You are on page 1of 14

Available online at www.sciencedirect.

com

Acta Materialia 58 (2010) 4039–4052


www.elsevier.com/locate/actamat

The evolution of crack-tip stresses during a fatigue overload event


A. Steuwer a,b, M. Rahman c, A. Shterenlikht d, M.E. Fitzpatrick c, L. Edwards c,e,
P.J. Withers f,*
a
ESS Scandinavia, University of Lund, 22350 Lund, Sweden
b
NMMU, Gardham Avenue, Port Elizabeth 6031, South Africa
c
Materials Engineering, The Open University, Walton Hall, Milton Keynes MK7 6AA, UK
d
Department of Mechanical Engineering, University of Bristol, Queen’s Building, Bristol BS8 1TR, UK
e
Institute of Materials Engineering, ANSTO, PMB1, Menai, NSW 2234, Australia
f
School of Materials, University of Manchester, Grosvenor Street, Manchester M1 7HS, UK

Received 26 February 2010; accepted 10 March 2010


Available online 10 April 2010

Abstract

The mechanisms responsible for the transient retardation or acceleration of fatigue crack growth subsequent to overloading are a
matter of intense debate. Plasticity-induced closure and residual stresses have often been invoked to explain these phenomena, but clo-
sure mechanisms are disputed, especially under conditions approximating to generalised plane strain. In this paper we exploit synchro-
tron radiation to report very high spatial resolution two-dimensional elastic strain and stress maps at maximum and minimum loading
measured under plane strain during a normal fatigue cycle, as well as during and after a 100% overload event, in ultra-fine grained
AA5091 aluminium alloy. These observations provide direct evidence of the material stress state in the vicinity of the crack-tip in thick
samples. Significant compressive residual stresses were found both in front of and behind the crack-tip immediately following the over-
load event. The effective stress intensity at the crack-tip was determined directly from the local stress field measured deep within the bulk
(plane strain) by comparison with linear elastic fracture mechanical theory. This agrees well with that nominally applied at maximum
load and 100% overload. After overload, however, the stress fields were not well described by classical K fields due to closure-related
residual stresses. Little evidence of overload closure was observed sometime after the overload event, in our case possibly because the
overload plastic zone was very small.
Crown Copyright Ó 2010 Published by Elsevier Ltd. on behalf of Acta Materialia Inc. All rights reserved.

Keywords: Plasticity-induced closure; Stress intensity factor; crack-tip stress field; Overload; Retardation

1. Introduction (i) Overloads retard, while underloads accelerate, the


rate of crack growth rate relative to the underlying
While most fatigue data is collected in a constant nom- constant amplitude rate.
inal crack-tip stress intensity range (DK = Kmax  Kmin), in (ii) A number of cycles (Nd) must be applied before the
practice it is commonplace for engineering components to original steady-state crack growth rate is re-estab-
experience a spectrum of loads. With this in mind, consid- lished following an over- or underload event.
erable research has been undertaken to characterise the (iii) The extent of the retardation effect depends on the
transient crack growth behaviour following individual overload ratio (OLR, KOmax/Kmax), the baseline value
over- or underload spikes in loading. The current state of of DK and the load ratio (R).
knowledge can be summarised thus [1]. (iv) Overloads can produce a very short initial accelera-
tion phase prior to prolonged retardation.

* These effects have previously been explained in terms of


Corresponding author.
E-mail address: philip.withers@manchester.ac.uk (P.J. Withers). plasticity-induced closure, crack-tip blunting, residual

1359-6454/$36.00 Crown Copyright Ó 2010 Published by Elsevier Ltd. on behalf of Acta Materialia Inc. All rights reserved.
doi:10.1016/j.actamat.2010.03.013
4040 A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052

Fig. 1. Schematic diagram of Newman’s strip model at KIMax and KIMin illustrating how plasticity-induced closure gives rise to a distribution of stresses
along the crack plane (after [11]), where a is the constraint factor and q = r the radius of the plastic zone.

compressive stresses and crack-tip branching or deflection. stress. Contrary to earlier work [7], Sadananda et al. [1] dis-
Plasticity-induced closure explanations have existed almost missed plasticity-induced crack closure, arguing that plas-
as long as crack closure has been proposed as a crack ticity always acts to open rather than close the crack.
growth retardation mechanism [2]. The argument is based The main problem with the crack-tip blunting explana-
on the fact that the overload creates a larger stretch in tion is that while it can reduce the stress intensity at the
the wake of the fatigue crack, which causes crack closure crack-tip, it cannot easily explain delayed retardation. It
as the crack-tip progresses through the overload plastic has been argued that reverse yielding ahead of the crack-
zone. Since the crack must first progress into the overload tip increases the size and magnitude of the compressive
plastic zone, this would explain why retardation is delayed. residual stress zone, thereby retarding crack growth. How-
Because of the larger plastic zone and the capacity for ever, the largest residual stresses will occur in the immediate
inward movement of material at the crack-tip under plane vicinity of the crack-tip, so it is difficult to explain delayed
stress, one might expect plasticity-induced closure to be retardation [8]. Further, the extent of the retardation phase
more significant in plane stress than generalised plane seems to be larger than the likely extent of the compressive
strain [3]. Under plane stress a potential mechanism of zone [9]. Especially for alloys with a propensity for planar
material transfer is obvious. Since out-of-plane deforma- slip, tensile overloads can promote crack deflection away
tion is not constrained, material can be transferred from from mode I (pure crack-opening). As a result, there is a
the thickness direction to the loading direction. However, short burst of accelerated growth, followed by retardation
the mechanism of material transfer postulated for plane due to the lower DKeff. However, such mechanisms are
stress is not admissible for plane strain. By definition, no not generally observed at the crack-tip, while overload
net out-of-plane contraction can occur and, therefore, it retardation is widely observed.
has been suggested that there can be no net axial stretch Regarding residual stress arguments, simple models
of material in the plastic wake behind the crack-tip, which such as Dugdale’s strip model [10] have been developed
implies no plasticity-induced crack closure [4]. The exis- to explain the evolution of crack-opening stress. These
tence of plasticity-induced crack closure under plane strain assume that crack-tip plasticity occurs in thin strips, with
conditions has thus been a topic of intense debate [5]. Fleck crack-tip plastic displacements and crack advance calcu-
and Newman have shown that closure does not occur for a lated step-wise by considering the region broken down into
bend specimen under plane strain conditions, while it does strips to simulate formation of the plastic wake. The tri-
occur for the middle crack tension geometry under plane axial constraints at the crack-tip are introduced into such
strain [6]. This may be due to the fact that the latter has models using a constraint factor (for which ideal values
a compressive T stress, while the former has a tensile T are 1 or 3 for plane stress and plane strain, respectively).
A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052 4041

The approach of Beretta and Carboni [11] is shown sche- tion was Al, 4.0% Mg, 1.2% Li, 1.0% C, 0.5% O and
matically in Fig. 1 where ac, at and aw are the constraint was produced from powders using a mechanical alloying
factors for compressive and tensile yielding and for com- process followed by hot isostatic pressing [22,23]. The
pressive yielding of the wake, respectively. billet was supplied as an upset-forged 42 mm thick plate
To date, experimental measurements of crack closure in the as forged (T1) condition. The material exhibited a
have relied mostly on either: (i) measuring some secondary yield stress of 450 MPa and a tensile strength of
property of the cracked body, such as compliance or electri- 505 MPa, values similar to those previously reported
cal resistance, or (ii) measurement of crack opening displace- for AA5091 in the T1 condition [24,25]. The tensile per-
ments on the surface of the cracked body. There are formance of Al–Mg–Li–C alloys is derived from signifi-
significant difficulties in interpreting secondary data in terms cant Hall–Petch strengthening due to their very fine
of crack-tip stresses, as the surface of a cracked body experi- grain structure, as well as solution and dispersion hard-
ences conditions of plane stress, in contrast to the bulk which ening. They typically possess a high Mg concentration
experiences conditions more akin to plane strain. in comparison with heat treatable Al–Li alloys, but the
In summary, the mechanisms of overload fatigue crack Li concentration is deliberately kept below 1.5%, to pre-
retardation are still the subject of debate. Some of the mech- vent formation of the age hardening d0 phase (Al3Li).
anisms relate to events in the crack wake, some ahead of the During mechanical alloying surface layers of oxides on
crack-tip. Crack-tip residual stresses under plane stress have the powder particles are repeatedly grown, then fractured
been measured at the surface using X-rays [12]. While the and dispersed within the particles. The grains, which are
measurement of stress fields in the bulk of cracked samples typically smaller than 1 lm, contain significant distribu-
by neutron diffraction has been possible for some years tions of fine (50–100 nm) precipitates, which are believed
[13–15], the spatial resolution (1 mm) has been insufficient to be Al4C3, Al2O3 and MgO. The Al2O3 and MgO
to resolve the immediate crack-tip field. Recently, very nar- phases derive from fragmentation of the surface oxides
row, but intense, beams of hard X-rays have become avail- from the as received materials. The Al4C3 phase, in con-
able at third generation synchrotron X-ray sources. This trast, derives only from reaction between the aluminium
has opened up the way for much higher spatial resolutions, and the process control agent used to aid the sintering
enabling crack-tip strains to be evaluated [16–19]. Recently, process [22,26]. The ultrafine grains promote homoge-
elegant synchrotron X-ray diffraction (SXRD) experiments neous deformation and prevent crystallographic cracking.
have been directed towards overload phenomena on 4 mm This fine microstructure also endows the alloy with low
thick samples [20], showing evidence of the overload on levels of crack closure and yield stress-insensitive fatigue
crack-tip stresses for cracks showing subsequent retardation. crack propagation behaviour so that the properties of the
SXRD also enables the study of cracks in much thicker sam- as forged T1 materials are indistinguishable from those
ples [17,21], presenting a unique opportunity to non-destruc- of the fully heat-treated condition [27]. The very small
tively determine the crack-tip elastic strain field under plane grain size makes it an ideal material for a high spatial
strain, i.e. in both the plastic and K-dominant zones. As high resolution investigation by energy dispersive X-ray dif-
energy X-ray diffraction utilizes small diffraction angles the fraction, since it allows very narrow slits to be used while
very high (lateral) spatial resolution can usually only be still providing a good powder diffraction pattern.
achieved in two orthogonal directions within bulk speci-
mens. However, if sufficiently thick cracked specimens are
used then, in principle, the full crack-tip stress field can be 2.2. The specimens
determined, since the conditions at the centre of such speci-
mens approximate well to generalised plane strain. Three 12 mm thick compact tension test pieces were
This paper reports what we believe to be the first mea- used, as defined in the fatigue crack propagation standard
surements of the two-dimensional crack-tip stress field in ASTM E647. Fatigue cracks were initiated in all the spec-
the plane strain region of thick samples at a resolution of imens and grown under conditions of p constant nominal
tens of microns. In order to examine the extent to which stress intensity range
p (DK I = 6 MPa m, p
R = 0.1, with
residual stresses arise during conventional fatigue and sub- KIMax = 6.6 MPa m and KIMin = 0.6 MPa m) using an
sequent to an overload results are reported at various MTS servo-hydraulic testing system. The crack length
points within the cycle for baseline fatigue, as well as dur- was monitored with a traveling microscope and the compli-
ing, after and well beyond a 100% overload event. ance of the specimen was recorded using a back face strain
gauge. The three specimens were examined under the fol-
2. Experimental details lowing conditions (see also Table 1):

2.1. The material studied  For specimen S1 (F) the crack was grown to a length of
26 mm.
The material used in this study is a powder metallurgy  For specimen S2 (FO) the crack was grown to 27.7 mm
Al–Li alloy based on alloy AA5091 obtained from Aero- and then p a 100% overload corresponding to
space Metal Composites Ltd., UK. The alloy composi- 13.2 MPa m (2KIMax) was applied.
4042 A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052

Table 1
p
Stress intensity factors (in units of MPa m) as nominally applied from the
far field load and inferred at the crack-tip by fitting LEFM solutions to the
strain maps.
Specimen Fatigue/ Applied crack-tip crack-tip
overload K Nom
I K eff
I K eff
II

S1 F 6.6 6.11 (0.21) 0.33 (0.53)


S2 FO 13.2 13.3 (1.83) 0.71 (0.38)
S3 FOF 6.6 6.2 (0.46) 0.06 (1.19)
The uncertainties given in parentheses as 2  standard deviation (95%)
confidence limits.

 For specimen S3 (FOF) the crack was grown p to


26.6 mm, where a 100% overload (13.2 MPa m) was
applied, after which the crack was extended by a further
0.1 mm, resulting in a final crack length of 26.7 mm.

For crack-tip stress field mapping these were transferred


to the synchrotron X-ray beam line and loaded in situ
using a modified Hounsfield tensometer mounted vertically
(Fig. 2a). For each test specimen the load was applied to
the correct level using the back face strain and the strain
gauge calibration obtained previously in the laboratory.
Partly as a result of the very small grain size, crack mor-
phologies under mode I loading in this material are typi-
cally very straight with the low surface roughness giving
rise to very low levels of roughness-induced closure. It
should be noted that at KIMax and 2KIMax the static for-
ward tensile plastic zones under plane strain would be
expected to be approximated by ry  K 2I =6pr2y [28], such
that for this alloy ry is equal to 11 lm and 45 lm, for KIMax Fig. 2. (a) Schematic of the diffraction geometry showing a CT specimen
with the crack plane horizontal and the two detectors measuring two
and 2KIMax, respectively, while the reverse compressive
directions of strain; note the coordinate system adopted for exx and eyy. (b)
plastic zones would be approximately four times smaller. Typical spectrum collected on ID15A in energy dispersive mode, where the
graph shows the original data points (+ symbols), the best fit Gaussian
2.3. The diffraction geometry and experimental set-up refinement (fitted line) and, displaced below these, their amplified
difference curve.
The experimental set-up comprised two solid-state X-
ray detectors mounted at 2h = 5°, one horizontally and
one vertically off-set. This configuration allowed the mea- approximate location of the crack-tip had previously been
surement of two essentially perpendicular directions of determined optically on the surface of the test piece and set
strain simultaneously (Fig. 2a; cf. [19]). The beam apertures initially to (0, 0). The measurement positions are indicated
were fixed at 25 lm in the vertical and horizontal directions by crosses overlaid on the strain map in Fig. 3a. (See below
to define the lateral spatial resolutions some 8  better for a detailed explanation of this figure.) The reference
than the best that had been achieved previously [18]. For frame was maintained during in situ loading by a 0.5 mm
each detector the receiving slits near the detector and those diameter steel pin that was inserted at a specified location
near the sample were both set at 40 lm. At low scattering near the notch of the computer tomography (CT) specimen
angles the strain resolution was sub-optimal, however, and scanned before each map.
the use of an ultrafine grained material and the large Ewald A typical spectrum and its best-fit refinement (using
sphere at high photon energies, as well as the multiple peak GSAS, see below) in energy dispersive mode are shown
refinement, helped to maintain good diffraction conditions in Fig. 2b. The spectra were very clear with low back-
and good strain resolution. The detectors and receiving grounds and no indication of strong texture. Initially a
slits were aligned and focused using the concentric tip of Gaussian-type peak profile was used for the refinement,
a small 0.5 mm diameter steel pin. The alignment between but, as indicated by the difference curve in Fig. 2b, there
the detectors was precise to within 20 lm. was a Lorentzian contribution to the profile. In order to
The strains were measured on a map of 25 lm pitch access the more complex peak profiles available in GSAS,
within 1 mm2 of the crack-tip, and on a coarser 100 lm the experimental data were first converted to a pseudo-
pitch over 2  2 mm2 further from the crack-tip. The angular dispersive scale [29]. Fitting was then undertaken
A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052 4043

-1.0 -0.5 0.0 0.5 -1.0 -0.5 0.0 0.5

0.75 a b 0.75

0.50
Strain
0.50
(10-6)
0.25 0.25
y (mm)

2,400
0.00 0.00
2,100
-0.25 -0.25 1,800
-0.50 -0.50 1,500
-0.75 -0.75 1,200
Measured
900
600
0.75 c d 0.75
300
0.50 0.50
0
0.25 0.25
y (mm)

-300
0.00 0.00
-600
-0.25 -0.25
-900
-0.50 -0.50 -1,200
-0.75 -0.75
Fitted
-1.0 -0.5 0.0 0.5
-1.0 -0.5 0.0 0.5
x (mm) x (mm)

Fig. 3. The raw measured data (top) and best fit (bottom) elastic strain fields (106) for (a) and (c) the crack growth (exx) and (b) and (d) the crack opening
p
(eyy) directions at KIMax = 6.6 MPa m for S1. In (a) the crosses mark the coarse and fine scan measurement locations. The LEFM best fit crack-tip strains
p p
correspond to a crack-tip stress intensity of (K eff eff
IMax = 6.11 MPa m, K IMax = 0.33 MPa m).

using GSAS [30] in intensity extraction mode [31], which order of magnitude smaller than the point-to-point varia-
fits all peak positions to the required parameters, in our tion, which is due to material (microstructural) and instru-
case simply the lattice parameter. Typically only the region mental factors. Stress in any three orthogonal directions
50–150 keV (or rather its equivalent in angular dispersive can be calculated from the corresponding elastic strains
terms) was included in the refinements. The strain was then using the relationship:
calculated using the relative variation of the refined lattice  
E m
parameter (a) across the sample in the usual way, i.e. rij ¼ eij þ ðe11 þ e22 þ e33 Þ ð1Þ
ð1 þ mÞ ð1  2mÞ
e = (a  a0)/a0. This approach has the advantage that, by
including many peaks in the refinement, the measured Following the arguments summarised in Suresh [35], a
strains are likely to be representative of the macroscopic crack in a fracture mechanics specimen is likely to undergo
response, free from significant elastic or plastic anisotropy plane strain conditions if its thickness (B) is greater than
effects [32,33]. The unstrained lattice parameter (a0) was 2.5 (K/ry)2, which gives for test specimen S1 B > 0.53 mm
obtained from the averaged far field measurement in the and for specimen S2 containing the overload crack
unloaded state [34]. This value agreed well with that B > 2.15 mm. These thicknesses will be even smaller for
obtained by applying a stress balancing condition on a suit- fatigue and, thus, it is justified to calculate the stress field
able part of the data, e.g. in the crack opening direction in both specimens using a plane strain approximation.
along the crack plane to the back face. For any subsequent Here we take this to mean that the elastic strain is in a state
diffraction peak width analysis a dedicated single peak fit- of plane strain [ezz = 0, rzz = m(rxx + ryy)].
ting routine was employed.
The counting times were of the order of 120 s per point, 2.4. Determination of crack-tip stress intensity
being limited primarily by the low counting rate (40 kHz)
of the detectors rather than the diffracted flux. The strain The advent of surface strain/displacement mapping
uncertainty (the uncertainty in the lattice parameter refine- techniques [36,37] opened the way for the development of
ment) at each point was of the order of 105, being an algorithms for the determination of crack-tip stress inten-
4044 A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052

p
sity factors directly from the (total) strain [38–41] or from fatigued (FOF) specimen
p p = 0.6 MPa m,
(S3) at KIMin
the displacement [42] fields thus measured. The advantage KIMean = 3.6 MPa m, KIMax = 6.6 MPa m.
of these methods is that one can determine the stress inten-
sity experienced by the crack-tip as against that nominally 3.1. Baseline fatigue (S1)
applied; the disadvantage is that most displacement or
strain mapping methods are applicable only at the surface The elastic strain fields recorded for the baseline fatigue
of metals and local strains derived from displacements tend cracked specimen (S1) at KIMax are shown for the crack
to be rather noisy due to numerical differentiation. By cou- opening [eyy(x, y)] and crack growth [exx(x, y)] directions
pling the theoretical linear elastic fracture mechanics in Fig. 3a and b. The measured mid thickness strain fields
(LEFM) crack-tip strain field to the (essentially) elastic have been compared against the predicted LEFM crack-tip
strain fields measured at mid thickness by SXRD we have strain distributions to derive the best fit mode I and mode
been able to infer the effective crack-tip stress intensity II stress intensity factors. These are summarised in Table 1
under plane strain directly from the crack-tip stress field and the best fit elastic fields are shown in Fig. 3 (bottom).
to compare against that nominally applied. From Table 1 it can be seen that the inferred crack-tip
stress intensities are in excellent agreement with those nom-
3. Experimental procedure and results inally applied both at maximum load and for the overload
case. These results suggest that, as one would expect, any
Strain maps over the mid plane were collected at various prior residual stresses have a negligible effect on the
stages of loading.
p For the fatigued test specimen
p (S1) at crack-tip stress intensity at KIMax and KIOMax and that
KIMin = 0.6
p MPa m, K IMean = 3.6 MPa m, K IMax = the measured fields shown in Fig. 3 are essentially what
6.6 MPa m, for the 100% fatigued overloaded
p (FO) test one would expect on the basis of LEFM.
specimenp (S2) at KIMin = p0.6 MPa m, Kmean = The strain fields [eyy(x, y), and exx(x, y)] measured at
3.6 MPa m, KIMax
p = 6.6 MPa m and KIOMax = KIMax for S1 have been converted to stress using Eq. (1) in
2KIMax = 13.2 MPa m and for the fatigued overloaded combination with the plane-strain assumption and are

-1.0 -0.5 0.0 0.5 -1.0 -0.5 0.0 0.5

0.75 a b 0.75 Stress


0.50
(MPa)
0.50
375
0.25 0.25
350
y (mm)

0.00 325
0.00
300
-0.25 -0.25 275
250
-0.50 -0.50
225
-0.75 -0.75 200
175
150

0.75 c d 0.75 125


100
0.50 0.50
75

0.25 0.25 50
25
y (mm)

0.00 0.00
0

-0.25 -0.25 - 25
- 50
-0.50 -0.50 - 75

-0.75 -0.75 -100

-1.0 -0.5 0.0 0.5 -1.0 -0.5 0.0 0.5


x (mm) x (mm)

Fig. 4. Measured mid thickness (a) crack growth (rxx), (b) crack opening (ryy), (c) through thickness (rzz) direction and (d) hydrostatic (rH) stress fields
for S1 loaded to KIMax.
A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052 4045

shown in Fig. 4, akin to Fig. 3. It can be seen that the ‘butter- residual stresses to be recorded by a 25 lm gauge dimen-
fly’ lobes evident in the crack opening strain map in Fig. 3b sion, noting that for the previous plane stress case the plas-
are not present in the corresponding crack-opening stress tic zone would be larger, or because closure is inherently
map (Fig. 4b). Significant peak stresses were found for all difficult under plane strain.
three stress components immediately ahead of the crack, The calculated residual stresses at KIMin are shown in
creating a triaxial tensile field as expected [ryy(x, y) > Fig. 5. There is some evidence of a 50–100 lm wide compres-
rzz(x, y) > rxx(x, y)] ahead of a crack-tip in plane strain. sive stress across the wake of the crack (Fig. 5d). However,
Although the measured von Mises stress was only around this should be treated with caution given the observation
140 MPa, it should be noted that the plastic zone is expected by Croft et al. [18] of similar low level ‘compressive’ strains
to be somewhat smaller than the gauge volume, so that the on the crack surfaces, even after splitting the sample in half.
peak stresses recorded at the crack-tip will be depressed by In their case these could have resulted from intergranular
averaging over the gauge volume. strains arising from crack-tip plasticity associated with using
The elastic strain and stress fields recorded upon the single diffraction peak (3 2 1). For the (3 2 1) plane the
unloading from p KIMax to the minimum load observed sense (compressive) of intergranular strain in the
KIMin = 0.6 MPa m are plotted in Fig. 5 at somewhat crack opening direction was in line with the axial plastic ten-
higher magnification. sion and the broadened diffraction peak profile in this
It is clear from Fig. 5a and b that, as one might expect, region. Such an explanation is less likely here as we have used
the elastic strains in the sample are much smaller at KIMin. the ‘full pattern profile refinement’ to obtain strain and,
It is not possible to identify a distinct reverse plastic zone in hence, our measurements are less likely to be affected by type
the immediate vicinity of the crack-tip. Our results are in II intergranular strains. Alternatively, the ‘stresses’ could be
contrast to those recorded by Akiniwa et al. [43] using sur- due to a slight error in strain free lattice spacing, however,
face X-ray diffraction (plane stress) for fine grained steel, since the stress in the crack opening direction (Fig. 5c) shows
who observed a residual stress consistent with a marked no such feature it may be a real effect of low level contact
reverse plastic zone in accordance with closure models stress. The observation of tensile crack opening and crack
[44]. There are two possible explanations; either the reverse growth direction stresses after the crack position might con-
plastic zone introduced here is too small to create sufficient firm very low level closure.

-0.4 -0.3 -0.2 -0.1 0.0 -0.4 -0.3 -0.2 -0.1 0.0

0.20
a b 0.20
0.15 0.15 strain
x1E-6
0.10 0.10
600
0.05 0.05 400
400
y (mm)

0.00 200
200
0.00
00
-0.05 -0.05 -200
-0.10 -0.10 -400
-600
-0.15 -0.15
-800
-0.20 -0.20
-0.25 -0.25

0.20
c d 0.20
0.15 0.15
stress
0.10 0.10 MPa

0.05 0.05 50
50
y (mm)

25
25
0.00 0.00 00
-0.05 -0.05 -25
-50
-0.10 -0.10
-75
-0.15 -0.15 -100

-0.20 -0.20
-0.25 -0.25
-0.4 -0.3 -0.2 -0.1 0.0 -0.4 -0.3 -0.2 -0.1 0.0
x (mm) x (mm)
p
Fig. 5. Close-up of the mid thickness elastic strain and stress fields at KIMin = 0.6 MPa m for S1. The elastic strains (106) in the (a) crack-growth (exx)
and (b) crack opening (eyy) directions. Inferred stresses in (c) the crack growth (rxx) and (d) the crack opening (ryy) directions. It should be remembered
that the strains were derived by multiple peak refinements.
4046 A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052

-0.5 0.0 0.5 -0.5 0.0 0.5


1.00 1.00
a 0.75
0.75

0.50 0.50

0.25 0.25
y (mm)

0.00 0.00

-0.25 -0.25
Stress
-0.50 -0.50
(MPa)
-0.75 -0.75
800
700
1.00 1.00
b 600
0.75 0.75
500
0.50 0.50 400
0.25 0.25 300
y (mm)

0.00
200
0.00
100
-0.25 -0.25
0
-0.50 -0.50
-100
-0.75 -0.75 -200
-300
x (mm) x (mm) -400
1.0 1.0
c
0.8 0.8

0.6 0.6

0.4 0.4
y (mm)

0.2 0.2

0.0 0.0

-0.2 -0.2

-0.4 -0.4

-0.6 -0.6

-0.5 0.0 0.5 1.0 -0.5 0.0 0.5 1.0


x (mm) x (mm)

Fig. 6. Maps (in MPa) of the crack growth direction stress (rxx) (LHS) and crack-opening stress (ryy) (RHS) around the crack-tip of the FO specimen (S2)
p p p
at (a) 2KIMax = 13.2 MPa m, (b) KIMean = 3.6 MPa m, (c) KIMin = 0.6 MPa m.

3.2. Overload case (S2) mate strengths for the alloy, as would be expected, the
maximum von Mises stress (370 MPa) was slightly less
The crack-opening stress fields for the fatigued andpthen than the yield stress due to the hydrostatic stress
100% overloadedp crack (FO) at 2KIMax p (13.2 MPa m), component.
KIMean (3.6 MPa m) and KIMin (0.6 MPa m) are shown As the load was reduced from the overload to the base-
in Fig. 6. The stress field at 2KIMax is clearly much more line mean stress, KIMean, it is clear that the field deviates sig-
intense than that shown in Fig. 3 for KIMax. Indeed, the nificantly from a simple ‘KI’ field. This is easily discerned
stress intensity representative of the crack-tip stress field from the compressive peak stress (ryy) at, or just behind,
at 2KIMax was determined directly from the crack-tip strain the crack-tip in Fig. 6b. This is in good agreement with
field and the inferred stress intensity factor was in excellent the behaviour sketched out in Fig. 1 and is more easily vis-
agreement with that nominally applied (see Table 1). Fur- ualised in the line graphs of Fig. 7. When considering this
ther, it is clear that the effective stress intensity was found plot it should be noted that the far field stresses under
with a good level of uncertainty. While the peak stress KIMin in both directions are slightly compressive, as is the
was considerably larger than both the typical yield or ulti- crack-opening stress at 2KIMax on the crack faces, suggest-
A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052 4047

1000

800 K=16.5 a K=16.5 b


K=3.6 K=3.6
600
K=0.6 K=0.6

Stress (MPa)
400

200

-200

-400

-600
-1.2 -0.8 -0.4 0 0.4 0.8 1.2 -0.8 -0.4 0 0.4 0.8 1.2
x (mm) x (mm)

Fig. 7. Variation in (a) crack growth stress (rxx) and (b) crack-opening stress (ryy) along the crack plane (at y = 0) around the crack-tip of the FO
p p p
specimen (S2) for 2KIMax = 13.2 MPa m, KIMean = 3.6 MPa m and KIMin = 0.6 MPa m. It is possible that overall the measured stresses are 50 MPa
too compressive (see text).

ing that the estimated stress free lattice parameter is prob- KIMin. It would appear that the compressive zone behind
ably slightly too large and that the stresses may be the tensile peak due to the combined action of the plastic
50 MPa too compressive. Provided there had been no wake and the reversed plastic zone (Fig. 1) fails to maintain
movement of the sample during unloading, Fig. 7 confirms a significant tensile crack-tip stress intensity (even given the
the presence of a significantly enhanced reverse plastic zone likely 50 MPa underestimation of the tensile stress), as the
ahead of the crack, as well as a small region of crack face peak tensile stress was not that much different to that at
contact behind the crack-tip, even at KIMean. Unfortu- Kmean for S1 measured during baseline fatigue. The tensile
nately, beam time constraints meant that it was not possi- peak fell further on unloading to KIMin. There is some evi-
ble to measure the evolution of the stress field over many dence that the stress ahead of the reverse plastic zone is
load steps during the unloading cycle and so it is not clear actually tensile (50 MPa) rather than zero, as it appears
at what applied stress a compressive stress in the crack-tip to be in Fig. 7, in view of the difficulty of accurately mea-
region began to develop, but it would appear to be at a suring the unstrained lattice parameter. The form of the
stress considerably larger than the baseline mean stress stress field was broadly in agreement with Dugdale-type
KIMean (in simple terms KIclosure > KIMean). Due to plastic calculations, as well as the finite element calculations
stretch within the plastic zone the maximum tensile stress reported by Wang et al. [45]. Furthermore, our overload
appears to be pushed some distance (100 lm) ahead of residual stress results in Fig. 6c are broadly similar to their
the crack-tip (cf. Fig. 1). This is slightly larger than the pre- predictions, taking into account growth of the fatigue
dicted extent of the overload forward plastic zone (rOy crack, summarised in Fig. 13 of their paper. Their work
60 lm) and considerably larger than the estimated reverse suggested that the reverse plastic zone is more important
plastic zone. The maximum tensile stress continues to fall than crack face closure in influencing crack growth. Recent
as the load is removed such that it is relatively small at work by Croft et al. [20] for thinner samples (more prone to

detector
channels
-1.0 -0.5 0.0 0.5 -0.5 0.0 0.5
Distance from crack-tip (y) [mm]

1.00 35
33
0.75
a b 0.75
20
31
0.50 0.50 19
29
0.25 0.25 18
27
0.00 0.00 17
25

-0.25 16
23
-0.25
15
21
-0.50 -0.50
19
-0.75 -0.75 17
15
-1.0 -0.5 0.0 0.5 -0.5 0.0 0.5
Distance from crack-tip (x) [mm] Distance from crack-tip (x) [mm]
p
Fig. 8. Maps of the Al (3 1 1) diffraction peak width (in units of detector channels) for (a) S1 at KI = 6.6 MPa m for lattice planes normal to the x-
p
direction and (b) S2 at KI = 13.2 MPa m for lattice planes normal to the crack opening y-direction. (Note that the type III contribution to the peak width
would not be expected to vary with strain measurement direction.)
4048 A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052

crack-tip plasticity) is in striking agreement with the results One way of examining the extent of the plastic zone, as
presented here. They observed a similar triangular profile well as the plastic wake, is to examine the diffraction peak
in the region of the crack-tip at KIMin and a compressive widths. Peak width is usually very sensitive to line broaden-
lobe constrained by a tensile region ahead of the crack- ing by plastic work, although it can also be affected by any
tip and a more striking one than we observed behind it. sharp gradient of elastic strains present within the gauge (as
Unfortunately, they did not monitor the unloading might occur around a crack-tip). The diffraction peak
response after overload, but they did monitor the loading widths of the (single peak) fitted Al (311) reflection at
response for the crack-tip stress file once it had propagated 2KIMax for S2 and KIMax for S1 are compared in Fig. 8.
further; which shows striking similarities to the present In agreement with the earlier estimation of the likely plastic
work. zone size, it is clear that it was not possible using a 25 lm

Kmin Kmax

0.75 a
0.50

0.25
y (mm)

0.00

-0.25 Stress
(MPa)
-0.50
350
-0.75
300
250
b 200
0.75
150
0.50
100
0.25 50
y (mm)

0.00 0
- 50
-0.25
-100
-0.50
-150
-0.75

.00

0.75 c
0.50

0.25
y (mm)

0.00

-0.25

-0.50

-0.75

-1.0 -0.5 0.0 0.5 -1.0 -0.5 0.0 0.5


x (mm) x (mm)
p p
Fig. 9. Stress maps around the crack-tip of the FOF specimen (S3) at KIMin = 0.6 MPa m (LHS) and KIMax = 6.6 MPa m (RHS) for the (a) crack
growth (rxx), (b) crack opening direction (ryy) and (c) normal stress (rzz) directions.
A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052 4049

gauge dimension to resolve the plastic zone at KIMax. Nor Comparing the stress fields at KIMax between S1 (fatigued
was there any significant evidence of a plastic zone gener- only) and S3 (FOF) in Figs. 4 and 9, respectively, the stress
ated in the crack wake as the crack had grown. In contrast, fields appear to be similar before and after the overload
broadening of the diffraction peak local to the crack-tip for cycle in both cases. Indeed, the crack-tip stress intensity
p
the overloaded sample (S2) is clearly evident in Fig. 8b, inferred from the full field strain map was 6.2 MPa m
despite the fact that energy dispersive diffraction typically (Table 1), which is in good correspondence with that nom-
has moderate instrumental resolution (peak width) com- inally applied, as well as with that observed at the crack-tip
pared with angular dispersive measurements [46]. This for baseline fatigue. This high level of correspondence was
broadening may in part be a consequence of the steep elas- emphasised by direct comparison of the two crack-tip
tic gradients found there (see also the discussion relating to plane profiles in Fig. 10a and b, where the profiles have
Fig. 11), but it is likely that plasticity has also paid a role. been displaced relative to one another to bring the crack
The peak in the diffraction peak width profile appears to locations into coincidence. The crack-opening stress at
have an extent of around 200 lm – considerably larger than KIMean for the FO (S2) and FOF (S3) cases are shown over-
the predicted plastic zone (45 lm), but comparable with laid in Fig. 10c. The effect of the overload cycle was marked
the reversed plastic zone inferred from Fig. 7 by comparing immediately after unloading (FO). In contrast to Croft
it with the schematic in Fig. 1. et al. [20], who observed a crack in a thinner sample, we
did not see a significant change in the baseline fatigue resid-
3.3. Baseline fatigue cycle sometime after overload (S3) ual stress field soon after overloading, nor any residual evi-
dence of the overload feature. This may be because the
Croft et al. [20] found that the effect of the overloaded plastic event created by the overload in our plane strain
plastic zone (which in their case was much larger) was case was very small. Results at Kmean for the FOF specimen
retained as the crack was grown through the retardation S3 (Fig. 10c and d) look very similar to that for baseline
zone and concluded that this was important in terms of fatigue, suggesting that the overload had not influenced
retarding crack growth rate. In our experiment we exam- the crack-tip stress fields, at least above the nominal Kmean
ined the crack-tip stress field after the crack had extended loading level. Furthermore, there appears no evidence that
a further 0.1 mm beyond the 100% (2KIMax) overload, i.e. there was any residual effect 0.1 mm behind the crack-tip in
just beyond the plastic zone arising from the overload. the stress maps of Fig. 9.
The crack-tip stress fields after the crack had grown To see if there was any evidence of the overload event,
0.1 mm beyond the 100% overload are shown in Fig. 9. the diffraction peak width maps for sample S3 (FOF) are

500

400
Kmax F a F b
FOF FOF
300
Stress (MPa)

200

100

- 100

- 200
300
Kmean FO
c FO
d
200
FOF FOF
Stress (MPa)

100

- 100

- 200

- 300
-1.2 -0.8 -0.4 0 0.4 0.8 1.2 -0.8 -0.4 0 0.4 0.8 1.2
x (mm)
x (mm)

Fig. 10. A comparison of the stresses for the F (S1), FO (S2) and FOF (S3) samples, each with the current crack position at x = 0 mm for the (a) crack
growth and (b) crack opening stresses at Kmax and the (c) crack growth and (d) crack opening stresses at Kmean.
4050 A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052

-1.0 -0.5 0.0 0.5

0.6 a 0.6
0.4 0.4
0.2 0.2
0.0 0.0

y (mm)
-0.2 -0.2
-0.4 -0.4 detector
channels
-0.6 -0.6
-0.8 -0.8 21.5
21
-1.0 -1.0
20.5
0.75 0.75
b 20

0.50 0.50 19.5


19
0.25 0.25
18.5
y (mm)

0.00 0.00
18
-0.25 -0.25 17.5
-0.50 -0.50

-0.75 -0.75

-1.00 -1.00
-1.0 -0.5 0.0 0.5
x (mm)

-0.3 -0.2 -0.1 0.0 0.1

0.05
12
c 0.05
02
0.00 0.00

-0.05 -0.05
19.5
y (mm)

1
5. 9
.5

-0.10 -0.10
19

-0.15 19 -0.15
19 91
-0.20 -0.20

-0.25 -0.25
19

19
-0.30 -0.30
-0.3 -0.2 -0.1 0.0 0.1
x (mm)
p p
Fig. 11. The width of the Al (3 1 1) diffraction peak (in channel numbers) for S3 (FOF) for: (a) KIMax = 6.6 MPa m, (b) KIMean = 3.6 MPa m and (c)
p
KIMin = 0.6 MPa m. The overload was applied when the crack was 100 lm behind the current location of the crack.

shown for KIMax, KIMean and KIMin in Fig. 11. A broaden- vicinity of a fatigue crack under plane strain conditions
ing of the peak width is clear at the crack-tip location at all at much higher spatial resolution than has been achieved
three load levels, but it is less than observed for the over- before [17]. From these measurements it has been possible
load case shown in Fig. 8. Further, the peak width at the to map the crack opening, transverse and normal stresses
crack-tip decreased with unloading, perhaps indicating that using an assumption of plane strain. Furthermore, because
its origin lay in the steep stress gradients there. In any a large part of the diffraction profile had been recorded at
event, evidence of the overload event as well as other crack each point, it was possible to use a Rietveld-style refine-
wake peak broadening become increasingly evident with ment so that the elastic strains determined are likely to
decreasing load. In particular, at KIMin broadening became be representative of bulk elastic strains/stresses. Addition-
significant at 100 lm behind the crack-tip (i.e. at the ally, by comparing the measured strain field with that pre-
overload location). Further, there is some indication of dicted by theory it has been possible, for the first time, to
some residual broadening event 0.65 mm behind the infer the crack-tip stress intensity right at the crack-tip
crack-tip, perhaps suggestive of crack face contact there to. within the bulk under conditions of plane strain.
p
During baseline fatigue (KIMax = 6.6 MPa m) the elas-
4. Conclusions tic strain field local to the crack was indicative of a stress
intensity very close to that nominally applied recording a
This paper presents two-dimensional maps of the elastic peak tensile stress (averaged over the 25 lm gauge volume)
crack-opening and crack growth direction strains in the of 450 MPa. Upon unloading, however, no significant
A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052 4051

residual stress was found at KIMin, perhaps because of the presence or absence of plasticity-induced closure in systems
difficulty of closure under plane strain or, alternatively, with a larger plastic zone. This will be the target of further
because the expected reverse plastic strain/zone size was work carried out deep within thick plane strain samples in
very small and therefore their effects were not detected. order to assess whether plasticity-induced closure events
As at KIMax, the strain field local to the crack-tip at can be significant under plane strain conditions.
100% overload was consistent p with the nominally applied
stress intensity (13.2 MPa m), i.e. the crack-tip was Acknowledgements
loaded to the nominal stress intensity. During unloading
compressive residual stresses were measured just in front These experiments were undertaken in 2005 at the ESRF
of the crack indicative of a significant reverse plastic zone, through the long-term project HS2252. We are grateful for
as well as some evidence of a small region of crack closure the support of staff on the beamline ID15A for experimen-
behind the crack-tip due to crack face contact. The broad tal support. M.E.F. is supported by a Grant through The
profile of these compressive strains compared well with Open University from the Lloyd’s Register Educational
those observed by Croft et al. [20] on thinner samples. Trust. P.J.W. is grateful for helpful discussions with Prof.
The peak compressive stress was around 450 MPa. David Nowell.
Slightly further ahead of the crack (around 100 lm) tensile
stresses were evident at KIMin, being maintained at no more
References
than 100 MPa. It was not possible to infer from our
unloading experiments the load at which the crack-tip [1] Sadananda K, Vasudevan AK, Holtz RL, Lee EU. Int J Fatigue
stress became compressive, but significant compressive 1999;21:S233–46.
stresses were present even at KIMean, due largely to reverse [2] Elber W. Eng Fract Mech 1970;2:37–45.
plastic yielding. In this context it should be noted that our [3] Alizadeh H, Hills DA, de Matos PFP, Nowell D, Pavier MJ, Paynter
RJ, et al. Int J Fatigue 2007;29(2):222–31.
residual stress curves were similar in form to the surface
[4] Fleck NA. Eng Fract Mech 1986;25(4):441–9.
stress study of overload made by Jagg and Scholtes [47], [5] Solanki K, Daniewicz SR, Newman JC. Eng Fract Mech
except that the effects occurred on a finer scale. This is, 2004;71(2):149–71.
of course, in part due to the fact that they studied plane [6] Fleck NA, Newman JC. In: Mechanics of fatigue crack closure, STP
stress, where plasticity-induced closure is much more dra- 982. Philadelphia, PA: ASTM; 1988. p. 319–41.
[7] Reynolds AP. Fatigue Fract Eng Mater Struct 1992;15:551–62.
matic and partly because they applied a higher maximum
[8] Fleck NA. In: Basic questions in fatigue, STP 924. Philadelphia,
stress intensity. After crack growth under baseline fatigue PA: ASTM; 1988. p. 157–83.
to 0.1 mm, beyond the overload event, the stress intensity [9] Suresh S. Eng Fract Mech 1983;18:577–93.
at KIMax was indistinguishable from that prior to the over- [10] Dugdale DS. J Mech Phys Solids 1960;8:100–4.
load event. [11] Beretta S, Carboni M. Eng Fract Mech 2005;72(8):1222–37.
[12] Allison JE. In: Fracture mechanics, STP 677. Philadelphia,
In all cases the stress in the wake of the crack appeared
PA: ASTM; 1979. p. 550–62.
to be compressive – this may be a valid observation or it [13] Allen AJ, Hutchings MT, Windsor CG, Andreani C. Adv Phys
may reflect an incorrect choice of the strain free lattice 1985;34(4):445–73.
parameter. Croft et al. [18] found a similar effect using a [14] Hutchings MT, Hippsley CA, Rainey V. Mater Res Soc Sympos Proc
single peak analysis. We tend to agree with the authors that 1990;166:317.
[15] Korsunsky AM, Fitzpatrick ME, Withers PJ. In: 10th Int conf on
in their case this was due to anisotropy arising from inter-
composite materials, vol. 1. Whistler: Vancouver; 1995. p. 545–552.
granular strains created by differential plastic flow between [16] James MN, Hattingh DG, Hughes DJ, Wei LW, Patterson EA, Da
differently oriented grains. It is rather more difficult to Fonseca JQ. Fatigue Fract Eng Mater Struct 2004;27(7):609–22.
explain in our case, because a Rietveld multiple peak anal- [17] Steuwer A, Santisteban JR, Turski M, Withers PJ, Buslaps T. J Appl
ysis was used. If the effect is due to an error in the strain Crystallogr 2004;37:883–9.
[18] Croft M, Zhong Z, Jisrawi N, Zakharchenko I, Holtz RL, Skaritka J,
free lattice spacing then the stresses reported here are
et al. Int J Fatigue 2005;27(10–12):1408–19.
50 MPa too compressive. Others have seen similar effects, [19] Steuwer A, Edwards L, Praithar S, Ganguly S, Peel M, Fitzpatrick
if at lower spatial resolution [48]. Further work needs to be ME, et al. Nucl Instrum Methods Phys Res 2006;246:217–25.
done both immediately after overload and after continued [20] Croft M et al. Int J Fatigue 2009;31:1669–77.
growth of the crack as it advances through the overload [21] Turski M, Bouchard PJ, Steuwer A, Withers PJ. Acta Mater
2008;56(14):3598–612.
plastic zone to establish whether crack closure or other
[22] Vine WJ, Pitcher PD, Tarrant AD. Mater Sci Technol 2001;17:802–6.
influences arising from the overload event can explain fati- [23] Pitcher PD, Bushby RS, Vine WJ, Smith AF, Tarrant AD. Mater Sci
gue crack retardation under plane strain conditions, requir- Technol 2001;17:807–14.
ing significant amounts of beam time. A great deal has [24] Bray GH, Reynolds AP, Starke EA. Metall Trans A 1992;23:3055–66.
already been learnt from the study of Croft et al. [18,20] [25] Bray GH, Kaisand LR, Starke EA. Fatigue Fract Eng Mater Struct
1995;18:551–64.
on thinner (plane stress) samples.
[26] Vine WJ, Pitcher PD, Tarrant AD. Mater Sci Technol
Finally, it should be noted that in our case the forward 2001;17:645–50.
and reverse plastic zones were very small and difficult to [27] Rao KTV, Ritchie RO. Metall Trans A 1991;22:191–202.
detect, even using the excellent spatial resolution achieved [28] Rice JR. In: Fatigue crack propagation. Philadelphia, PA: ASTM;
here (25 lm). It would certainly be easier to confirm the 1967.
4052 A. Steuwer et al. / Acta Materialia 58 (2010) 4039–4052

[29] Steuwer A, Santisteban JR, Turski M, Withers PJ, Buslaps T. Nucl [38] Stanley P, Chan WK. In: Proc I Conf Fat Eng Mater Struct. Lon-
Instrum Methods Phys Res B 2005;238B:200–4. don: Institution of Mechanical Engineers; 1986.
[30] Larson AC, Von Dreele RB. GSAS: generalized structure analysis [39] Shterenlikht A, Garrido FAD, Crespo PL, Withers PJ, Patterson EA.
system. Los Alamos, NM: Los Alamos National Laboratory; 1985. p. Adv Exp Mech 2004;1(2):107–12.
151–152. [40] Hild F, Roux S. C R Mecanique 2006;334:8–12.
[31] Lebail A, Duroy H, Fourquet JL. Mater Res Bull 1988;23(3):447–52. [41] Yoneyama S, Morimoto Y, Takashi M. Strain 2006;42:21–9.
[32] Daymond MR, Bourke MAM, Von Dreele RB, Clausen B, Lorentzen [42] Lopez-Crespo P, Shterenlikht A, Patterson EA, Yates JR, Withers PJ.
T. J Appl Phys 1997;82(4):1554–6. J Strain Anal Eng Des 2008;43(8):769–80.
[33] Hutchings MT, Withers PJ, Holden TM, Lorentzen T. Introduction [43] Akiniwa Y, Tanaka K, Kimura H. In: Residual stresses VII,
to the characterisation of residual stresses by neutron diffraction. Lon- proceedings, 2005. p. 118–23.
don: CRC Press; 2005. p. 424. [44] Newman JC. In: Methods and models for predicting fatigue crack
[34] Withers PJ, Preuss M, Steuwer A, Pang JWL. J Appl Crystallogr growth under random loading, STP 748. Philadelphia, PA: ASTM;
2007;40:891–904. 1981. p. 53–84.
[35] Suresh S. Fatigue Maters, Cambridge Solid Sci Series. Cambridge, [45] Wang H, Buchholz F-G, Richard HA, Jägg S, Scholtes B. Comput
UK: Cambridge University Press; 1991. p. 617. Mater Sci 1999;16:104–12.
[36] Patterson EA, Olden EJ. Fatigue Fract Eng Mater Struct [46] Otto JW. J Appl Crystallogr 1997;30:1008–15.
2004;27:623–36. [47] Jagg S, Scholtes B. Zeitschrift Fur Metallkunde 2005;96(7):768–74.
[37] Quinta da Fonseca J, Mummery PM, Withers PJ. J Micros [48] Liljedahl CDM, Zanellato O, Fitzpatrick ME, Lin J, Edwards L. Int J
2005;218:9–21. Fatigue 2010;32:735–43.

You might also like