You are on page 1of 9

Procedia Engineering 55 ( 2013 ) 636 644

1877-7058 2013 The Authors. Published by Elsevier Ltd.


Selection and peer-review under responsibility of the Indira Gandhi Centre for Atomic Research.
doi: 10.1016/j.proeng.2013.03.307
6
th
International Conference on Cree
Damage Mechanisms and
Very Hi
H
Department of Materials Science & Engineering, Un
Abstract
The primary cause of fatigue damage is usually intim
material and details of the loading conditions, variou
of fatigue damage evolution in low and high cycl
discussed. Various forms of fatigue damage and life-
will be illustrated. In particular, the fatigue propert
microstructural mechanisms and the question of the e

2013 The Authors. Published by Elsevier Ltd. Sele
Centre for Atomic Research

Keywords: Fatigue crack initiation; LCF; HCF; UHCF; fatig
1. Introduction
The history of fatigue research dates back to
became an increasingly important problem. This
behaviour of steel specimens subjected to cycl
kinds of metals and other materials have been
has remained a topic of prime interest, academ
materials and their use in more and more deman
Only after Ewing and Humfrey had shown th
developed at the surface during cyclic loadin
microstructural processes that lead to fatigue
deformation generally leads to microstructural
After initial cyclic hardening or softening, some
of fatigue damage become apparent which ultim
and the responsible life-controlling microstructu

Corresponding author:
E-mail address: mughrabi@ww.uni-erlangen.de
ep, Fatigue and Creep-Fatigue Interaction [CF-6

Fatigue Lives: From the Low to th
igh Cycle Regime
ael Mughrabi


niversity of Erlangen-Nrnberg,Martensstr. 5, 91058 Erlangen, German
mately related to some form of cyclic slip irreversibilities. Depen
s types of fatigue crack initiation are observed. Here, typical ex
le fatigue (LCF, HCF) and ultrahigh cycle fatigue (UHCF)
-controlling microstructural mechanisms in different metals and
ties of ultrafine-grained (UFG ) materials will be discussed, a
existence of a fatigue limit in the UHCF regime will be considere
ection and/or peer-review under responsibility of the Indira Gand
gue lives; fatigue limit
o the middle of the 19
th
century, when failures of railway
s was the motivation for Whlers first systematic studies
lic loading [1]. Since then, fatigue damage and failures
studied extensively. Until today, research on fatigue heh
mically and with regard to the development of new imp
nding fields of application.
hat, in fatigued steels, microcracks initiated in slip bands
ng [2], were the mechanisms of cyclic deformation a
failures studied in more detail. Today, it is clear that
changes by the accumulation of repeated cyclic micros
etimes followed by a steady state of cyclic saturation, firs
mately lead to fatigue failure. The evolution of fatigue d
ural mechanisms vary from material to material and depe
6]
he
ny
ding on
xamples
will be
d alloys
and the
ed.
dhi
y axles
s of the
s of all
haviour
proved
which
nd the
cyclic
strains.
st signs
damage
end on
Available online at www.sciencedirect.com
2013 The Authors. Published by Elsevier Ltd.
Selection and peer-review under responsibility of the Indira Gandhi Centre for Atomic Research.
637 Hael Mughrabi / Procedia Engineering 55 ( 2013 ) 636 644
the conditions of cyclic loading. In ductile metals, some sort of cyclic strain localization (e.g. in persistent slip
bands: PSBs) frequently occurs and leads to the initiation of cracks at the surface, compare [3]. Other forms of
fatigue damage are also quite common, e.g. at grain boundaries. On the other hand, in heterogeneous materials
containing inclusions or pores, the latter can be sites of beginning subsurface fatigue damage and can even
control fatigue life n the regime of very low loading amplitudes and correspondingly very high fatigue lives,
when fatigue damage at the surface develops only very slowly. Materials such as high-strength steels
containing inclusions show this behaviour in the regime of ultrahigh cycle fatigue (UHCF) or very high cycle
fatigue (VHCF).
The present work focuses on just a few of the characteristics of fatigue behaviour. The different types of
fatigue life diagrams for low and high cycle fatigue (LCF, HCF) will first be described, followed by a
discussion of the significance of fatigue crack initiation relative to fatigue crack propagation. Then, the
microstructural mechanisms of cyclic strain localization in PSBs will be discussed. The final two sections are
devoted to two topics of quite recent interest, namely the fatigue of ultrafine-grained (UFG) materials and
fatigue in the UHCF regime of very low loading and exceptionally high fatigue lives.
2. Fatigue life representations Whler (S-N), Coffin-Manson and Basquin Plots
Since the pioneering experiments of Whler, it has become customary to plot fatigue life data in so-called
Whler or S-N diagrams (S: stress, N: number of cycles). It is interesting to note that, to the best knowledge of
the author, Whler never presented his data in graphical form bu usually only in lengthy tables [1]. Figure 1
shows schematically a Whler (S-N) plot in which the stress amplitude /2 (: stress range) is plotted
against the number of cycles to failure, N
f
. In this particular case, a plateau (stress level R
w
) is indicated in the
low-stress (long life) HCF regime, implying that a fatigue limit exists in the sense that at stress levels below the
plateau, fatigue life should become infinite. For decades, the general opinion was that, in particular in the case
of ferritic steels, a true fatigue limit existed in the HCF range.


Fig. 1. Schematic Whler (S-N) fatigue life curve (with a fatigue limit), hysteresis loops are indicated. After [3].
Only in recent years has it been questioned whether such a fatigue limit extends beyond the HCF into the
UHCF range, compare Section 6. While it had become clear after the work of Ewing and Humfrey that there
must be some microplastic deformation occurring during fatigue, the plastic strain amplitude had not been used
in the description of fatigue life until Coffin [4] and Manson [5] formulated the fatigue life law named after
them in the fifties of the last century. The role played by the plastic strain amplitude
pl
/2 (
pl
: plastic strain
range) is indicated in Fig. 1, showing typical open hysteresis loops (stress vs. strain ), for different levels of
the stress amplitude. The widths of these loops at zero stress (identical to
pl
) decreases as the stress level
decreases but vanishes only below the fatigue limit at which it is still finite (some 10
-5
to 10
-4
), indicating that
such low plastic strain amplitudes are non-damaging, probably because they are almost reversible in a
microstructural sense and do not cause irreversible damaging changes at the surface.
638 Hael Mughrabi / Procedia Engineering 55 ( 2013 ) 636 644
Following the work of Morrow [6], very useful fatigue life laws based on cyclic-stress-strain relationships
can be formulated. In addition to the Whler or S-N plot of stress amplitude /2 vs. N
f
, it is common to make
also use of the Basquin relation for the HCF range:
b
f
f el
) 2 (
'
2 2
N
E E

=

(1)
and the Coffin-Manson law for the LCF regime:
c
f f
pl
) 2 ( '
2
N

(2)
Here,
el
/2 is the elastic strain amplitude, E is Youngs modulus;
f
and
f
are the fatigue strength and
fatigue ductility coefficients, respectively, and b and c are the fatigue strength and the fatigue ductility
exponents, respectively. Since the total strain is
t =

pl
+
el
, these two laws add up to the total strain fatigue life
relationship, compare [6]:
c
f f
b
f
f
pl
el t
) 2 ( ' ) 2 (
'
2 2 2
N N
E


+ =

(3)
which corresponds in the asymptotic limit of large strain amplitudes to the Coffin-Manson relation, whereas, at
very low amplitudes, it approaches the relation between stress amplitude and number of cycles to failure as in
the Whler-type S-N plot. For this reason, the use of this total strain fatigue life relation is sometimes favoured
and very useful, see Section 5.
3. Early fatigue damage: the french curve initiation of critical cracks
As early as in 1933, French proposed the so-called critical damage curve for materials of known S-N
curve exhibiting a fatigue limit [7]. This approach assumes that the material is initially defect-free. The French
curve which is obtained in a two-stage fatigue experiment, compare [8], represents cum grano salis the
number of cycles necessary for the initiation of a crack of a critical length which is roughly equal to the critical
size of non-propagating cracks. Depending on the details, the lengths of these initially formed microcracks in a
fatigued steel have been found by Klesnil [8] to lie in the range between just a few and some ten microns. The
number of cycles separating the French curve from the S-N curve at a particular stress level can be considered
to represent the period of crack propagation after microcracks of critical size have been initiated.
Figure 2 shows, as an example from Klesnils work, a French curve obtained on the Czechoslovak low-
carbon steel SN 12 010 [8]. In this case, microcracks of a few microns in length were noted at stress levels
below the French curve, and longer microcracks of ca. 100 m length, lying in persistent slip bands (PSBs),
were observed at higher stress levels. In the present context, it is important to note, compare Fig. 2, that the
period of crack propagation occupies a large fraction of fatigue life at high stress levels in the LCF range and
decreases continuously with decreasing stress level until it becomes almost negligible at the HCF level of the
fatigue limit. Thus, in HCF, fatigue crack initiation occupies a major fraction of fatigue life, as found in some
standard text-books [9] and in more detail in [10]. Moreover, this should be even more true important in UHCF
at very high N
f
, see Section 6.

639 Hael Mughrabi / Procedia Engineering 55 ( 2013 ) 636 644

Fig. 2. Example of French curve for steel SN 12010, compared with fatigue life curve. Failures or no failure are indicated by
upward/downward. arrows, respectively. From [10], after M. Klesnil [7]
4. Cyclic strain localization in persistent slip bands
Cyclic strain localization in PSBs and the initiation of fatigue cracks in fcc metals have received much
attention since the pioneering work of Thompson et al. [11]. In PSBs, the local plastic strain amplitude is
enhanced by two orders of magnitude or more, compared to the surrounding matrix. Fatigue crack initiation
occurs at the surface at the sites of emerging PSBs, where complex extrusions develop. While it is true that the
importance of PSBs is more or less confined to fatigued simple ductile materials and is of much less relevance
with respect to the fatigue resistance of structural materials of practical importance, the work on PSBs has
fascinated many researchers and has led to a generally improved understanding of dislocation behaviour in
fatigued metals, as discussed, for example, in [3,12,13,14].
With the aid of TEM it became possible to characterize in detail the dislocation distributions and to identify
the governing dislocation mechanisms of cyclic strain localization in PSBs of fcc single crystals with the so-
called ladder structure or wall structure. Based thereupon, Essmann et al.

[12] developed a semi-
quantitative model (called EGM model subsequently) of the evolution of the surface relief of PSBs at the sites
of emerging PSBs, in fair agreement with observations. In the model, the steady-state equilibrium of the defect
content in cyclic saturation is analyzed semi-quantitatively, taking into account the production and annihilation
of dislocations and point defects. For details, the reader is referred to [3,12,13]. The underlying assumptions are
so general that the predictions, perhaps with slight modifications, will also apply to other materials in which
PSBs do not necessarily have the ladder structure. In passing, it should be noted that Brown had developed a
detailed model of the dislocation processes in PSBs, paying special attention to the stress singularity at the
surface site of an emerging PSB [15]. In a later extension of the model, Brown and Nabarro [16] arrived at a
model of extrusion formation that, in its net result, is very similar to the EGM model.


Fig.3. Schematic illustration of evolution of surface relief at emerging PSB in the absence of diffusional effects. a) Interfacial dislocations
accumulate at PSB-matrix interfaces. b) Rapid formation of static extrusions by emergence of interfacial PSB-matrix dislocations.
c) Gradual roughening of surface of extrusion by random slip in PSB. After [3,12,13].
initiation
growth
Failure
French
Curve
initiation
growth
Failure
French
Curve
c) c)
640 Hael Mughrabi / Procedia Engineering 55 ( 2013 ) 636 644
Figure 3 shows in a strongly simplified view the subsequent stages of the dislocation pattern in a PSB, as
envisaged in the EGM model. In this picture, the dislocation dipoles which prevail in a high density are
omitted. Then, in cyclic saturation, essentially only so-called PSB-matrix interface dislocation layers are left, as
shown in Fig. 3a. During subsequent cyclic deformation, these interface dislocations glide out step by step,
leaving slip steps at the surface behind which constitute extrusions, as shown in Fig. 3b. The EGM model
predicts the initial rapid growth of these so-called static extrusions of height e. The extruded volume matches
precisely the volume of vacancy-type defects that accumulate in cyclic saturation. The local vacancy defect
concentration
sat
v
C
in the dislocation walls (volume fraction fw 0.1) of the ladder structure is established as
the result of a balance between annihilations of very close (vacancy) edge dipoles and formation of edge
dipoles.
sat
v
C
is estimated as 610-3. Slipping out of so-called PSB-matrix interface dislocations leads to a
one-dimensional elongation of the PSB lamella in the direction of the Burgers vector b, creating one static
extrusion of height e on each side of the crystal:
e 0.5
sat
v
C f
w
310
-4
D, (4)
where D is the specimen diameter measured in the direction of the Burgers vector. Thus, it is predicted that
the extrusion heights will be the larger, the thicker the fatigued single crystal is. In the case of a polycrystal, D
would represent the grain size, and the predicted height of an extrusion in a surface grain would be twice as
large as according to eq. (4), because only one extrusion would form. Nonetheless, extrusions in polycrystals
should be much smaller than in single crystals because the grain size is generally much smaller than the
specimen diameter. The described purely mechanical material extrusion process can operate in particular at low
temperatures, as had been observed, compare [3,13,14]. Continued cyclic deformation will lead to a gradual
surface roughening of extruded PSB material, see Fig. 3c, which can be estimated semi-quantitatively [3,12,13]
and which should be similar in polycrystals and in single crystals. Stage I shear cracks are initiated at the PSB-
matrix interfaces and/or in locations of stress enhancement in the roughened extrusion surface.
Since the EGM model deals almost exclusively with the formation of extrusions and not of intrusions, some
clarifying comments on the role of intrusions are in place. Before the advent of TEM and before high-
resolution SEM techniques had become available, the very early models of the surface profiles of PSBs were
constructed to predict so-called extrusion-intrusion pairs, as summarized in [13]. Today, it is clear these
views are no longer tenable. Thus, the EGM model essentially predicts only extrusions. Intrusion-like
deepenings which can be regarded as microcracks are considered to arise as a consequence of the extrusions
that develop first. Polk, in an attempt to extend the EGM-model [17], considered also the migration of point
defects from the PSB band to the surrounding matrix. A more detailed discussion of the model and its
consequences is found in the recent review by Man et al. [18,19]. Essentially, the Polk model then predicts, in
addition to the extrusions forming at the surface in the centre of the PSB, intrusions at both PSB-matrix
interfaces. Man et al. have studied the PSB surface profiles, mainly on fatigued polycrystalline stainless steel,
in great detail, using in particular atomic force microscopy (AFM), as summarized in [19]. While it cannot be
denied that the authors succeeded to reveal many previously unknown details, it is noted, in passing, that their
observations would have been even more informative had the studies been performed on monocrystalline
material, in which the effects would have been significantly larger. Nonetheless, the author maintains that,
most intrusions revealed by Man et al. were in fact embryonic cracks. The reasons are that the bulk of the
experimental observations indicate clearly that the dominant effects observed are as follows. 1) At the very
beginning, bulgy extrusions form. 2) Only later, intrusion-like deepenings, described as very fine and thin in
comparison with extrusions [19] are observed to develop. 3) They are not formed on both sides of the
extrusion but at the side of the extrusion where the emerging active slip plane is inclined to the surface at an
acute angle [19]. This corresponds to the PSB-matrix interface at which crack initiation is expected to occur
preferentially [12,15,16].
These experimental facts suggest strongly that the intrusions are much less dominant than the extrusions
and that they do not develop by a mechanism comparable to that of extrusion formation, as proposed by Polk.
641 Hael Mughrabi / Procedia Engineering 55 ( 2013 ) 636 644
Rather, much of the experimental evidence suggests that intrusions develop as a consequence of the
previously formed extrusions and are in fact embryonic stage I shear cracks.
5. UFG Materials
In recent years, bulk UFG (ultrafine grain size: below one micron) materials have attracted considerable
attention. These materials have attractive mechanical properties, in particular extraordinary high strength and
appreciable ductility. They are produced by severe plastic deformation techniques such as ECAP (equal
channel angular pressing), compare [20]. The fatigue properties of these materials are interesting n several
respects, compare [21-23]. In UFG copper, fatigue damage occurs in the form of macroscopic shear banding,
and cyclic softening accompanied by bimodal grain coarsening is observed, compare Fig. 4.

Fig. 4. Cyclic strain localization and shear banding in UFG copper fatigued at RT. Stress axis horizontal. Surface observations. a) Shear
bands in patches. b) Extended shear bands. c) TEM observation of locally coarsened grain/subgrain microstructure. From [22,23].
It is interesting that the fatigue lives of UFG materials, when plotted in HCF and UHCF Whler (S-N)
diagrams, are always larger than those of the conventional grain (CG) size counterparts, whereas the opposite is
true, when the fatigue life data are plotted in LCF Coffin-Manson fatigue life diagrams. The latter is believed to
be a consequence of the reduced ductility of the ECAP-processed material. This behaviour can be understood
in a total strain fatigue life diagram [22,23]. Considering UFG materials as strong and CG materials as
ductile materials, as shown in Fig. 5, it follows that the fatigue life curves of both classes of materials are
expected to intersect. Thus, the fatigue lives of UFG materials should be larger than those of CG materials in
the HCF regime (corresponding to the Whler-type S-N plot), but smaller in the LCF range (corresponding to
the Coffin-Manson plot). This prediction has been verified for UFG specimens of -brass (see Fig. 6),
aluminium and UFG copper [22].

Fig.5. Illustration of fatigue lives of strong UFG and more ductile CG materials in a total strain fatigue life diagram. From [23].
642 Hael Mughrabi / Procedia Engineering 55 ( 2013 ) 636 644
Fig. 6. Crossing of fatigue life curves of UFG a
The ductility of UFG materials can be enh
strength. Then a bimodal grain size distributio
(compare Fig. 4c), develops which, in the case
LCF range. Figure 7 shows fatigue life da
microstructures and, in addition, with a bimodal
different microstructures) are shown in a Whle
7b). In the first case, ECAP-processed UFG c
copper and CG copper. In the second case, one
by CG copper and, finally, UFG copper. The
fatigue-resistant than the CG copper in both cas
bimodally coarsened UFG copper exhibits hig
amplitudes investigated [24]. Such improveme
structures have so far not been reported for oth
bimodal UFG copper was more fatigue-resistant
strain fatigue life diagram, the bimodally coar
copper in the whole range of amplitudes inve
introduction of bimodal grain structures have so

Fig.7. Fatigue life data of commercial purity ECAP-proc
copper. a) Whler (S-N) plot, stress amplitude /2 vs. nu

pl
/2 vs. N
f
.From [24].
10
3
10
10
-3
10
-2
/2 = co
/2 = co

t
/
2
r
10
3
10
10
-3
10
-2
/2 = co
/2 = co

t
/
2
r

and CG -brass in total strain strain fatigue life diagram. After [22].
hanced by a mild annealing treatment at the expense of
on, similar to the one that evolves during cyclic deform
of copper, leads to a significant fatigue life enhancement
ata of commercial purity copper with CG and UFG
l grain distribution [24]. The data (lumped together for th
er-type S-N plot (Fig. 7a) and also in a Coffin-Manson plo
copper exhibits the longest fatigue lives, followed by bi
finds the longest fatigue lives for the bimodal copper, fo
e important result is that the bimodal UFG copper was
ses. As a consequence, in a total strain fatigue life diagra
gher fatigue lives than the CG copper in the whole ra
ens of the fatigue strength by introduction of bimodal
her UFG materials than copper. The important result is th
t than the CG copper in both cases. As a consequence, in
rsened UFG copper exhibits higher fatigue lives than th
estigated [24]. Such improvemens of the fatigue stren
far not been reported for other UFG materials than coppe

essed UFG copper, compared with bimodally coarsened UFG copper an
umber of cycles to failure, N
f
. b) Coffin-Manson plot, plastic strain amp
See text and original publication for details.
4
10
5
10
6
10
7
nst.,
pl
/2 = const., CG--brass, d 40 m
nst.,
pl
/2 = const., UFG--brass, d 750 nm
reversals to failure, 2 N
f
-brass
UFG, strong
CG, ductile
4
10
5
10
6
10
7
nst.,
pl
/2 = const., CG--brass, d 40 m
nst.,
pl
/2 = const., UFG--brass, d 750 nm
reversals to failure, 2 N
f
-brass
UFG, strong
CG, ductile
f some
mation
t in the
G grain
e three
ot (Fig.
imodal
llowed
s more
am, the
nge of
l grain
hat the
a total
he CG
gth by
er.
nd CG
plitude
643 Hael Mughrabi / Procedia Engineering 55 ( 2013 ) 636 644
6. Ultrahigh Cycle Fatigue (UHCF)
Laboratory fatigue tests are usually confined to some 107 cycles. However, in practice, there are many
examples of failures after much larger numbers of cycles, with Nf in the range of 109 cycles or more, e.g. in
vibrating or rotating machine components. Research in the UHCF regime has become an important topic,
referred to as UHCF or VHCF, since the International Conference Fatigue Life in Gigacycle Regime (later
called VHCF-1) in the year 1998. The problem at the heart of UHCF is whether a traditional fatigue limit in
HCF extrapolates into the UHCF range and poses the question whether a fatigue limit exists in the UHCF
regime.
Early research focused primarily on high-strength steels which did not show a fatigue limit beyond the HCF
regime, but failed by subsurface so-called fish-eye fractures in UHCF. The cracks initiated at internal
heterogeneities, in particular, non-metallic inclusions of ca. 10 m diameter [25,26]. Bathias [25] has explained
in a schematic diagram how the crack initiation frequency and sites vary from LCF to HCF and then UHCF,
compare Fig. 8. In the LCF range, many crack initiation sites develop at the surface. In the HCF regime,
surface damage still occurs but in a less severe manner. Thus the crack initiation sites are similar but less severe
and fewer. Finally, in the UHCF range, at even smaller stress amplitudes, surface damage becomes more or
less negligibly small. Then, cracks that initiate at internal heterogeneities can slowly grow to the surface and
become life-controlling.


Fig. 8. Schematic of shift of fatigue crack initiation sites from surface to interior [25].

Fig. 9. Schematic fatigue life diagram for LCF, HCF and UHCF ranges. After [26,27].
Nishijima and Kanazawa [26] have interpreted the HCF/UHCF fatigue life data in terms of a dual-stage
fatigue life diagram, based on the superposition of two separate fatigue life plots describing fatigue life for
surface and subsurface failure, respectively. Summarizing, the fatigue lives can then be represented
schematically over the whole range, from LCF to UHCF, as shown in Fig. 9 [27], with approximate boundaries
separating the regimes of HCF and UHCF, on the one hand, and surface and subsurface internal failures on the
other hand.
More recently, attention was also focused on UHCF of simple single-phase materials without internal
heterogeneities. In the case of copper fatigued ultrasonically for more than 10
10
cycles well below the
traditional PSB threshold, fatigue-induced surface roughening, cyclic strain localization in PSB-like bands and
644 Hael Mughrabi / Procedia Engineering 55 ( 2013 ) 636 644
stage I shear cracks were observed [28], as proposed in an earlier model of UHCF of ductile fcc metals [27],
compare Fig. 10. The question whether these cracks were non-propagating or whether they could ultimately
lead to failure is still under discussion. In Fig.10a, in a FIB section parallel to the stress axis, the fatigue-
induced surface roughness is clearly recognizable and can be evaluated quantitatively to yield a very low cyclic
slip irreversibility of just some 10
-4
[27]. However, the cumulative irreversible cyclic shear strain is
nonetheless very large (>30!!) because of the very high number of cycles. Fig.10b shows clearly that cyclic
strain localization has occurred as a consequence of the very high number of loading cycles in spite of the fact
that the loading amplitude was below the traditional PSB threshold.


Fig.10. Copper polycrystal fatigued at RT below the traditional ultrasonic PSB threshold to N = 1.5910
10
cycles, : stress axis. a) Surface
roughness profile and stage I crack initiation (FIB). b) Lamellae of slip localization (SEM). The stress axis is horizontal. From [28].
References
[1] A.Whler, Zeitschrift fr Bauwesen, 20(1870)73 .
[2] J.A.Ewing and J.C.W.Humfrey, Phil. Trans. Roy. Soc. 200(1903)241.
[3] H.Mughrabi: in Dislocations and Properties of Real Materials (Conf. Proc.), Book No. 323, The Institute of Metals, London, (1984),
p. 244
[4] L.F.Coffin, Trans. ASME 76(194)931.
[5] S.S.Manson, Nat. Advis. Comm. Aero., 1954, Technical Note 2933.
[6] J.D.Morrow, in ASTMSTP 378; Cyclic Plastic Strain Energy and Fatigue of Metals, American Society for Testing of Materials,
Philadelphia, (1965), p. 45.
[7] H.J.French, Trans. Am. Chem. Soc. Steel Treatment 21(1933)899.
[8] M.Klesnil, Metal Treatment and Drop Forging 55(1965)55.
[9] S.Suresh, Fatigue of Materials. Cambridge University Press, 2
nd
edition, 1998, p. 224.
[10] H.Mughrabi, in Proc. of 5
th
International Conference on Very High Cycle Fatigue (VHCF-5), (VHCF-5), Berlin, 2011, edited by
C.Berger and H.-J.Christ, Deutscher Verband fr Materialforschung und -prfung e.V. (DVM), 2011, p. 53.
[11] N.Thompson, N.J.Wadsworth, N.Louat,. Phil. Mag. 1(1956)113.
[12] U.Essmann, U.Gsele, H.Mughrabi, Phil. Mag. A 44(1981)405.
[13] H.Mughrabi, Metall. and Mater. Transactions A 40(2009)1257.
[14] Z.S.Basinski, S.J.Basinski, Progr. Mater. Sci. 36(1992)89.
[15] L.M.Brown, in Modelling of Material Behaviour and Design, edited by J.D.Embury and A.W.Thompson, TMS-AIME, 1990, p. 175.
[16] L.M.Brown and F.R.N.Nabarro, Phil. Mag. 84(2004)441.
[17] J.Polk, Mater. Sci. Eng. A 92(1987)71.
[18] J.Man, K.Obrtlik, J.Polk, Phil. Mag. 89(2009)1295.
[19] J.Man, P.Klapetek, O.Man, A.Weidner, K.Obrtlik, J.Polk, Phil. Mag. 89(2009)1337.
[20] R.Z.Valiev, R.K.Islamgaliev, I.V.Alexandrov, Progr. Mater. Sci. 45(2000)103.
[21] S.R.Agnew, A.Vinogradov, S.Hashimoto, J.R.Weertman, J. Electron. Mater. 28(1999)1038.
[22] H.Mughrabi, H.W.Hoeppel, Int. J. Fatigue 23(2010)1413.
[23] H.Mughrabi, H.W.Hppel, Mater. Res. Soc Symp. Proc. B2.1.1- B2.1.12(2001)634.
[24] M.Korn, R.Lapovok, A.Bhner, H.W.Hppel, and H.Mughrabi, Kovove Materialy 49(2011)1.
[25] C.Bathias, Int. J. Fatigue 23(2001)S143
[26] S.Nishijima and K.Kanazawa, Fatigue & Fract. Engng. Mater. Struct. 22(1999)601.
[27] H.Mughrabi, Int. J. Fatigue, 28(2006)1501.
[28] A.Weidner, D.Amberger, F.Pyczak, B.Schnbauer, S.Stanzl-Tschegg, H.Mughrabi, Int. J. Fatigue, 32(2010)872.
1 m 1 m
1 m
b
1 m 1 m
1 m
b
1 m 1 m
1 m
b
1 m 1 m
1 m
b

You might also like