You are on page 1of 22

Engineering Fracture Mechanics 198 (2018) 2–23

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Fatigue strength and fracture mechanics – A general perspective


U. Zerbst a,⇑, M. Madia a, M. Vormwald b, H.Th. Beier b
a
Bundesanstalt für Materialforschung und -prüfung (BAM), 9.1, D-12205 Berlin, Germany
b
Technische Universität Darmstadt, Materials Mechanics Group, D-64287 Darmstadt, Germany

a r t i c l e i n f o a b s t r a c t

Article history: Common fracture mechanics based fatigue considerations are usually limited to the resid-
Received 22 March 2017 ual lifetime determination of so-called long cracks. The extension of this concept to the
Received in revised form 13 April 2017 total lifetime, as in the S-N curve approach, requires an adequate description of short crack
Accepted 18 April 2017
propagation which cannot be based on the DK concept, and it must consider the crack clo-
Available online 20 April 2017
sure phenomenon as well as its gradual build-up at the short crack stage. Further, it has to
provide a meaningful definition of initial crack dimensions and a solution for the multiple
Keywords:
crack problem at stress levels higher than the fatigue limit as it is specific for some config-
Fatigue strength
Endurance limit
urations such as weldments. This paper aims at a discussion of all these points and offers
Fracture mechanics possible solutions which are illustrated by examples taken from the German IBESS project
Short crack propagation on fracture mechanics based determination of the fatigue strength of weldments, the
Multiple cracking results of which will be discussed in more detail in this Special issue.
Weldments Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction

For constant amplitude loading, the fatigue properties of materials (and structures) are usually described by S-N or e-N
curves (Wöhler curves). It has been found that some materials such as ferrous alloys show a distinct lower limit in the S-N
curve in the range beyond 106–107 loading cycles. This is designated by the term ‘‘fatigue limit” or ‘‘endurance limit”. If a
specimen or component is loaded at stress levels below this limit no failure will occur no matter how many cycles it expe-
riences. This concept has been challenged at the background of very high cycle fatigue (108 or more loading cycles) where a
further drop of the S-N curve was found even for materials for which a distinct fatigue limit in the high cycle fatigue regime
(up to 107 cycles) was attested in the past. This is state-of-the-art today (see, e.g., [1]) although it is still not clear if all mate-
rials are concerned in the same way [2]. No detailed discussion on this topic shall be provided in this paper. When the term
‘‘fatigue limit” is used here, it characterizes the stress range or amplitude at 107 loading cycles.
Whilst the material S-N curve is usually determined on unnotched specimens and a stress ratio R = rmin/rmax = 1, the
component S-N curve, besides the material influence, is affected by a number of factors such as mean stress or stress ratio,
notches, surface roughness and impairments. In practical application those effects are taken into account by semi-empirical
correction factors on the fatigue limit.
In contrast to the S-N curve concept, fracture mechanics has the potential for implicitly taking into account such factors.
However, in common applications e.g. in the framework of a damage tolerance concept [3], it is restricted to the determi-
nation of a residual lifetime, i.e., the time that a pre-existing crack need to grow to its critical size. The dimensions of the
pre-existing cracks are usually defined by the detection limit of the non-destructive testing method applied in quality

⇑ Corresponding author.
E-mail address: uwe.zerbst@bam.de (U. Zerbst).

http://dx.doi.org/10.1016/j.engfracmech.2017.04.030
0013-7944/Ó 2017 Elsevier Ltd. All rights reserved.
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 3

Nomenclature

a crack length (crack depth for surface cracks)


ap primary pit depth (Fig. 22)
as secondary pit depth (Fig. 22)
ai initial crack depth (for fracture mechanics analysis), Fig. 14
anotch depth of a secondary notch, e.g., an undercut
a0 El Haddad parameter in Eq. (1) and Fig. 2
a⁄ correction term for modified El Haddad’s model, Eqs. (8) and (9)
b Burges vector, Eq. (3)
c half crack length at surface (semi-elliptical crack)
2cp primary pit width (Fig. 22)
2cs secondary pit width (Fig. 22)
C, n fit parameters of the da/dN-DK curve in the Paris regime, Fig. 4
d1 crack depth defining crack arrest at fatigue limit, Figs. 2, 9 and 10
d2 crack depth for the transition between short and long cracks, Figs. 2 and 9
da/dN fatigue crack propagation rate
E modulus of elasticity
K stress intensity factor (K-factor)
Kmax maximum K-factor in a loading cycle (Fig. 5)
Kmin minimum K-factor in a loading cycle (Fig. 5)
Kop K-factor at crack opening (Fig. 5)
Kt stress concentration factor
N number of loading cycles
R loading ratio (=rmin/rmax or Kmin/Kmax)
Re yield strength
Rm tensile strength
Ti component of the traction vector for determining the J integral (Eq. (5))
ui displacement vector components for determining the J integral (Eq. (5))
U crack closure factor (DKeff/DK)
U(a) crack closure factor for mechanically short cracks, Fig. 6
ULC crack closure factor for long cracks, independent of crack length, Fig. 6
We strain energy density (Eqs. (5) and (6))
a0 , n0 fit parameters of the Ramberg-Osgood description of the cyclic stress-strain curve (Eq. (4))
Da crack extension, Figs. 6, 7, 8 and 10
DCTOD cyclic crack tip opening displacement
DJ cyclic J integral
DK K-factor range (Kmax  Kmin)
DKeff effective K-factor range (=Kmax  Kop)
DKp plasticity-corrected DK, Eq. (10)
DKth fatigue crack propagation threshold
DKth,eff intrinsic fatigue propagation threshold, Fig. 8a
DKth,LC fatigue propagation threshold in the long crack regime, Fig. 8a
DKth,op DKth component induced by the crack closure effect, Fig. 8a
De strain range (emax  emin)
Dr stress range (rmax  rmin)
Dre fatigue limit range
Drth threshold stress range, Eq. (1)
e strain
emax maximum strain in the fatigue cycle
emin minimum strain in the fatigue cycle
C path for J integral determination (Eq. (5))
m Poisson’s ratio
r stress
r0,cyc reference stress, Eq. (4)
ra stress amplitude (=1/2Dr)
re fatigue limit (stress amplitude)
rw re at R = 1
rmax maximum stress in the fatigue cycle
rmin
p ffiffiffiffiffiffiffiffiffiffi minimum stress in the fatigue cycle
area square root of the crack area
4 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Abbreviations
cyc cyclic
FCP fatigue crack propagation
KT Kitagawa-Takahashi diagram
LC long crack

control after manufacturing or in regular inspections in service. It will, therefore, depending on the material, be in the order
of millimeters which – in the following – will be referred to as long cracks (the term will be explained in Section 3). Any frac-
ture mechanics approach that potentially could be applied to fatigue strength determination has to address at least there
challenges: (a) It has to adequately describe so-called short crack propagation (which cannot be based on the common long
crack concepts such as the da/dN-DK curve for principle reasons), (b) it has to provide a meaningful definition of the initial
crack dimensions as the starting point for an S-N curve relevant (residual) lifetime analysis, and (c) it has to cope with the
problem of multiple cracks for load levels higher than the fatigue limit for configurations where this is relevant. It is the aim
of the present paper to provide an introduction to these problems which also forms the basis of the fracture mechanics
approach for determining the S-N curve of weldments as the topic of this Special Issue.

2. Fatigue strength

2.1. The fatigue limit

Yukitaka Murakami, in his book ‘‘Metal Fatigue” [4], answers the question ‘‘What is a fatigue limit?” this way: ‘‘A fatigue
limit is the threshold stress for crack propagation and not the critical stress for crack initiation” as it was thought in former
times. Likewise, Keith Miller, in a distinct paper [5] demonstrated that the actual scenario at the background of the fatigue
limit is not that of ‘‘not forming” cracks but of arresting them. In engineering alloys a large number of microcracks is initiated
at various microstructural features such as slip bands, e.g., at grain or twin boundaries or at inclusions. In the following, these
microcracks shall be designated as microstructurally short cracks because their early propagation is strongly affected by the
microstructure of the material (see also Section 3). One consequence is that the crack experiences phases of acceleration,
deceleration and arrest. The conventional fatigue limit can be defined as that stress range Dre below which even the largest
of an extended number of originally propagating microstructurally short cracks is arrested. Note, however, that crack arrest
can also occur at a later stage of crack propagation when the crack is physically short or even at the transition from the short
to the long crack range (Section 3). An example of an arrested crack is shown in Fig. 1.

2.2. The Kitagawa-Takahashi diagram (KT diagram)

An early method which combines S-N based fatigue strength and fracture mechanics in a semi-empirical way is the
Kitagawa-Takahashi diagram [7] which is a log-log plot of the threshold stress range Drth against the crack length a. It is
shown in Fig. 2(a) in its original and Fig. 2(b) in the version proposed by Miller [5] which takes into consideration crack prop-
agation below the fatigue limit Drth = Dre such as mentioned above.
The KT diagram may be subdivided into three regions. For long cracks (right hand side) the diagram shows a slope of 1/2
in double logarithmic scaling. Here, the da/dN-DK concept is applicable. For very small cracks (microstructurally short; left
hand side) the threshold stress range Drth refers to the fatigue limit of the S-N curve approach which, as mentioned above,

Fig. 1. Non-propagating crack in A1N railway axle steel. The crack stopped within a grain. It experienced more than 108 loading cycles; according to [6].
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 5

Fig. 2. Kitagawa-Takahashi diagram; (a) original version [7]; (b) version with respect to the growth characteristics of microstructurally short cracks
according to Miller [5].

is defined by crack arrest. In between these two regions there is a range which can be designated as that of mechanically/-
physically short cracks. These terms will be explained in more detail in Section 3. Here it shall only be mentioned that this
intermediate range can be described by fracture mechanics, but not by the linear elastic DK concept. Note, however, that
the KT diagram does not use elastic-plastic fracture mechanics for describing this branch of the curve but, instead, applies
a semi-empirical equation proposed by El Haddad et al. [8]:
 1=2
Drth ðaÞ a0
¼ ð1Þ
Drth ða ! 0Þ ða þ a0 Þ

with a0 referring to the intersection point of extended straight lines left and right in the double-logarithmic plot. The black
dot at a = d1 refers to the crack arrest event defining the fatigue strength.

3. Stages of fatigue crack propagation and lifetime

Generally, the lifetime of cyclically loaded structures made of technical alloys consists of three periods: fatigue crack ini-
tiation, fatigue crack propagation and final fracture. Frequently and for pragmatic reasons, the crack initiation period is
defined as the time or number of loading cycles up to the first detectable crack. However, with respect to the physical dam-
age process this ‘‘engineering” initiation period can be subdivided into three distinct stages. Consequently, the complete life-
cycle of a fatigue crack, from its nucleation to component failure, passes through the stages shown in Fig. 3.

3.1. Crack nucleation

Crack initiation (or nucleation) in a narrower sense is the range within which a crack first develops due to the accumu-
lation of irreversible plastic deformation. This can happen at pre-existing defects which act as stress concentrators due to
stiffness mismatch with the matrix material and/or micro notch effects or – in rare cases – at the defect free surface. In
the very high cycle fatigue regime the fatigue initiation site is found beneath the surface but also in conjunction with some
micro-structural irregularities [4]. The definition of the end of the initiation phase is a bit academic because it is hard to
define when exactly an original defect has become a crack.
6 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Fig. 3. Crack length scales of the life cycle of a component subjected to cyclic loading. (a) According to [11]; slightly modified, numbers are rough estimates
only; (b) crack propagation stages including the thresholds of fatigue; according to [5].

3.2. Microstructurally short crack propagation

Subsequent to its nucleation, the crack is still short compared to the characteristic dimensions of the microstructure,
mainly the grain size. Therefore, it is designated as microstructurally short. Its propagation is discontinuous, i.e., characterised
by acceleration and deceleration phases, e.g., when the crack has to overcome microstructural barriers, see Fig. 2(b). Com-
mon barriers are grain or phase boundaries where the neighboring units show different crystal orientations which affect the
development of the plastic zone ahead of the crack tip. Not all neighboring grains are of significantly different crystal orien-
tation because of which the effect is not uniform along the crack front when this becomes larger. As a result, the maximum
size of non-propagating cracks at the fatigue limit is usually larger than one grain size (e.g., Murakami reports it as three
average grain sizes for an annealed 0.13% C steel [4]). If the applied stress range is high enough, the crack, irrespective of
its size, will continue to extend.

3.3. Mechanically/physically short crack propagation

At the next stage, when the crack front length exceeds a certain value and encloses, depending on the material, a larger
number of grains, the crack will become a mechanically short one. The influence of the local microstructural features is declin-
ing – or, better, it is averaged out – and the crack propagation becomes more continuous. The crack is still short but its
dimensions are now in the order of characteristic mechanical dimensions such as the size of the plastic zone ahead of the
crack tip. Since the crack might be fully embedded within the plastic zone, fracture mechanics concepts based on the cyclic
linear elastic stress intensity factor DK cannot be applied but have to be replaced by elastic-plastic concepts such as DJ or
DCTOD.
Like the microstructurally short crack, the mechanically short one can be arrested (e.g., due to the stress gradient in a notch
[9]) but now a further aspect comes into play: the gradual build-up of crack closure effects, which presupposes a certain
crack size. This aspect is usually characterized by the term physically short crack. The crack closure will be addressed in
Section 4.1.

3.4. Long crack propagation

Eventually a crack which has not been arrested before becomes a long one. The transition is defined by the fully developed
crack closure effects, i.e., the stress or K-factor in the loading cycle above which the crack is open has reached a constant
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 7

value. At that stage crack propagation can be described by the da/dN-DK diagram, however, corrected for the crack closure
effects, until the maximum load in terms of the stress intensity factor, Kmax, is so high that it causes interspersed events of
monotonic crack extension.

3.5. Component failure

Failure of the component occurs when the crack reaches its critical state. This can be defined by Kmax (or its elastic-plastic
equivalent) exceeding the monotonic fracture resistance of the material but also by other characteristics which cause the
loss of function of the component. Note that the exact definition of the failure criterion, in many cases, is of minor relevance
since the crack propagation is very rapid at this stage and there is no much difference in terms of residual lifetime between
different definitions.
Comparing the different stages of crack propagation, it is frequently found that neither crack nucleation nor long crack
propagation but the short crack regime decisively contributes to the overall lifetime of the component [10].

4. Application of fracture mechanics to long and short crack propagation

4.1. Long crack propagation

4.1.1. The da/dN-DK diagram


The backbone of fracture mechanics description of long crack propagation is the da/dN-DK diagram. No extended expla-
nation is needed here. The double-logarithmic diagram consists of three regions: the threshold region, the so-called Paris
region and the region were fatigue propagation moves towards fracture. The main influencing factors within the regions
are given in Fig. 4. Perhaps the most astonishing observation is that the curve, in the Paris region, is only slightly influenced
by the microstructure of the material within whole material classes such as steels or aluminium alloys.

4.1.2. Crack closure effects


An information that cannot be gained from Fig. 4 is that the da/dN-DK curve is affected by the R-ratio which is an indirect
measure of the mean stress in the loading cycle. With increasing R-ratio (and mean stress) the curve, for a given DK, is
shifted to higher crack propagation rates da/dN. This is commonly explained by the approach of crack closure first intro-
duced by Elber [12]. In practical analysis, it is taken into account by replacing the cyclic crack driving force DK = Kmax  Kmin
by DKeff = Kmax  Kop with Kop being the K-factor above which the crack is open.
The crack closure effect is caused by a number of mechanisms () which shall be briefly discussed in the following. Basi-
cally, crack propagation only occurs in that part of the loading cycle during which the crack is open. For a stress ratio of
R = 1 that means that DKeff can be no larger than DKeff = 1/2 DK. However, in reality, DKeff is even smaller due to further
effects. The most important ones are the plasticity-induced, the roughness-induced and the oxide-induced ones, for the discus-
sion of further variants such as closure due to viscous liquids, martensitic transformations, metal particles from fretting or
hydrogen-induced deformation see [11,13,14]. The first three mechanisms shall be very briefly explained, for a more detailed
discussion see the citations above, [15] and also the review on the fatigue crack propagation threshold of the authors in [16].

Fig. 4. The three regions of the long crack da/dN-DK diagram for describing fatigue propagation.
8 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Fig. 5. Crack closure mechanism. (a) Definition of the loading parameters; (b) plasticity-induced mechanism; (c) roughness-induced mechanism; (d) oxide-
induced mechanism.

4.1.2.1. The plasticity-induced crack closure effect (Fig. 5b). When the fatigue crack propagates, i.e., the crack tip is shifted fur-
ther into the ligament, the plastically stretched material, i.e. the plastic zone formed ahead of the crack tip, remains at the
crack faces in the wake of the crack. There it forms a ‘‘thickening” particularly near the side surfaces where a plane stress
state prevails. This, of course, affects crack closure in that it increases the load above which the crack is open. The
plasticity-induced crack closure effect plays a role in the threshold as well as the Paris regions of the da/dN-DK diagram. Like
the other closure effects, discussed below, it increases with lower R-ratios and diminishes and finally disappears at high R.

4.1.2.2. The roughness-induced crack closure effect (Fig. 5c). The roughness-induced crack closure effect is caused by the asper-
ity of the crack faces and enhanced by a shear mode component of the displacement due to potential deflection and branch-
ing of the crack. Roughness-induced crack closure plays a major role near the fatigue crack propagation threshold and in
mixed-mode loading situations.

4.1.2.3. The oxide-induced crack closure effect (Fig. 5d). Comparable to the roughness-induced crack closure phenomenon, the
oxide- or oxide-debris-induced mechanism plays a major role near the fatigue crack propagation threshold for surface
cracks. In moist atmospheres oxide layers are built-up at the crack surfaces. At low R-ratios, the thickness of this might fur-
ther increase by breaking and re-forming the oxide layer due to friction, i.e., debris is formed within the crack. Note that
there is some indication [17] that oxide-induced crack closure disappears under LCF conditions, since the oxide-debris is
pressed into the crack wake during the compressive half cycle.

4.1.3. Long crack threshold against fatigue crack propagation


In the context of fracture mechanics-based S-N curve determination, the threshold against fatigue propagation, DKth, is of
paramount interest. In a wider frame it is the last threshold for crack arrest after the crack has passed the other barriers men-
tioned in Section 3, and it is closely related to the crack closure effects described above. For a detailed review on DKth the
reader is referred to the review paper of the authors in [16].
The principle is quite easy: At loads DKeff below DKth the crack is not able to propagate. Thus, the parameter could be
understood as an endurance limit if it were not referred to long cracks in this section. These cracks have already experienced
nucleation and short crack propagation, this way having ‘‘used up” the major part of their lifetimes. Note that short cracks
grow at nominal load levels at which long cracks would arrest. The reasons for this will be explained in Section 4.2.
An important observation is that the threshold DKth consists of two components, an intrinsic one, DKth,eff, and a second
one, DKth,op, caused by the crack closure effects:

DKth ¼ DK th;eff þ DKth;op ð2Þ


The intrinsic component DKth,eff is identical to the closure-free DKth, e.g. determined at a high R-ratio, typically at R  0.7.
Therefore, it is designated with the index ‘‘eff”. It has long been recognized (see the overview in [16]) that it is correlated
with the modulus of elasticity, E, which in turn is a function of the crystal lattice of the material, and by the Burger’s vector,
b, as a geometrical measure of dislocations. A more recent correlation, proposed in [18], is given by:
qffiffiffiffiffiffi
3
DKth;eff ¼ E  j bj ð3Þ
4
Typical values of DKth,eff are 2.4–2.6 MPa m1/2 for steels, 0.9–1.9 MPa m1/2 for aluminium alloys and 1.7–2.5 MPa m1/2 for
titanium alloys. For a more extended overview see [19]. It is of major relevance that the intrinsic threshold is obviously inde-
pendent of material parameters related to the microstructure as well as of the R-ratio and the crack length. This is com-
pletely different with respect to the crack closure part DKth,op which is affected by all these parameters but always in the
context of the crack closure effects, i.e., in a different way than other material parameters such as the strength [16].
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 9

4.2. Short crack propagation

Short cracks are different from long cracks in that their propagation cannot be described by the linear elastic fracture
mechanics, and in that the crack closure effects are not yet fully developed but show a transient behavior depending on
the crack length. This shall be discussed in the following. Note that this discussion is restricted to mechanically and physically
short cracks. Micromechanically short cracks cannot be described by common fracture mechanics approaches. Instead con-
cepts designated by Miller as ‘‘micro-mechanical fracture mechanics” have to be employed (see, e.g. [20]).

4.2.1. Elastic-plastic crack driving force for cyclic loading


The linear elastic DK concept cannot be applied to short cracks since the plastic zone size is in the same order of magnitude
as the crack length or – in other words – the crack is even embedded in the plastic zone. As an alternative, elastic-plastic con-
cepts such as the cyclic J integral DJ or a cyclic crack tip opening displacement DCTOD can be used. Various authors, including
one of this paper, have shown that the da/dN-crack driving force curves of short and long cracks coincide on the basis of DJ, if
the latter is properly defined and the crack closure effect is adequately considered [21–24]. The determination of DJ will be the
topic of another paper in this Special issue [25]. Therefore, only some basic ideas will be given here.
It is well known that the J integral has been developed for non-linear elastic deformation behaviour because of which its
application is limited to proportional and monotonically increasing loading. These conditions are not fulfilled in the case of a
growing crack. Why then the application of DJ to fatigue crack propagation should be possible? Saxena, in [26], argues, based
on earlier work of Lamba [27], that a path-independent DJ integral is obtained when it is defined for the increasing (or
decreasing [24,25]) load portion of the (cyclic) load deflection curve only. This can be described by a Ramberg-Osgood like
equation
 n0
Dr Dr
De ¼ þ 2  a0  ð4Þ
Ecyc 2  r0;cyc
where the factor 2 refers to the loading path from the lower to the upper reversal points of the hysteresis loop. In this case
the DJ integral is formulated as
Z  
@ ðDui Þ
DJ ¼ DWe  dy  DTi   ds ð5Þ
C @x
with
Z
 

DWe ¼ Drij  d Deij ð6Þ

Note that, unlike DK = Kmax  Kmin, the parameter DJ – Jmax  Jmin. The ‘‘D‘‘ terms refer to the arguments in the equations,
i.e., the load input parameters. The requirement for a limited amount of crack extension (typical in monotonic R-curve deter-
mination based on J) is also fulfilled since the crack increment during one loading cycle is small (ca. 0.1 mm maximum). The
path independency of the DJ integral defined this way has empirically been proven (e.g., [22,27,28]) and its application is
common today (see the extended overview in [29]). There are, however, also limitations [26], e.g., DJ may be path-
dependent when the condition of a stabilized cyclic stress-strain curve is not fulfilled.
Within the application in this Special issue an analytical approach was developed for the analytical determination of DJ,
see [25], which is based on an R6-type methodology [30], see also [31]. This is slightly different but similar to an approach
proposed by McClung et al. [32]. An important input parameter for determining the ligament yielding correction, which
forms the centrepiece of such methods, is the stabilised cyclic stress strain curve of the material (e.g, the heat affected zone
in weldments) in which short crack extension takes place [33].

4.2.2. Crack length dependency of the crack closure effects and the threshold Dkth
When the crack closure effect is defined by a dimensionless parameter U (=DKeff/DK), its development with increasing
crack length is illustrated in Fig. 6. At the beginning, when the crack is very short, e.g. it is just formed by decohesion of a
non-metallic particle from the matrix material, no crack closure effect exists and U = 1. In order to build-up the effects, some
loading history is necessary, and that means a certain dimension of the wake of the propagating crack. This forms during
crack propagation when the plastically stretched zone within the wake (or other features like oxide debris) will accumulate.
The parameter U gradually decreases to U < 1 until it reaches a crack length-independent constant value ULC. The transition
to ULC defines the transition from the short to the long crack propagation regimes. Note, that only surface cracks are consid-
ered here.
It is essential for the fracture mechanics model for generating S-N curves presented within this Special issue that the
development of the crack closure effect in terms of U is mirrored in the crack extension (Da) dependency of the fatigue crack
propagation threshold [34] (Fig. 7).
UðaÞ  1 DKth ðaÞ  DKth;eff
¼ ð7Þ
ULC  1 DKth;LC  DKth;eff
10 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Fig. 6. Schematic crack size dependency of the dimensionless closure parameter U.

Fig. 7. Parallel development of the closure factor U and the fatigue crack propagation threshold with increasing crack length, schematic view.

Fig. 8. Cyclic R-curve; (a) intrinsic and crack opening component, DKth,eff and DKth,op, of the fatigue crack propagation threshold DKth and (b) cyclic R-
curves obtained for S355NL steel (base metal); according to [34], see also [35] within this Special issue.

The curve at the right hand side of Fig. 7 is designated as ‘‘cyclic R-curve” and is separately illustrated in Fig. 8 along with
an example of one of the steels investigated within the IBESS cluster project. No detailed information on the cyclic R-curve
and its determination will be added here since this is the topic of another paper within this Special Issue [35]. The only point
to be mentioned is that it is the crack closure component DKth,op only, which increases during crack propagation whilst the
intrinsic value, DKth,eff, stays constant such as described above.
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 11

5. Bringing together fatigue strength and fracture mechanics

5.1. Kitagawa-Takahashi diagram and cyclic R-curve

The cyclic R-curve and the Kitagawa-Takahashi diagram according to [8] are complementary approaches (Fig. 9) when Eq.
(1) is slightly modified by introducing an additional term a⁄ [34,36] such that the threshold is described by
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Da þ a 
DKth ðDaÞ ¼ DKth;LC  ð8Þ
Da þ a þ a0

with a⁄ simply being determined by


 2
DKth;eff =DKth;LC
a ¼ a0   2 ð9Þ
1  DKth;eff =DKth;LC

The term a⁄ is introduced to fulfill the condition that DK th ¼ DK th;eff for Da ¼ 0 which is not part of the original El Haddad
model [8] to the Kitagawa Takahashi diagram. Whilst the latter is a semi-empirical approach, the cyclic R-curve provides
some physical background.

5.2. Cyclic R-curve analysis

The relation between fatigue strength and fracture mechanics is more generally illustrated in Fig. 10 which comprises (a)
the S-N curve, (b) the fatigue propagation threshold Kth including (c) its crack length dependency in the physically short crack
range, (d) the Kitagawa-Takahashi diagram and (e) the principle of a cyclic R-curve analysis which finally ties together the
different concepts. What they all have in common is that they define the fatigue strength by crack arrest. This is illustrated
by the dark dot, which shows up in Fig. 10(a), (d) and (e). In principle, this point could be obtained from the Kitagawa-
Takahashi diagram. However, its determination is not straightforward there because there is no distinct transition point
of the curve at d1.
A realistic alternative is a cyclic R-curve analysis. The principle is illustrated in Fig. 11. It can be compared to the well-
known R-curve analysis for monotonic loading which also consists of three elements. The elements of the cyclic analysis are:

(a) The crack driving force DKp determined for different applied loads Dr and different crack sizes. Within the model
introduced in this Special issue, DKp is formally derived from DJ by

Fig. 9. Correspondence between the Kitagawa-Takahashi diagram (left) and the cyclic R-curve (right). Sections A, B and C refer to the crack growth stages of
microstructurally short (A), mechanically short (B) and long cracks (C).
12 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Fig. 10. The relation between fatigue strength and fracture mechanics. (a) S-N curve, (b) fatigue propagation threshold DKth including (c) its crack length
dependency in the physically short crack range, (d) the Kitagawa-Takahashi diagram and (e) the principle of a cyclic R-curve analysis.

qffiffiffiffiffiffiffiffiffiffiffiffiffi
DKp ¼ DJ  E 0 ð10Þ

with E0 being E for plane stress and E/(1  m2) for plane strain conditions. The assumed surface crack geometry is semi-
elliptical for finite-life fatigue strength but semi-circular at the stress level referring to the endurance limit where only
one relevant crack exists (see Section 5.3).
(b) The second element of the fracture mechanics triangle is the cyclic R-curve as described above, which has
(c) its starting point in Fig. 11 at (ai; DKth,eff). The crack depth ai marks the size of a pre-existing crack which is needed for
fracture mechanics analysis.

If ai is chosen such that the crack driving force and the R-curve just fulfill the tangency criterion, it refers to the size of the
crack, which is just at the arrest-non arrest transition or, in other words, it is the largest non-propagating crack.

5.3. The fatigue strength

At this point it has to be distinguished between the finite life fatigue strength and the endurance limit respectively the
fatigue limit for N = 107 as defined within the present approach. Let’s begin with the latter. When the endurance limit is
defined as that stress range Dre or stress amplitude re below which even the largest of an extended number of cracks
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 13

Fig. 11. Schematic view of a cyclic R-curve analysis. The transition from arrest to non-arrest is given by that crack driving force curve (referring to load 2 in
the example) which tangentially touches the cyclic R-curve.

(which were previously able to grow) is arrested, the assumption of just one relevant crack at the transition from arrest to
non-arrest makes sense. That this is also the case in reality is illustrated in Fig. 12, where S-N tests of steel butt weldments
were interrupted after about 1/4 to 1/3 of the overall lifetime [37]. Subsequently the number of cracks (all cracks and only
those deeper than 50 lm) along the weld toe was determined. It showed up that the number of cracks increased with the
load level up to more than hundred (when only the 50 lm deep ones are counted) at stress amplitudes of 200 MPa. However,
at a stress amplitude of about 100 MPa which, in the experiments, roughly referred to the fatigue limit at N = 107, the count-
ing pointed to the order of zero to one crack. The weld toe length considered here was 50 mm.
In the frame of the present approach, the R-curve analysis was applied to the determination of the fatigue limits of weld-
ments [39]. In a first step, the initial crack size of a semi-circular surface crack in a tensile loaded flat and smooth plate was
determined. For this, the applied loading curve DKp-a was established for stepwise increased crack sizes, a, and a stress
amplitude ra referring to the endurance limit re of the material (which, as was shown in [33], could also be estimated from
the ultimate tensile strength Rm of the material). Then, with its starting point at the ordinate fixed to DKth,eff the known cyc-
lic R-curve was shifted along the abscissa such that the crack driving force and cyclic R-curve just touched one each other at
one point (tangency). The correspondent starting point of the cyclic R-curve at the level of DKth,eff was the initial crack size ai.
The calculations were performed for R = 1 and the resulting initial crack size ai was assumed to be a material dependent
parameter.

Fig. 12. Experimentally determined number of cracks along the 50 mm weld toe of butt welds, obtained at 1/4 to 1/3 of the overall lifetime [37], see also
[38].
14 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Fig. 13. The division of the weld toe into equidistant sections. Each of it is characterized by its own weld geometry and stress-depth profile. In addition,
each section contains an initial crack the size of which also is determined statistically.

With the cyclic R-curve, the now known initial crack size ai and the intrinsic threshold DKth,eff it was possible to deter-
mine the component fatigue limits. For this, the cyclic R-curve was fixed on both the abscissa (at a = ai) and the ordinate (at
DKth = DKth,eff) and crack driving force curves DKp(a) of the welded components were determined for different applied loads,
which were determined for the through thickness stress profiles of the real weld geometries including the toe notches The
load for which the tangent criterion with the cyclic R-curve was satisfied defined the fatigue limit of the weldment under
consideration. Note that all analyses were performed in a stochastic way taking into account the variability of the initial
crack size and of the geometry parameters at the weld toe. An example with respect to the initial crack size [39] will be pro-
vided in Section 6.2.1.

5.4. Multiple crack propagation in the finite life regime

The determination of the fatigue strength in the finite life regime was further complicated compared to those of the
endurance limit because of the phenomenon of multiple crack propagation, cf. Fig. 12. The reason was that not only the ini-
tial crack size was a statistical parameter but, in addition, the local weld toe geometry was not homogeneous but showed
some variation along the toe [37]. The principle, how the analysis was performed, is introduced in more detail in [39]
and shall only be briefly highlighted here. The weld toe was subdivided into equidistant sections each of which characterized
by an individual geometry (Fig. 13) with the latter being provided in terms of empirically obtained statistical distributions of
the weld toe radius, the flank angle and the secondary notch depths. Note that Fig. 13, as a general scheme, also includes
excess weld metal and material parameters as statistical input data, which was not realized in the IBESS approach so far.
The through thickness stress distributions were determined by parametric equations developed within the IBESS cluster pro-
ject (see [39]). At the beginning of the analysis, the crack growth was determined individually for those cracks which were
able to propagate but beyond a certain crack size, interaction effects between adjacent cracks had to be taken into account.
Crack coalescence was simulated with the result that the number of cracks stepwise decreased such as did the a/c ratio until
one long surface crack along the weld toe was the result. This pattern is confirmed by empirical evidence within the own
project and the literature on weldment fatigue (for a review see [40]).

6. The initial crack size for fracture mechanics analyses

6.1. General aspects

Within the approach presented in this Special issue the initial crack depth ai was defined as that depth of a semi-circular
crack that was just at the limit between arrest and non-arrest at the stress level of the material’s endurance limit. Note, how-
ever, that this approach cannot be generalized. E.g., in [41], investigating an aluminium alloy, the authors have identified
inclusions as crack initiation sites. It showed up that the cohesion of these inclusions to the matrix material was very poor.
They were rather large (with a maximum depth of about 50 lm) and this size could immediately be taken as the initial crack
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 15

Fig. 14. Definition of the initial crack size for fracture mechanics analyses by either crack arrest or initial flaws, whichever provides larger values.

length. As the consequence, a differentiated view on defining initial cracks is due. The general scheme is presented in Fig. 14.
The initial crack size determined for crack arrest provides a lower limit which is relevant only if no initial flaws larger than
this exist, provided these flaws can be treated as initial cracks. With respect to such flaws, it can be distinguished between
material defects and geometric irregularities. In the following, a brief discussion on this topic will be provided.

6.2. Material defects

6.2.1. Non-metallic inclusions


Non-metallic inclusions are typical initial flaws, e.g., in high strength steels and wrought aluminium alloys, for overviews
see [4,42,43]. Whilst the inclusion size is reported as the dominant parameter other characteristics such as its shape and ori-
entation with respect to the loading direction, its distance from the surface, its adhesion to the matrix material and stiffness
mismatch (e.g., different E moduli) between inclusion and matrix can also play a role (see the discussion in [44] at the back-
ground of the failure of railway axles for which inclusions have been identified as a potential root cause). Inclusions may also
play a role in environmentally-assisted crack initiation and early growth due to the electrochemical mismatch between
inclusion and matrix [45].
In principle, indigenous and exogenous inclusions can be distinguished with the first ones being the result of reactions
such as desoxidation in molten or solidifying steel and the second ones being the result of mechanical incorporation of slags
from the mold powder or particles from the refractory lining [44,46]. Besides individual inclusions, inclusion clusters or
band-type inclusion formations can occur which are usually more dangerous because of their larger dimension. A thorough
discussion of the various inclusion types and their effects in calcium treated steels is provided in [46]. The effect of inclusions
on the S-N curve of spring steel is shown in Fig. 15 [47].
Fig. 16 shows initial crack sizes, ai, and crack sizes at crack arrest of two steels of different strengths obtained by the crack
arrest approach described in Section 4.2. As can be seen, the predicted crack sizes are significantly smaller for the high
strength than for the low strength steel. This observation, in conjunction with Fig. 14, might explain why the fatigue limit
of high strength steels tends to be influenced by inclusions whereas that of low strength steels doesn’t: Whilst the crack size
at arrest becomes smaller with higher strength, the inclusion size does not follow this trend. Thus one should expect the
inclusions taking over the role of the initial defect size when the strength of the material is high enough. On the other hand,
larger inclusions than the existing ones would be necessary for having the same effect in lower strength materials.
A last question to be addressed in that context is how the information on the inclusion size in the material can be
obtained. The optimum method is the evaluation of the fracture surfaces of fatigued specimens, see, e.g. [41]. Sometimes
16 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Fig. 15. The effect of the diameter of nonmetallic inclusions on the fatigue limit of a high strength spring steel; according to [47].

Fig. 16. Initial crack sizes and crack sizes at crack arrest obtained by the crack arrest model described in Section 4.2 for low strength and high strength
steels (base materials).

specimen preparation is accompanied by additional measures such as tempering [47] or utilizing hydrogen embrittlement
[48] before testing. Less accurate but time-saving is the metallographic investigation. Note, however, that this has to be com-
bined with a statistics of extremes method [4,48] in order to avoid misleading results. This is because fatigue is a weakest-
link phenomenon, i.e., it is triggered by the largest defect (or the largest defects) present in the highly loaded volume which
also explains part of the well-known size effect as well as the loading type effect (tension versus bending) of the fatigue
strength. The fatigue limit is decreased when the highly loaded volume becomes larger and the chance is higher to trigger
an even larger inclusion [49], see also the discussion on the size effect of the fatigue strength in [44]. Two examples of statis-
tics of extreme values of casting defects and inclusions are shown in Fig. 17.

6.2.2. Pores, cavities, micro-shrinkages


This kind of initial flaws is typical for casts, e.g. of aluminium alloys [50–56], sintered materials [57], materials manufac-
tured by additive manufacturing, e.g., selective laser melting (for a review on this see [58]) and, to a lesser extent, for weld-
ments [40]. In nodular cast iron there might be a competition between the graphite nodules and micro-shrinkage pores with
the latter being preferential crack initiation sites [59], see also [60–62,51]. Note that pores, cavities and micro-shrinkages are
much larger in size than the non-metallic inclusions discussed above. An issue reported for pearlite gray iron is multiple
cracking [63]. In structures manufactured by selective laser melting cavities formed by non-melted powder and clusters
of pores are a common problem [64,65].
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 17

Fig. 17. Examples of statistics of extreme values; (a) casting defects in a superalloy; according to [50]; (b) and (c) non-metallic inclusions in low strength
steel, IBESS project, this issue [37].

Fig. 18. Classical crack initiation sites in manual welds; according to [66].

6.2.3. Weld defects


Classical weld defects in manual arc (stick) welded mild steels are slag inclusions, spatter and microcracks as material and
undercuts as geometric defects (Fig. 18). We will briefly come back to the latter at the end of Section 6.3.1. Note, however,
that no such defects have been found at the toes of the weldments investigated in the IBESS project cluster which are
reported in this Special issue. What is, however, found is distinct multiple crack propagation at stress levels above the endur-
ance limit such as shown in Fig. 12. This feature is well known for weldments. Fig. 19 illustrates the differences with respect
to manual and automatic welds. Obviously the nonuniform local geometry of the weld toe profiles have an impact on the
complete pattern of the propagation of the multiple cracks this way influencing the lifetime. Automatic welds tend to a lar-
ger number of initial cracks which earlier transform to a long surface crack by coalescence than found in manual welds
[66,67]. Note that this can have a detrimental effect on lifetime. On the other hand it should be easier to generate uniform
smooth notches and flatter flank angles by automatic welding, which would be beneficial with respect to fatigue strength.
See also the discussion in [40].

6.2.4. Defective microstructures


Surface-near areas of defective microstructures can also be crack initiation sites. Since this is a very wide field, no detailed
discussion is possible here. Examples are surface damage introduced during shot peening on a decarburized layer in spring
steel [4], an a phase area close to the surface in a titanium alloy which was observed to be a crack initiation site [68] and a
‘‘shattered” microstructure area in an aluminium alloy due to voids, gas porosity, shrinkage and dendritic grain coarsening
[69].
18 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Fig. 19. Typical fracture surfaces of transverse butt welded plates. (a) Automatic weld and (b) manual weld, according to [66].

6.2.5. Microcracks and crack like features


Finally, microcracks may form due to residual stresses during manufacturing or caused by other reasons. Sometimes, e.g.,
in friction stir welding (where it can be a major problem, see, e.g. [70,71]) or selective laser melting of aluminium alloys
[72,73], oxide layers can get into the melt pool where they, during solidification, fragment and disperse. The fragments
can act as microcracks or, at least, as sites for microcrack generation due to micro-notch effects and different stiffness com-
pared to the matrix material. Examples of surface cracks formed in materials manufactured by selective laser melting are
provided in [74].

6.3. Geometrical irregularities

6.3.1. Surface roughness


It is not necessary to write much about the effect of surface roughness on the fatigue strength. The effect is that of local
stress concentrations at the surface which promotes crack nucleation and early growth this way lowering the fatigue
strength. That it may have different impacts at different load levels of the S-N curve is illustrated in Fig. 20 for steel. The
roughness effect increases with the ultimate tensile strength of the material (keyword: notch sensitivity), an effect which,
among others, is taken into account by the German FKM Guidelines [75]. Note that this is also found by applying the model
presented in this Special issue to the simulation of the fatigue limit of steels of different strengths [76]. Sometimes, as in the
case of friction stir welding, surface roughness introduced by the manufacturing process may cause microcracking even
before loading [77].
Note that within the fracture mechanics approach presented within this Special issue the term ‘‘secondary notches” is
used as a collective term for geometrical irregularities which, besides roughness includes features such as undercuts at
the weld toe or edge points of weld-re-solidification ripples. During the simulations it became obvious that the secondary
notches were of major importance with regard to the fatigue strength [39], see also [76].

Fig. 20. Schematic S-N curves for machined and polished as well as for as-forged surfaces of steel; according to [78].
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 19

6.3.2. Corrosion pits


Another kind of surface impairment is corrosion pits which usually arise during service operation. A typical example is
found for uncoated railway axles usually of freight cars. Since it is known that these pits can act as crack initiation sites,
a number of studies have been carried out to quantify the actual state in various railway fleets (e.g. [79], see also the over-
view in [44]). An example is provided in Fig. 21 where an impression is given on the size of such defects. Corrosion pits can
show quite different geometries (deep and narrow or wide and shallow) [80] because of which it is not a trivial task to deter-
mine stress concentration factors for them. In Fig. 22 [81] the effect of secondary corrosion pits within the main dimples is
simulated for simplified geometrical conditions. The effect of the corrosion pit depth on the S-N curve of a turbine steel blade
is shown in Fig. 23 [82]. Fatigue initiated by corrosion pits is frequently combined with multiple cracking, see [83] as an
example of an aluminium alloy.

6.3.3. Mechanical surface impairment


Mechanical surface impairments are damages due to impact events and scratches.

6.3.3.1. Impact notches. Examples for the first one are flying ballast impact notches [44]. Note that impact damage can also
promote stress corrosion crack initiation, e.g., by local damage of coatings and that it introduces a complex residual stress
state. An example of a fatigue crack in a railway axle which initiated from a ballast indentation is reported in [84].

6.3.3.2. Scratches. The effect of scratches is similar to that of the surface roughness, i.e., scratches form stress concentrators.
In addition, they cause strongly localized plastic deformation and residual stresses which have an effect on crack nucleation
and early crack propagation [85]. An example for the influence of scratches on the S-N curve is shown in Fig. 24.

Fig. 21. Statistical distribution of corrosion pit sizes of three ex-service railway axes made of A1N steel: axles 1–2 were freight axles whilst axle 3 was from
a regional train; according to [79].

Fig. 22. The effect of secondary corrosion pits on stress concentration using simplified geometrical conditions, (a) Kt without secondary pit and (b) Kt with
secondary pit; according to [81].
20 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

Fig. 23. Experimental S-N curves on specimens with corrosion pits of different depths at (a) R = 0.05 and (b) R = 0.4, steam turbine blade steel; according to
[82].

Fig. 24. Effect of notch depth and notch root radius of scratches on the fatigue life of an aluminium alloy, rmax = 200 MPa, R = 0.1; according to [86].

6.3.4. Fracture mechanics treatment of geometrical defects


The distinction between material and geometrical defects in the previous sections brings up the question of how to han-
dle this in a fracture mechanics analysis. In principle there are two options, e.g. [61]: (a) The defects are treated as notches at
the crack driving force side of the component. This requires great effort since the effect of the flaws on the local stress state
has to be established which is usually done by finite element analyses. (b) The defects are treated as cracks what, of course, is
much easier. Note, however, that for doing so, the crack length has to be large enough compared to the notch radius [4]. Both
alternatives (a and b) have been used in the past, e.g. [61,87–91].
Within the IBESS project, the results of which are the topic of this Special issue, both philosophies were applied with
respect to the treatment of the secondary notches (Fig. 25), however, (a) treating the secondary notch as notch (Fig. 25 b)
was realized for 10 mm thick plates only. For this, stress-depth profiles have been provided in table format for butt and cru-
ciform welds loaded by tension and bending for a wide range of geometrical input parameters of the weld toe such as illus-
trated in Fig. 25a [92]. The major part of the analyses was performed based on the option of Fig. 25c [39], i.e., the initial crack
length used in the analyses was ai + anotch. Both parameters were used as statistical variables.

7. Summary

A discussion has been provided on using fracture mechanics to the determination of the fatigue strength and life of com-
ponents. This comprised the description of short crack propagation based on an elastic plastic concept for the crack driving
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 21

Fig. 25. Treatment of the secondary notch depth in the present approach; (a) the most important geometrical parameters considered; (b) the secondary
notch is treated as notch. For this case through thickness stress profiles are provided for 10 mm thick plates; (c) the notch depth anotch is treated as part of
the initial crack length.

force and an approach for describing the gradual build-up of the crack closure effects, a discussion on the adequate definition
of the initial crack size needed for fracture mechanics considerations and the multiple crack phenomenon at load levels
above the endurance limit. The paper provides basic information for better understanding a number of papers of this Special
issue which introduce a methodology for fracture mechanics based determination of the fatigue strength of weldments.

Acknowledgements

This work was part of the DFG/AiF research cluster ‘‘IBESS’’. The authors gratefully appreciate the funding by the AiF net-
work (Arbeitsgemeinschaft industrieller Forschungsvereinigungen) and the German Research Foundation (Deutsche
Forschungsgemeinschaft).

References

[1] Sakai T, Nakagawa A, Oguma N, Nakamura Y, Ueno A, Kikuchi S, et al. A review on fatigue fracture modes of structural metallic materials in very high
cycle regime. Int J Fatigue 2016;93(Part 2):339–51.
[2] Haibach E. A likely explanation of very high cycle fatigue in steel. In: 5th Int conf on very high cycle fatigue, Berlin; 2011.
[3] Zerbst U, Schwalbe K-H, Ainsworth RA. An overview of failure assessment methods in codes and standards. In: Milne I, Ritchie RO, Karihaloo B, editors.
Comprehensive structural integrity (CSI), vol. 7. Amsterdam: Elsevier; 2003. p. 4–48 [Chapter 7.01].
[4] Murakami Y. Metal fatigue. Effects of small defects and nonmetallic inclusions. Oxford: Elsevier; 2002.
[5] Miller KJ. The two thresholds of fatigue behavior. Fatigue Fract Eng Mater Struct 1993;16:931–9.
[6] Beretta S, Carboni M, Madia M. Modelling of fatigue thresholds for small cracks in a mild steel by ‘‘Strip-Yield” model. Eng Fract Mech
2009;76:1548–61.
[7] Kitagawa H, Takahashi S. Applicability of fracture mechanics to very small cracks or the cracks in the early stage. In: Proc 2nd intern conf mech behav
mater, Boston, ASM, Cleveland, Ohio; 1976. p. 627–31.
[8] El Haddad MH, Smith KN, Topper TH. Fatigue crack propagation of short cracks. Trans ASME, J Eng Mater Technol 1979;101: 42–6.
[9] Hertel O, Vormwald M. Statistical and geometrical size effects in notched members based on weakest-link and short-crack modelling. Eng Fract Mech
2012;95:72–83.
[10] Polak J. Cyclic deformation, crack initiation, and low-cycle fatigue. In: Ritchie RO, Murakami Y, editors. Comprehensive structural integrity, volume 4:
cyclic loading and fracture. Elsevier; 2003. p. 1–39.
[11] Tanaka K. Fatigue crack propagation. In: Ritchie RO., Murakami Y, editors. Comprehensive structural integrity; volume 4: cyclic loading and fracture.
Elsevier; 2003. p. 95–127.
[12] Elber W. Fatigue crack closure under cyclic tension. Eng Fract Mech 1970;2:37–45.
[13] Tanaka K, Akinawa Y. Modelling of fatigue crack growth: mechanistic models. In: Ritchie RO, Murakami Y, editors. Comprehensive structural integrity;
volume 4: cyclic loading and fracture. Elsevier; 2003. p. 165–89.
[14] Suresh S. Fatigue of materials, 2nd ed. Cambridge: Cambridge University Press; 2003.
[15] Pippan R, Hohenwarter A. Fatigue crack closure: a review of the physical phenomenon. Fatigue Fract Eng Mater Struct 2017;40:471–95.
[16] Zerbst U, Vormwald M, Pippan R, Gänser H-P, Sarrazin-Baudoux C, Madia M. About the fatigue crack propagation threshold of metals as a design
criterion – a review. Eng Fract Mech 2016;153:190–243.
[17] Schlitzer T. Hitherto unpublished results.
[18] Pokluda J, Pippan R, Vojtek T, Hohenwarter A. Near-threshold behavior of shear-mode fatigue cracks in metallic materials. Fatigue Fract Eng Mater
Struct 2014;37:232–54.
[19] Hadrboletz A, Weiss B, Stickler R. Fatigue threshold of metallic materials – a review. In: Carpinter A, editor. Handbook of fatigue crack propagation in
metallic structures. Elsevier; 1994. p. 847–82.
[20] Pippan R., Riemelmoser FO. Modelling of fatigue growth: Dislocation models. In: Ritchie RO, Murakami Y, editors. Comprehensive structural integrity;
volume 4: cyclic loading and fracture. Elsevier; 2003. p. 191–207.
[21] Dowling NE. Crack growth during low-cycle fatigue of smooth axial specimens. ASTM STP 637; 1977. p. 97–121.
[22] Tanaka K. The cyclic J-integral as a criterion for fatigue crack growth. Int J Fract 1983;22:91–104.
[23] Heitmann H, Vehoff H, Neumann P. Random load fatigue of steels – service life prediction based on the behaviour of microcracks. In: Proc int conf on
application of fracture mechanics to materials and structures, Freiburg, Germany; 1983.
22 U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23

[24] Vormwald M, Seeger T. The consequences of short crack closure on fatigue crack growth under variable amplitude loading. Fatigue Fract Eng Mater
Struct 1991;1(3):205–25.
[25] Tchoffo Ngoula D, Madia M, Beier H-Th, Vormwald M, Zerbst, U. Cyclic J integral – numerical and analytical investigations on weldments. Eng Fract
Mech 2018;198:24–44.
[26] Saxena A. Nonlinear fracture mechanics for engineers. In: Fatigue crack growth under large-scale plasticity. CRC Press; 1998. p. 267–305 [Chapter 9].
[27] Lamba HS. The J-integral applied to cyclic loading. Eng Fract Mech 1975;7:693–703.
[28] Wüthrich C. The extension of the J-integral concept to fatigue cracks. Int J Fract 1982;20:R35–7.
[29] McClung RC, Chell GG, Lee Y-D, Russell DA, Orient GE. Development of a practical methodology for elastic-plastic and fully plastic fatigue crack growth.
NASA Report NASA/CR-1999-209428; 1999.
[30] R6, Revision 4. Assessment of the integrity of structures containing defects. Barnwood (Gloucester): British Energy Generation Ltd (BEGL); 2009.
[31] Zerbst U, Schödel M, Webster S, Ainsworth RA. Fitness-for-service fracture assessment of structures containing cracks. A workbook based on the
European SINTAP/FITNET procedure. Elsevier; 2007.
[32] McClung RC, Chell GG, Lee Y-D, Russel DA, Orient GE. A practical methodology for elastic-plastic fatigue crack growth. ASTM STP 1296; 1997. p. 317–
37.
[33] Kucharzcyk P, Madia M, Zerbst U, Schork B, Gerwin P., Münstermann S. Fracture mechanics based prediction of the fatigue strength of weldments:
material aspects. Eng Fract Mech 2018;198:79–102.
[34] Zerbst U, Madia M. Fracture mechanics based assessment of the fatigue strength: approach for the determination of the initial crack size. Fatigue Fract
Eng Mater Struct 2014;38:1066–75.
[35] Maierhofer J, Kolitsch S, Pippan R, Gänser H-P, Madia M., Zerbst U. The cyclic R curve – determination, problems and limitation. Eng Fract Mech
2018;198:45–64.
[36] Vormwald M, Heuler P, Seeger T. A fracture mechanics based model for cumulative damage assessment as part of fatigue life prediction. ASTM STP
1122. Philadelphia: American Society for Testing and Materials; 1992. p. 28–43.
[37] Schork B, Kucharzcyk P, Tchuindjang D, Kaffenberger M, Bernhard J, Madia M., et al. The effect of the local weld geometry and material defects on crack
initiation and fatigue strength. Eng Fract Mech 2018;198:103–122.
[38] Zerbst U, Madia M, Schork B. Fracture mechanics based determination of the fatigue strength of weldments. Proc Struct Integr 2016;1:10–7.
[39] Madia M, Zerbst U, Beier HTh, Schork B. The IBESS model – elements, realization, validation. Eng Fract Mech 2018;198:171–208.
[40] Zerbst U, Ainsworth RA, Beier HTh, Pisarski H, Zhang ZL, Nikbin K, et al. Review on fracture and crack propagation in weldments – a fracture mechanics
perspective. Eng Fract Mech 2014;132:200–76.
[41] Zerbst U, Madia M, Hellmann D. An analytical fracture mechanics model for estimation of S-N curves of metallic alloys containing large second
particles. Eng Fract Mech 2011;82:115–34.
[42] Murakami Y. High and ultrahigh cycle fatigue. In: Ritchie RO, Murakami Y, editors. Comprehensive structural integrity; volume 4: cyclic loading and
fracture. Elsevier; 2003. p. 41–76.
[43] Lambrighs K, Verpoest I, Verlinden B, Wevers M. Influence of non-metallic inclusions on the fatigue properties of heavily cold drawn steel wires. Proc
Eng 2010;2:173–81.
[44] Zerbst U, Beretta S, Köhler G, Lawton A, Vormwald M, Beier HTh, et al. Safe life and damage tolerance aspects of railway axles – a review. Eng Fract
Mech 2013;98:214–71.
[45] Murtaza G, Akid R. Empirical corrosion fatigue life prediction models of a high strength steel. Eng Fract Mech 2000;67:461–74.
[46] Juvonen P. Effects of non-metallic inclusions on fatigue properties of calcium treated steels PhD. Espoo: Helsinki Univ of Technology; 2004.
[47] Rödling S, Fröschl J, Hück M, Decker M. Einfluss nichtmetallischer Einschlüsse auf zulässige HCF-Bemessungskennwerte. Deutscher Verband für
Materialforscnung und -prüfung (DVM), Report No. 137; 2010. p. 135–45.
[48] Murakami Y. Material defects as the basis for fatigue design. Int J Fatigue 2012;41:2–10.
[49] Beretta S, Ghidini A, Lombardo F. Fracture mechanics and scale effects in the fatigue of railway axles. Eng Fract Mech 2005;72:195–208.
[50] Kunz L, Lukaš P, Konecna R, Fintova S. Casting defects and high temperature fatigue life of IN 713LC superalloy. Int J Fatigue 2012;41:47–51.
[51] Zhu X, Jones JW, Allison JE. Effect of frequency, environment, and temperature on fatigue behavior of E319 cast aluminum alloy: stress-controlled
fatigue life response. Metall Mater Trans A 2008;39:2681–8.
[52] Mayer H, Papakyriacou M, Zettl B, Stanzl-Tschegg SE. Influence of porosity on the fatigue limit of die cast magnesium and aluminium alloys. Int J
Fatigue 2003;25:245–56.
[53] Yi JZ, Zhu X, Jones JW, Allioson JE. A probabilistic model of fatigue strength controlled by porosity population in a 319-type cast aluminum alloy: Part II.
Monte-Carlo simulation. Metall Mater Trans A 2007;38:1123–35.
[54] Jang Y, Jeong Y, Yoon C, Kim S. Fatigue life prediction for porosity containing cast 319–T7 aluminum alloy. Metall Mater Trans A 2009;40:1090–9.
[55] Jang Y, Jin S, Jeong Y, Kim S. Fatigue crack initiation mechanism for cast 319-T7 aluminum alloy. Metall Mater Trans A 2009;40:1579–87.
[56] Tijani Y, Heinrietz A, Stets W, Voigt P. Detection and influence of shrinkage pores and nonmetallic inclusions on fatigue life of cast aluminum alloys.
Metall Mater Trans A 2013;44:5408–15.
[57] Polasik SJ, Williams JJ, Chawla N. Fatigue crack initiation and propagation of binder-treated power metallurgy steels. Metall Mater Trans A
2002;33:73–81.
[58] Zerbst U, Hilgenberg K. Schadensentwicklung und Schadenstoleranz von SLM-gefertigten Strukturen. In: Richard HA, Schramm B, Zipsner T, editors.
Additive Fertigung von Bauteilen und Strukturen. Wiesbaden: Springer Vieweg; 2017.
[59] Shirani M, Härkegård G. Damage tolerant design of cast components based on defects detected by 3D X-ray computed tomography. Int J Fatigue
2012;41:188–98.
[60] Endo M, Yanase K. Effects of small defects, matrix structures and loading conditions on the fatigue strength of ductile cast irons. Theor Appl Fract Mech
2014;69:34–43.
[61] Costa N, Machado N, Silva FS. A new method for prediction of nodular cast iron fatigue limit. Int J Fatigue 2010;32:989–95.
[62] Nasr A, Bouraoui Ch, Fathalláh R, Nadot Y. Probabilistic high cycle fatigue behaviour of nodular cast iron containing casting defects. Fatigue Fract Eng
Mater Struct 2009;32:292–309.
[63] Fash JW. Fatigue crack initiation and growth in gray cast iron. Fracture control program report no. 35. Urbana (IL): University of Illinois; 1980.
[64] Kasperovich G, Hausmann J. Improvement of fatigue resistance and ductility of TiAl6V4 processed by selective laser melting. J Mater Proc Technol
2015;220:202–14.
[65] Strantza M, Vafadari R, de Baere D, Vrancken B, van Paepegem W, Vandendael I, et al. Fatigue of Ti6Al4V structural health monitoring systems
produced by selective laser melting. Materials 2016;9:106. doi: http://dx.doi.org/10.3390/ma9020106.
[66] Otegui JL, Kerr HW, Burns DJ, Mohaupt UH. Fatigue crack initiation from defects at weld toes in steel. Int J Press Vess Piping 1989;38:385–417.
[67] Gurney TR. Fatigue of welded structures. 2nd ed. Cambridge University Press; 1979.
[68] Chan KS, Koike M, Mason RL, Okabe T. Fatigue life of titanium alloys fabrikated by additive layer manufacturing techniques for dental implants. Metall
Mater Trans A 2012;44:1010–22.
[69] Mayer H, Papakyriacou M, Zettl B, Vacic S. Endurance limit and threshold stress intensity of die cast magnesium and aluminium alloys at elevated
termperatures. Int J Fatigue 2005;27:1076–88.
[70] Mishra RS, Ma ZY. Friction stir welding and processing. Mater Sci Eng 2005;R50:1–78.
[71] Zhou C, Yang X, Luan G. Effect of oxide array on the fatigue property of friction stir welds. Scr Mater 2006;54:1515–20.
[72] Louvis E, Fox P, Sutcliffe CJ. Selective laser melting of aluminium components. J Mater Process Technol 2011;211:275–84.
U. Zerbst et al. / Engineering Fracture Mechanics 198 (2018) 2–23 23

[73] Olakanmi EO, Cochrane RF, Dalgarno KW. A review on selective laser sintering/melting (SLS/SLM) of aluminium alloy powders: Processing,
microstructure, and properties. Progr Mater Sci 2015;74:401–77.
[74] Vrancken B, Wauthle R, Kruth J-P, van Humbeeck J. Study of the influence of material properties on residual stress in selective laser melting. In: Proc
solid freedom fabrication symposium, Austin; August 2013. p. 1–15.
[75] Rennert R, Kulling E, Vormwald M, Esderts A, Siegele D. FKM guideline. Analytical strength assessment of components made of steel, cast iron and
aluminium materials in mechanical engineering. 6th ed. Frankfurt/Main: VDMA Verl; 2013.
[76] Madia M, Zerbst U. Application of the cyclic R-curve method to notch fatigue analysis. Int J Fatigue 2016;82:71–9.
[77] James MN, Hattingh DG, Bradley GR. Weld tool travel speed effects on fatigue life of friction stir welds in 5083 aluminium. Int J Fatigue
2003;25:1389–98.
[78] McKelvey SA, Fatemi A. Surface finish effect on fatigue behavior of forged steel. Int J Fatigue 2012;36:130–45.
[79] Beretta S, Carboni M, Lo Conte A, Palermo E. An investigation of the effects of corrosion on the fatigue strength of AlN axle steel. Proc Inst Mech Eng Part
F: J Rail Rapid Transit 2008;222:129–43.
[80] Shan-hua X, Wang Y-D. Estimating the effects of corrosion pits on the fatigue life of steel plate based on the 3D profile. Int J Fatigue 2015;72:27–41.
[81] Cerit M, Genel K, Eksi S. Numerical investigation on stress concentration of corrosion pit. Eng Failure Anal 2009;16:2467–72.
[82] Schönbauer BM, Stanzl-Tschegg SE, Perlega A, Salzman RN, Rieger NF, Turnbull A, et al. The influence of corrosion pits on the fatigue life of 17-4PH
steam turbine blade steel. Eng Fract Mech 2015;147:158–75.
[83] Van der Walde K, Brockenbrough JR, Craig BA, Hillberry BM. Multiple fatigue crack growth in pre-corroded 2024-T3 aluminum. Int J Fatigue
2005;27:1509–18.
[84] Australian Transport Safety Bureau. Derailment of XPT passenger train ST22 Harden, New South Wales; 9 February 2006. http://www.atsb.gov.
au/media/24374/rair2006002_001.pdf.
[85] Zhixin Z, Weiping H, Miao Z, Qingchun M. The fatigue life prediction for structure with surface scratch considering cutting residual stress, initial
plasticity damage and fatigue damage. Int J Fatigue 2015;74:173–82.
[86] Cini A, Irwing PE. Transformation of defects into fatigue cracks; the role of Kt and defect scale on fatigue life of non-pristine components. Proc Eng
2010;2:667–77.
[87] Lukas P, Kunz L, Weisse B, Stickler R. Non-damaging notches in fatigue. Fatigue Fract Eng Mater Struct 1986;9:195–204.
[88] Tanaka K, Nakai Y, Yamashita M. Fatigue growth threshold of small cracks. Int J Fract 1981;17:519–33.
[89] El Haddad MH, Topper TH, Smith KN. Prediction of non-propagating cracks. Eng Fract Mech 1997;11:573–84.
[90] Amiri M, Arcari A, Airoldi L, Naderi M, Iyyer N. A continuum damage mechanics model for pit-to-crack transition in AA2024-T3. Corros Sci
2015;98:678–87.
[91] Savaidis G, Savaidis A, Zerres P, Vormwald M. Mode I fatigue crack growth at notches considering crack closure. Int J Fatigue 2010;32:1543–58.
[92] Zerbst U. Schlussbericht zum IGF-Vorhaben ‘‘Analytische bruchmechanische Ermittlung der Schwingfestigkeit‘‘ (IBESS – A3), Appendix B; 2016.
https://opus4.kobv.de/opus4-bam/frontdoor/index/index/docId/35989.

You might also like