack tip does not contribute to the crack extension, which has to be taken into account in the relation between CTOD and da/dN . In the authors opinion for each ductile metal or alloy there exists a unique relation between CTOD and da/dN as long as the static fracture does not dominate the processes in front of the crack tip. There are materials, where there exists a simple linear relation between da/dN and CTOD as shown in examples in this paper; however, there may be materials with a progressive increase of da/dN with increasing CTOD,55 the crack tip blunting angle and the ratio of the part of CTOD, which generates the new fracture surface and the part of CTOD, which is caused by a deformation of the crack flanks in the vicinity of the crack tip may cause a deviation from the linear relation between CTOD and da/dN . In short and long cracks in ductile metals the crack propagates by the same blunting and re-sharpening mechanism independent of small- or large-scale yielding. Even in the near threshold regime the intrinsic process remains the same, only the discrete nature of plasticity begins to play a dominant role. The focus of this paper was devoted to intrinsic fatigue crack propagation mechanisms in ductile metallic materials. However, the presented examples have demonstrated the importance of crack closure. Even in the low cycle

fatigue regime the part of the amplitude, where the crack flanks are in contact, is significant. It was shown that the variation of the effect of crack closure as a function of the plastic strain amplitude is essential for the explanation of the low cycle fatigue behaviour by means of fatigue crack propagation.
REFERENCES
1 Ritchie, R. O. (1988) Mechanisms of fatigue crack propagation in metals, ceramics and composites: role of crack tip shielding. Mater. Sci. Eng. 103, 1528. Zapffe, C. A. and Worden, C. O. (1951) Fractographic registrations of fatigue. Trans. ASM 43, 958969. Laird, C. (1967) The Influence of Metallurgical Structure on the Mechanisms of Fatigue Crack Propagation. Fatigue Crack Propagation, ASTM STP 415, Philadelphia PA, 131. Pelloux, R. M. N. (1970) Crack extension by alternating shear. Eng. Fract. Mech. 7, 235247. Neumann, P. (1974) Geometry of slip processes at a propagating fatigue crack. 2. Acta Met. 22, 11661178. Tomkins, B. and Biggs, W. D. (1969) Low endurance fatigue in metals and polymers - Part 3. The Mechanisms of failure. J. Mater. Sci. 4, 544553. McEvily, A. J. and Boettner, R. C. (1963) On fatigue crack propagation in F.C.C. metals. Acta Met. 11, 725743. Furukawa, K., Murakami, Y. and Nishida, S. (1998) A method for determining stress ratio of fatigue loading from the width and height of striation. Int. J. Fatigue 20, 509516. Tomkins, B. (1996) The mechanism of stage II fatigue crack growth. Fract. Eng. Mater. Struct. 19, 12951300. Kolednik, O. (1981) A contribution to stereophotogrammetry with the scanning electron microscope. Pract. Metallogr. 18, 562573. Stampfl, J., Scherer, S., Berchthaler, M., Gruber, M. and Kolednik, O. (1996) Determination of the fracture toughness by automatic image processing. Int. J. Fract. 78, 35 44. Semprimoschnig, C. O. A., Stampfl, J., Pippan, R. and Kolednik, O. (1997) A new powerful tool for surveying cleavage fracture surfaces. Fatigue Fract. Eng. Mater. Struct. 20, 15411550. Bichler, C. and Pippan, R. (1999) Direct observation of the formation of striations. Engineering Against Fatigue (Edited by J. H. Begon, M. W. Brown, R. A. Smith, T. C. Lindley and B. Tomkins), Balkema, Rotterdam, p. 211218. Siegmund, T., Kolednik, O. and Pippan, R. (1990) Direct measurement of the cyclic crack-tip deformation. Z. Metallkde. 81, 677683. Pippan, R. (2007) Effect of single overloads in ductile metals: a reconsideration. Eng. Fract. Mech. 74, 13441359. Newman, J. C. Jr. (1999) Advances in Fatigue Crack Closure Measurement and Analysis (Edited by R. C. McClung and J. C. Newman), ASTM STP 1343, West Conshohocken, PA, pp. 128144. Wanhill, R. J. H. (1975) Fractography of fatigue crack propagation in 2024-T3 and 7075-T6 aluminum alloys in air and vacuum. Met. Trans. 6A, 15871596. Bichler, C. (1997) Crack tip deformation at constant and variable amplitude. Thesis, Montanuniversit at Leoben.

2 3

4 5 6

7 8

9 10

11

12

13

14

15 16

17

18

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

16

R . P I P P A N et al.

19 20

21

22

23

24

25

26

27 28

29

30

31

32

33

34

35

36

37 38

Pippan, R. (1991) Dislocation emission and fatigue crack-growth threshold. Acta Mater. 39, 255262. Riemelmoser, F. O., Pippan, R. and Stuwe, H. P. (1997) A comparison of a discrete dislocation model and a continuous description of cyclic crack tip plasticity. Int. J. Fract. 85, 157168. Riemelmoser, F. O., Gumbsch, P. and Pippan, R. (2001) Dislocation modelling of fatigue cracks: an overview. Mater. Trans. 42, 213. Wilkinson, A. J., Roberts, S. G. and Hirsch, P. B. (1998) Modelling the threshold conditions for propagation of stage I fatigue cracks. Acta Metall. Mater. 46, 379390. Bjerk en, C. and Melin, S. (2003) A tool to model short crack fatigue growth using a discrete dislocation formulation. Int. J. Fatigue 25, 559566. Hanson, P. and Melin, S. (2005) Dislocation-based modelling of the growth of a microstructurally short crack by single shear due to fatigue loading. Int. J. Fatigue 27, 347356. Deshpande, V. S., Needleman, A. and Van Der Giessen, E. (2001) A discrete dislocation analysis of near-threshold fatigue crack growth. Acta Mater. 49, 31893203. Ohr, S. M. (1985) An electron microscope study of crack tip deformation and its impact on the dislocation theory of fracture. Mater. Sci. Eng. 72, 135. Rice, J. R. and Thomson, R. (1974) Ductile versus brittle behavior of crystals. Phil. Mag. 29, 7397. Pippan, R., Berger, M. and Stuwe, H. P. (1987) The influence of crack length on fatigue crack-growth in deep sharp notches. Metall. Trans. 18A, 429435. Pippan, R., Riemelmoser, F., Weinhandl, H. and Kreuzer, H. (2002) Plasticity-induced crack closure under plane-strain conditions in the near-threshold regime. Phil. Mag. A82, 32993309. Riemelmoser, F. O., Pippan, R. and Stuwe, H. P. (1998) An argument for a cycle-by-cycle propagation of fatigue cracks at small stress intensity ranges. Acta Mater. 46, 1793 1799. Pippan, R. and Riemelmoser, F. O. (1998) Visualization of the plasticity-induced crack closure under plane strain conditions. Eng. Fract. Mech. 60, 315322. Hansson, P. and Melin, S. (2008) Simulation of simplified zigzag crack paths emerging during fatigue crack growth. Eng. Fract. Mech. 75, 14001411. Murakami, Y., Kanezaki, T., Aline, Y. and Matsuoka, S. (2008) Hydrogen embrittlement mechanism in fatigue of austenitic stainless steels. Metall. Mater. Trans. 39A, 1327 1339. Lynch, S. P., Radtke, T. C., Wicks, B. J. and Byrnes, R. T. (1994) Fatigue-crack growth in nickel-based superalloys at 500700-degrees-C.1. waspaloy. Fatigue Fract. Eng. Mater. Struct. 17, 297311. Gach, E. F. (1999) Effect of environment on the crack tip deformation during cyclic loading. Diploma thesis, Montanuniversit at Leoben. Vehoff, H. (1980) Crack propagation and cleavage initiation in Fe-2.6%-Si single crystals under controlled plastic crack tip opening rate in various gaseous environments. Acta Mater. 28, 265272. Pelloux, R. M. N. (1969) Mechanisms of formation of ductile fatigue striations. Trans. Am. Soc. Metals 62, 281285. Tvergaard, J. V. (2004) On fatigue crack growth in ductile materials by crack-tip blunting. J. Mech. Phys. Solids 52, 21492166.

39

40

41

42

43 44

45 46 47

48 49 50

51

52

53

54

55

56

57

Pippan, R. (1991) Threshold and effective threshold of fatigue crack propagation in ARMCO iron II: the influence of environment. Mater. Sci. Eng. A138, 1522. Pippan, R., Tabernig, B., Gach, E. and Riemelmoser, F. O. (2002) Non-propagation conditions for fatigue cracks and fatigue in the very high-cycle regime. Fatigue Fract. Eng. Mater. Struct. 25, 805811. Mayer, H., Papakyriacou, M., Pippan, R. and Stanzl-Tschegg, S. (2001) Influence of loading frequency on the high cycle fatigue properties of AlZnMgCu1.5 aluminium alloy. Mater. Sci. Eng. A314, 4854. Pippan, R. and Weinhandl, H. (in press) Discrete dislocation modelling of near threshold fatigue crack propagation. Int. J. Fatigue. Suresh, S. and Ritchie, R. O. (1984) Propagation of short fatigue cracks. Int. Metals Rev. 29, 445477. Davidson, D., Chan, K., McClung and Mudak, S. (2003) Small fatigue cracks. Comprehensive Structural Integrity, Vol. 4 (Edited by R. O. Ritchie and Y. Murakami), Elsevier, pp. 129164, Oxford, UK. Suresh, S. (1998) Fatigue of Materials. Cambridge University Press, Cambridge, UK. Zelger, C. (2007) In-situ Untersuchung des Wachstums kurzer Risse. Diplomarbeit, Montanuniversit at Leoben. Rice, J. (1967) Mechanics of Crack Tip Deformation and Extension by Fatigue. ASTM STP 415, Philadelphia, pp. 247309. Manson, S. S. (1954) National Advisory Commission on Aeronautics: Report 1170, Cleveland, OH. Coffin, L. F. (1954) Trans. Am. Soc. Mech. Eng. 76, 931951. Dowling, N. E. (1977) Crack growth during low-cycle fatigue of smooth axial specimens. In: Cyclic Stress-strain and Plastic Deformation Aspects of Fatigue Crack Growth. Special Technical Publication 637, pp. 97121. Philadelphia: American Society for Testing and Materials. Krupp, U., Floer, W., Lei, J., Christ, H., Schick, H. J. and Fritzen, C. P. (2002) Mechanisms of short-fatigue-crack initiation and propagation in a beta-Ti alloy. Phil. Mag. A82, 33213332. Pippan, R., Hageneder, P., Knabl, W., Clemens, H., Hebesberger, T. and Tabernig, B. (2001) Fatigue threshold and crack propagation in gamma-TiAl sheets. Intermetallics 9, 8996. Pippan, R., Tesch, A., Hageneder, P., Hebesberger, T., Kestler, H. and Clemens, H. (2001) Fatigue crack propagation in -TiAl sheets: intrinsic and extrinsic contributions to the fatigue resistance. Proc. of International Symposium on Structural Intermetallics 2001 TMS (Edited by K. J. Hemker, D. M. Dimiduk, H. Clemens, D. Darolia, H. Inui, J. M. Larsen, V. K. Sikka, M. Thomas and J. D. Whittenberger), pp. 381389. Pippan, R., Bichler, C., Tabernig, B. and Weinhandl, H. (2005) Overloads in ductile and brittle materials. Fatigue Fract. Eng. Mater. Struct. 28, 971981. Tanaka, K., Hoshide, T., Yamada, A. and Taira, S. (1979) Fatigue crack-propagation in blaxilal stress-fields. Fatigue Eng. Mater. Struct. 2, 181194. Tatschl, A. and Kolednik, O. (2003) A new tool for the experimental characterization of micro-plasticity. Mater. Sci. Eng. A339, 265280. Motz, C. and Pippan, R. (2001) Deformation behaviour of closed-cell aluminium foams in tension. Acta Mater. 49, 24632470.

c 2010 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 34, 116

S-ar putea să vă placă și