Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233354763

Effect of dislocation hardening on monotonic and cyclic strength of severely


deformed copper

Article  in  Philosophical Magazine A · February 2012


DOI: 10.1080/14786435.2011.630693

CITATIONS READS

11 1,445

4 authors, including:

Alexei Vinogradov
Norwegian University of Science and Technology
217 PUBLICATIONS   5,659 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Cold Brittleness Investigation View project

vasilyevez@susu.ru View project

All content following this page was uploaded by Alexei Vinogradov on 03 May 2014.

The user has requested enhancement of the downloaded file.


Philosophical Magazine
2011, 1–24, iFirst

Effect of dislocation hardening on monotonic and cyclic strength


of severely deformed copper
A. Vinogradov*, M. Maruyama, Y. Kaneko and S. Hashimoto

Department of Intelligent Materials Engineering, Osaka City University,


Osaka 558-8585, Japan
Downloaded by [Monash University Library] at 14:56 13 November 2011

(Received 18 May 2011; final version received 1 October 2011)

The present study aims at clarifying the role of dislocation strengthening in


fatigue of materials manufactured by severe plastic deformation (SPD)
techniques. Employment of single crystals hardened via equal channel
angular pressing (ECAP) helps to minimise or completely eliminate the
effect of high angle boundaries on strengthening and fatigue behaviour.
Both monotonic strength and high cycle fatigue (HCF) resistance were
improved significantly after the first ECAP pressing, when low-angle
dislocation configurations dominate in the microstructure. The essential
role of dislocation accumulation during severe plastic deformation is
highlighted for both tensile and fatigue strength (SPD). Dilute alloying of
copper by silver stabilises the deformation microstructure and further
improves the fatigue properties considerably.
Keywords: equal channel angular pressing; fatigue; copper; electron
backscattering diffraction; dislocation structure

1. Introduction
Grain refinement down to a submicrocrystalline or nano-scale is regularly claimed as
the major factor responsible for the enhanced properties (yield stress, ultimate
strength, ductility, fatigue, etc.) of ultra-fine grain (UFG) materials manufactured by
severe plastic deformation (SPD) [1,2]. Several hypotheses have been proposed
regarding the mechanisms controlling the strength of UFG materials manufactured
by SPD. The most popular view is based on the Hall–Petch hardening dominance
caused by the small dimensions of highly misoriented domains within a polycrystal
[3,4]. This is supported by numerous experimental observations showing that the
angles of misorientation and the fraction of high-angle boundaries (HABs) increases
steadily after intensive plastic straining to the effective shear strains 48 [5,6]. Valiev
et al. [3] attributed the increased strength to a specific ‘‘non-equilibrium’’ state of
grain boundaries with high long-range stress fields after SPD (see also, [6–8]).
Alternatively, the high strength of severely deformed metals can be explained in
terms of strain hardening associated with accumulation of dislocations [9–17] and
disclinations [8] in the cell/fragment/grain boundaries. The latter approach offers a
self-consistent, entirely dislocation-based scenario of cell structure evolution,

*Corresponding author. Email: alexei@imat.eng.osaka-cu.ac.jp

ISSN 1478–6435 print/ISSN 1478–6443 online


ß 2011 Taylor & Francis
http://dx.doi.org/10.1080/14786435.2011.630693
http://www.tandfonline.com
2 A. Vinogradov et al.

showing that it is the dislocation cell size that determines the final grain size in the
material undergoing large strain deformation [18].
In the present communication, we highlight the significance of dislocation
hardening, which should not be underestimated when discussing strength-related
properties including the high cycle fatigue (HCF) life of SPD materials. Structure
refinement is a key feature of any SPD process aimed at ultimate strengthening
paired with comparatively good ductility [1–4]; however, the contributions of other
hardening mechanisms may be significant and should be considered.
Single crystals have long been recognised as model materials for studying many
fundamental aspects of dislocation strengthening and elementary mechanisms of
plastic deformation and fracture. This is because the crystal rotations and dislocation
structures can easily be traced with increasing strain under various deformation
Downloaded by [Monash University Library] at 14:56 13 November 2011

modes, such as in tension [19,20], cold rolling [21–25], channel die compression
[26–29] and equal channel angular pressing (ECAP) [30–39]. In addition, the complex
stress states associated with strain compatibility in the vicinity of grain boundaries in
polycrystals can be eliminated. Single crystals of Cu [30–34,37], Al [35,36], Nb [38]
and Cu bi-crystals [39] have been used in investigations aimed at understanding the
mechanisms of HABs formation and the role of initial orientation in structure
refinement during ECAP. The strong influence of the initial crystallographic
orientation on the microstructural evolution and crystal fragmentation has been
demonstrated convincingly in all studies. Despite considerable effort, the relationship
between the microstructure formed during SPD and properties is still unclear. In
particular, the effect of SPD on fatigue is of concern due to its crucial importance for
applications and the assessment of mechanisms of plasticity and fracture of UFG
structures.
The expanding experimental database of fatigue behaviour in UFG materials is
providing comprehensive coverage of fatigue properties under different conditions
[40–42]. However, data of this type are still limited. Most relevant fatigue studies
have been performed on UFG (grain size of 200–300 nm) specimens, which have
undergone intensive straining through 8–16 ECAP-passes [43–45]. SPD gives rise to a
typical UFG structure with grain dimensions in the order of 200 nm in copper and its
alloys. An enhancement of HCF life has been systematically observed in parallel with
monotonic strength. While this has been attributed to the strengthening effect of
HABs created during ECAP, the role of strain hardening associated with dislocation
accumulation inside the grain and in the cell and sub-boundaries with low angles of
misorientation has either been ignored or discussed only occasionally. In this
context, it is worth noting that there is a wealthy of publications showing that the
most impressive strengthening of 200–300% is usually achieved during straining to
relatively moderate strains, e.g. during the first ECAP-pressing to an equivalent
strain of 1.15, when the low angle boundaries dominate in the microstructure [46,47].
The formation of HABs, which is associated with further intensive straining to 8–16
or more passes, leads to only modest, if any, additional strengthening. HCF life
enhancement has been frequently noticed after ECAP and attributed to the
strengthening effect caused by HABs formed during SPD. However, the role of
dislocation accumulation and substructure strengthening seems inescapable when
discussing the mechanical behaviour of materials manufactured by SPD.
Particularly, since HCF life and monotonic strength are interrelated [40,41,48,49],
Philosophical Magazine 3

the influence of dislocation accumulation on fatigue should be better understood. A


comprehensive understanding of hardening mechanisms in the microstrain region is
crucial for the development of any realistic microstructurally based fatigue life
model. However, experiments have yet to be conducted under conditions which
permit a straightforward interpretation in terms of the strengthening mechanisms
involved. To meet this goal, we have investigated the mechanical behaviour, both
monotonic and cyclic, of copper single crystals with carefully selected crystallo-
graphic orientations subjected to one ECAP-pressing. Tailoring dislocation micro-
structures through simple shear deformation of orientation-controlled single crystals
to only one ECAP pass allowed us to fabricate a material with mechanical properties
comparable or superior to its polycrystalline counterparts deformed to the same or
even higher strains. High HCF properties were observed although the strengthening
Downloaded by [Monash University Library] at 14:56 13 November 2011

of single crystals was governed primarily by dislocation storage in configurations


with only low angles of misorientation. Since the thermo-mechanical stability of
dislocation configurations created during processing is of vital importance for
cyclic properties, we will demonstrate that microalloying furnishes an effective means
of controlling dislocation pinning, structure stabilisation and fatigue life
improvement.

2. Experimental procedure
Three copper single crystals of 99.96% purity, differing in their crystallographic
orientation, were grown in the form of 14  14  140 mm3 billets by a Bridgman
method. The normal direction of the shear plane (NSP) and shear direction (SD) is
related to the coordinates (ID, TD, ED) of the ECAP die, as shown in Figure 1a.
Initial crystallographic orientations of the specimens are shown in Figure 1b–d.
(1) The initial orientation of the first single crystal billet was (100) parallel to the
intrusion direction (ID) in the ECAP frame of reference (Figure 1), promoting
intensive multiple dislocation glide. (2) In the second orientation, the ID is parallel to
the h112i axis of the single crystal, the {1 1 1} slip plane is parallel to the theoretical
ECAP shear plane and the h110i slip directions are aligned with the ECAP SD, i.e.
single slip along the shear plane is promoted by this crystallographic orientation.
(3) The third crystal is oriented with the h011i direction parallel to ID, which is
supposed to facilitate a coplanar slip [19,20]. These crystals are further labelled as
#100, #211 and #011, respectively. They were subjected to one ECAP-pressing at
room temperature through the die with the minimised friction in rectangular
channels with hardened walls and sharp outer corner. This allowed us to impose
macroscopic simple shear deformation onto a working billet in good agreement with
the plastic flow patterns described by Segal [50,51]. The plastic deformation zone
was narrow, being confined to the theoretical shear plane without wide-spread
fan-shaped regions around the outer corner.
Our microstructural observations were concentrated on the analysis of
misorientations created during ECAP. In addition, the evolution of microstructure
during HCF tests was characterised by electron back-scatter diffraction (EBSD) and
electron channel contrast imaging (ECCI). ECCI observations and crystal orienta-
tion measurements were made using a JEOL JSM-6500FE field emission scanning
4 A. Vinogradov et al.
Downloaded by [Monash University Library] at 14:56 13 November 2011

Figure 1. ECAP schematics and stereographic projections showing the initial crystal
orientations with respect to the ECAP directions.

electron microscope (FE-SEM) equipped with an EDAX-TSL EBSD detector.


For large-area OIM observations, the imaging area was set at 320  1800 mm2 and
individual measurements were taken with incremental step of 1.5 mm. For finer
observations and measurements of the angles of misorientations, the imaging area
was set at 120  100 mm2 with a step size of 0.2 mm. The sections from the central part
of the billet were observed from the TD direction.
To prepare samples for mechanical tests and microstructural observations, the
billets were sectioned into three 3-mm thick slices parallel to the TD plane from the
central part of the billet. The most peripheral slices, which had been in the contact
with the die walls, were not used for sample preparation to eliminate possible effects
of the die friction-affected zone. Samples for mechanical testing were shaped by
spark erosion to have a square cross-section of 3  3 mm2 and 6 mm gauge length.
They were mechanically and electrolytically polished to a mirror-like finish and then
subjected to microstructural observations prior to and after fatigue tests. Tensile
tests were performed on a screw-driven testing machine with a nominal strain rate of
" ¼ 3:0  104 s1 . The HCF symmetric push–pull fatigue experiments were carried
out using dog-bone-shaped specimens mounted on a hydraulic testing frame
operating at 10 Hz under load control at constant stress-amplitudes at room
temperature. To clarify the effect of dislocation hardening on low-cycle fatigue
Philosophical Magazine 5
Downloaded by [Monash University Library] at 14:56 13 November 2011

Figure 2. Orientation image mapping and surface topographic images showing the micro-
structure created during ECAP of Cu single crystals with different initial crystallographic
orientations. Colours relate to the crystallographic orientations shown in the stereographic
triangle. The {111} pole figures with crystal orientations detected by EBSD after pressing are
shown for comparison.

(LCF) life, several ECAPed single- and poly-crystalline specimens were tested under
constant total D"t/2 or plastic D"pl/2 strain amplitudes ranging from 5  104 to
1.1  102 at 0.15 Hz. Conventional polycrystalline copper samples of the same
purity were processed by ECAP from a well-annealed state (800 C, 2 h in vacuum)
for comparison.

3. Results and discussion


3.1. Microstructural characterisation of single crystals after ECAP
The characteristics of slip systems and microstructure evolution have been detailed
for similar Cu and Al single crystals during ECAP [9–16], where both TEM and
EBSD observations were undertaken at various locations across the billet. The
mechanical behaviour of #100 and #211 single crystals has only been discussed in a
short communication [52]. According to the above studies, the localisation of plastic
deformation into macroscopic shear bands, such as those seen in the EBSD image
quality maps in Figure 2b and d, is a typical phenomenon in face-centred cubic (fcc)
metals at large strains [27–29].
In agreement with previous results cited above, all samples in the present study
reveal the strong dependence of deformation microstructures on the initial crystal
orientation. Textures, represented by discrete pole figures in Figure 2, are
characterised by typical shear components of fcc metals which deviate from the
position of ideal shear as is commonly observed in ECAPed metals [30,31,53–55].
In line with the findings of Miyamoto et al. [30,31] in Cu and Grosdidier et al. [53]
6 A. Vinogradov et al.

and Skrotzki et al. [54] in Ni, the #100 and #211 crystals are macroscopically
fragmented by deformation bands (Figure 2a–d). This creates a large orientation
scatter, which is revealed by the {111} pole figures (see insets in Figure 2). The
orientation spread is particularly strong in the #100 specimen. In contrast to the
#100 and #211 specimens, the #011 billet appears macroscopically homogeneous
without any deformation bands or highly-misoriented regions. Thus, the ‘‘single
crystal nature’’ is conserved in this sample although its crystallographic orientation
rotates reasonably from the initial orientation during processing, which is clearly
confirmed by the corresponding {111} discrete pole figure. This figure demonstrates
a very low orientation spread of less than 10 in the #011 specimen after ECAP.
More detailed TEM and EBSD microstructural observations have been reported in
[30,31] where the authors argued that the specific microstructure in the #011
Downloaded by [Monash University Library] at 14:56 13 November 2011

specimens is due to cell structures developing uniformly as a consequence of fairly


uniform slip in crystals with a given orientation.
Due to the disposition of highly symmetrical slip systems, the cube orientation is
unstable during room temperature plastic deformation and can exhibit macroscopic
shear and deformation bands under various loading conditions, such as tensile or
compressive straining, rolling, channel-die compression, ECAP, etc. [21–29, 53–57].
The EBSD analysis of different regions in deformed #100 and #211 single crystals
shows that the matrix is subdivided by dislocation walls generating small angles of
misorientations whereas large misorientations are observed between the matrix and
deformation bands as well as between the fragments within the deformation band.
Based on the EBSD data, the average deformation band width in the #100 and #211
specimens is 350  60 and 50  20 mm, respectively, with an interband spacing of
70  50 and 120  100 mm, respectively. Grosdidier et al. [53] and Skrotzki and
co-workers [54,55] have argued that the presence of the side walls, along which the
intensive friction is observed, is responsible for significant heterogeneity along the
transverse direction of the billet. Similar reasons for macroscopic shear banding in
metals subjected to ECAP have been given by Segal [51] from the viewpoint of solid
mechanics.
Preliminary TEM observations [30–32] have shown that the matrix in all
specimens is composed of dislocation cells with low angles boundaries. A rather
equiaxed cellular structure with cell dimensions in the range 0.2-0.5 mm was found in
the #100 and #211 specimens. In the #110 specimen, cells with dense dislocation
walls were elongated along the SD with an interboundary spacing of 0.4-0.5 mm. The
macroscopic deformation bands, which are clearly seen in Figure 2a and b in the
ECAPed #100 and #211 single crystals, respectively, consist of the clustered
microbands of 0.3–0.5 mm in width. Selected-area diffraction pattern analysis
confirmed the low angles of misorientations between the microbands as well as
between the cells in the matrix.
Correlated misorientation angle distributions obtained between neighbouring
points are presented in Figure 3. Low angle boundaries (LABs) are clearly
dominant in the microstructure. Note that Figure 3 underestimates the fraction of
low-angle boundaries, since the EBSD technique can resolve misorientations of
hi2 . Our EBSD observations are in good agreement with previous TEM
investigations on ECAPed Cu and Al single crystals [30–37]. Inter-band distances
of 100 and 200 mm on average are observed in the #100 and #211 samples,
Philosophical Magazine 7
Downloaded by [Monash University Library] at 14:56 13 November 2011

Figure 3. Uncorrelated distributions of angles of misorientation between adjacent regions in


single crystals after a single pass ECAP.

respectively. However, no deformation bands and no high angle boundaries are


seen in the #011 crystal (Figure 2e and f). An average effective cell diameter of
0.30  0.07 mm has been observed commonly in all crystals, in conformity with
TEM observations reported previously [30–37]. The orientation maps in Figure 2
and misorientation angle distributions in Figure 3 conclusively demonstrate that
the vast majority of boundaries can be classified as LABs in all specimens, which is
consistent with the results of small area TEM observations. In fact, the formation
of high angle boundaries is not expected at the modest shear strain imposed on the
sample during one ECAP-pass, ¼ 2. The development of high angle grain
boundaries is more commonly observed after SPD up to strains as high as
¼ 16–24 [1,3].
Concluding this section, although a few boundaries with high angle of
misorientation delineate the matrix and the deformation bands in #100 and
#211 crystals, no HABs have been found in the #011 crystal. Hence, it is
assumed that the high-angle grain boundary effect on the mechanical behaviour
of these samples is minimised or nullified, at least in the #011 crystal.
Therefore, the prepared samples are suitable for the present work, i.e. to clarify
the role of dislocation storage for fatigue life. For further in-depth discussions
on the microstructure and texture formation in single crystals during SPD,
readers are encouraged to consult the original papers given in the reference
section.
8 A. Vinogradov et al.

3.2. Tensile properties


Tensile stress–strain curves for the ECAP single crystals are shown in Figure 4 and
compared with the reference data obtained for an ordinary well annealed
polycrystalline sample subjected to one ECAP pass. All tests were repeated at least
twice using the specimens shaped from different parts of the billet. No meaningful
difference was found for the properties of specimens prepared from the centre or
from the periphery of the billet, which agrees with the uniformity of the
microstructure reported above. Mechanical properties are summarised in Table 1.
Downloaded by [Monash University Library] at 14:56 13 November 2011

Figure 4. Tensile stress–strain curves for Cu and Cu–0.07 (wt%) Ag single crystals after a
single ECAP.

Table 1. Tensile properties and fatigue limit of copper single- and poly-crystals subjected to
ECAP.

Sample (499.96% Cu)  0.2 (MPa)  UTS (MPa)  (%)  f (MPa)

SC Cu #112 ECAP 1 pass 250 330 25 90


SC Cu #100 ECAP 1 pass 330 405 25 75
SC Cu #011 ECAP 1 pass 260 275 20 105
SC Cu-Ag #100 ECAP 1 pass 380 400 20 110
PC Cu ECAP 1 pass 305 380 16 95
(from fully annealed state) [40]
PC Cu UFG ECAP 8 B [36] 390 440 22 80
PC Cu UFG ECAP 12 Bc [32,35] 480 512 17 115
PC Cu UFG ECAP 16 Bc [35] 350 413 15 110
PC fine grain Cu d ¼ 3.4 mm [54] 80 245 31* 60
fine grain Cu d ¼ 15 mm [54] 41 230 31* 60
coarse grain Cu d ¼ 150 mm [54] 15 250 28* 60
PC Cu [41] 220–480 60–80

Note: SC, single crystal; PC, polycrystals; *uniform true strain.


Philosophical Magazine 9

All single crystals show a spectacular increase in yield stress  0.2 after ECAP
compared to their as-grown counterparts. Among all the specimens, the #100 crystal
promotes multiple slip and rapid hardening during ECAP due to its ‘‘hard’’ initial
crystallographic orientation h100i parallel to the ID. It is not surprising that the
strength of single crystals deformed to one ECAP-pass, "eff  1.15, is lower than that
of polycrystals deformed much more severely to 8 and 16 passes, "eff  9 and 17,
respectively (Table 1). However, it is interesting that the specific strength charac-
teristics of the #100 sample are notably higher than those for the reference
polycrystal subjected to one ECAP-pass, which is obviously associated with the
effect of initial crystallographic orientation.
Downloaded by [Monash University Library] at 14:56 13 November 2011

3.3. Fatigue life


Texture-induced differences in slip characteristics are known to exert strong effects
on fatigue [58–60]. Llanes et al. [58] and Peralta et al. [59] have shown that ‘‘harder’’
crystallographic orientations, such as h111i or h100i, aligned with the loading axis
promote a high cyclic hardening in coarse grain copper polycrystals. In contrast, a
softer cyclic stress–strain behaviour is observed as a result of dominant single slip at
the same grain size. Hence, the effects of crystallographic texture are comparable or
even stronger than the effects of grain size on cyclic hardening in coarse grain
polycrystals.
Results of stress-controlled fatigue testing are displayed in Figure 5 as S–N or
W|hler plots. The fatigue life improvement in single crystals after one ECAP pass is
remarkable if compared to their annealed counterparts (see the comprehensive
results reported in [60–63]). The fatigue behaviour, specific dislocation patterning,

Figure 5. S–N fatigue life of Cu and Cu–0.07 (wt%) Ag single crystals after a single pass
ECAP versus reference data for conventional well-annealed polycrystals and UFG materials
manufactured by ECAP via a large number of pressings.
10 A. Vinogradov et al.

fatigue limit and relationship between fatigue of conventional annealed poly- and
single- crystals have been intensively studied and understood [60–70]. Summarising
the results of numerous studies, Li et al. [64] concluded that orientation exerted some
effect on the plastic strain fatigue limit, owing to the different dislocation structures
formed during cyclic deformation of differently oriented single crystals. At variance
with the conclusion drawn by Kettunen [61], they found that the fatigue limit
improved by activation of multiple slip. Lukáš and Kunz [60] have shown that
the S–N curves of Cu single crystals overlap those of polycrystals converted by the
Taylor factor M ¼ 3.06. The agreement was found to be particularly good in the
high-cycle regime where the stress fatigue limit  f0 was often in the range 60–80 MPa
for pure Cu at 107 cycles [49,65–67]. It is not surprising that the conventional fatigue
limit of strain-hardened single crystals surpassed that of annealed single- or poly-
Downloaded by [Monash University Library] at 14:56 13 November 2011

crystals (Table 1 and Figure 5). Among the three types of ECAPed crystals, the #100
specimen showed the longest life at higher stress amplitudes (150 MPa) in the LCF
region. In addition, its number of cycles to failure Nf was comparable or slightly
longer than that of the reference UFG polycrystal deformed up to eight passes [44].
However, as the stress amplitude was reduced to 80 MPa, the fatigue life of the #100
crystal became shorter than that of other crystals. On the other hand, the #211 single
crystals showed a notably longer HCF life and higher conventional fatigue limit than
annealed polycrystals. Moreover, its fatigue life was even longer than that of the
reference polycrystals produced by ECAP to eight passes. The highest fatigue
resistance at low stress amplitudes was demonstrated by the #011 specimen having
the homogeneous structure without any deformation bands and/or HABs. Overall,
the results of fatigue life measurements on single crystals fall within the scatter of
experimental data available in the literature for UFG polycrystals of comparable
purity manufactured by multiple ECAP (Figure 6). Since the hardening effect
associated with HABs is minor or absent in the single crystalline specimens under
investigation, enhancement of their fatigue life could be attributed to the excess
dislocation storage in LABs and cell structures formed during the single ECAP-pass.
Nonetheless, the effect of microstructure is not limited to the final strength. The
different microstructures respond differently to the imposed monotonic or cyclic
strains, as will be discussed in the next section. In brief, unlike the #011 crystal with a
uniform structure resulting in a uniform slip, the structural heterogeneity in the #100
and #211 crystals resulted in the significant heterogeneity of slip concentrated
primarily in the bands with preferred crystallographic orientations. This intensive
slip in the predefined regions determines the overall poor mechanical stability of the
microstructure and promotes recrystallisation and dynamic grain-coarsening during
cycling. In addition, the concentrated slip in heterogeneous structures results in
premature crack initiation and inferior fatigue performance in comparison with the
#110 single crystal of conventional ECAPed polycrystal, where the slip behaviour is
much more uniform.
A considerable improvement in HCF life is commonly observed in ECAP
materials compared to non-deformed specimens. Although intensive hardening
occurs by simple shear on the plane of intersection between the channels, formation
of HABs is not usually expected in polycrystals after imposing an equivalent strain of
1.15. Despite the reasonable experimental scatter, which is typical for fatigue data,
and despite the differences in slip behaviour caused by the differences in the
Philosophical Magazine 11
Downloaded by [Monash University Library] at 14:56 13 November 2011

Figure 6. (a) Cyclic softening curves obtained under total strain amplitude control for the
#110 ECAPed single crystal and reference polycrystals pressed once through the ECAP die
under the same conditions. (b) Coffin–Manson plot for the same materials. Data adapted
from the literature for UFG Cu processed by multipass ECAP are plotted for comparison.

microstucture of poly- and single-crystals, it is evident that the fatigue life of


ECAPed single-pass poly- and #110 single-crystals is virtually the same over a wide
range of stress amplitudes. It is evident that, at the reasonably modest strains
imposed on a working specimen during a single ECAP pass, the fatigue life is
controlled predominantly by dislocation storage in low angle configurations. At the
smallest D/2 values near the conventional fatigue limit, the fatigue life of #112 and
#011 single crystals can be even somewhat longer than that of the polycrystal pre-
deformed by ECAP under the same conditions. Vinogradov [40] and Estrin and
Vinogradov [41] discussed the traditional HCF fatigue improvement strategy based
on a correlation between monotonic and cyclic strength as applied to UFG metals
and alloys. While the high cyclic performance of UFG materials is improved
12 A. Vinogradov et al.

considerably with a grain size reduction to the sub-micrometer scale by SPD, these
materials are consistently inferior in their ability to sustain cyclic loads in the LCF
regime under high applied stress or strain amplitudes. The detail of fatigue life in
terms of the empiric Coffin–Manson and Basquin relations with plastic and elastic
strain amplitudes, D"pl/2 and D"el/2, respectively, has been widely discussed [42]. It
was demonstrated that the fatigue ductility coefficient "0f in the Coffin–Manson
expression:
D"pl
¼ "0f ð2Nf Þc ð1Þ
2
(c is the Coffin–Manson exponent) is drastically reduced in UFG materials
compared to that of its coarse grain counterparts, while the fatigue strength
Downloaded by [Monash University Library] at 14:56 13 November 2011

coefficient f0 , which is related to monotonic strength in Basquin’s law:


D"pl f0
¼ ð2Nf Þb ð2Þ
2 E
(b is the Basquin exponent, E is the Young’s modulus) shows the opposite trend and
is increased significantly. Consequently, in the HCF regime, which is dominated by
the elastic portion of the cyclic hysteresis loop according to Basquin’s law, the
increased fatigue strength coefficient accounts for the enhancement in fatigue life. On
the other hand, in the LCF regime, which is dominated by the plastic part of the
cyclic hysteresis loop according to the Coffin–Manson law, the decreased fatigue
ductility coefficient determines deterioration of fatigue life after intensive cold
working and grain refinement by SPD. Both the increase in strength and the
reduction in ductility relate to the ability of dislocations to move, i.e. to the
microstructural obstacles created primarily by the internal boundaries. Considering
the interactions between dislocations and grain boundaries, Höppel et al. [42] argued
that the reduced fatigue lives in the regime of intermediate to high plastic strain
amplitudes (low cycle fatigue regime) are due to a limitation in the possible slip
length of the dislocation slip path by the UFG structure. While the interaction of
gliding dislocations with high-angle grain boundaries during cyclic deformation
often results in grain boundary cracking due to high non-relaxed local stresses ahead
of dislocation pile-ups or PSBs, the low-angle boundaries are known to be immune
from crack nucleation owing to plastic strain compatibility between neighbouring
domains (sub-grains or cells). In this context, it is of interest to consider some
preliminary results obtained during strain controlled testing of ECAPed single
crystals in the LCF regime. The #110 single crystals, which are characterised by the
most homogeneous and stable microstructure (see the next section for details) and
which show the best HCF performance among other single crystals, were chosen for
this type of testing. Results, represented in Figure 6, show the typical cyclic
hardening/softening curves (a) and the Coffin–Manson plot (b) for both single- and
poly-crystalline specimens. Progressive cyclic softening is observed in ECAPed
specimens, as is commonly observed for cold-worked pure metals (Figure 6a) [42,71].
The stress amplitude in single crystals is consistently smaller than in polycrystals
under comparable conditions, as expected. However, if one compiles the current data
and the data available in the literature for multiple ECAPed Cu polycrystals on the
same Coffin–Manson plot (Figure 6b), all results fall within the same reasonably
Philosophical Magazine 13

Table 2. Parameters of the Basquin and Coffin–Manson relation derived from stress- and
strain-controlled tests.

Basquin law Coffin–Manson

Specimen f0 b "0f c

CG Cu [72]* 1.02 0.62  0.04


[73] 308 0.075 2.18 0.66
UFG 8Bc ECAP Cu 630  65 0.13  0.02 0.22  0.06 0.50  0.05
ECAP 1 pass Cu 537  51 0.11  0.01 0.23  0.06 0.54  0.04
ECAP 1 pass #011 Single crystal 602  42 0.11  0.02 0.31  0.07 0.52  0.04

Note: *Parameters were calculated from original data presented in the cited study.
Downloaded by [Monash University Library] at 14:56 13 November 2011

narrow scatter band (bearing in mind the usual scatter of fatigue data commonly
linked to variability in the microstructures, chemical compositions and sensitivity to
possibly different surface and experimental conditions), as shown by dashed lines.
Moreover, the slope of all data – adapted from the literature and present results –
corresponding to the Coffin–Manson exponent c is similar, 0.5, which is slightly
smaller that of conventional coarse grain Cu (Table 2). Moreover, the results shown
in Figure 6b agree well with findings by Lukáš et al. [73] who examined the Coffin–
Manson behaviour of UFG Cu of different purity: the same value for the c exponent
can be inferred from their data. Moreover, they have shown that, similarly to
Figure 6b, all results in the Coffin–Manson plot fall within a narrow band regardless
of purity, number of ECAP passes or strain path. Taking all these data in
consideration, one can suggest that the generic mechanism controlling fatigue life in
the range of intermediate strain amplitudes is similar in these materials regardless of
the details of their microstructures. Hence, the LCF life appears to be controlled
primarily by the cyclic ductility coefficient, which cumulatively accounts for virtually
all differences in microstructures and cyclic deformation behaviour. Evidently,
within the scatter of the experimental results, no appreciable difference is found in
the LCF life of UFG samples and single- and poly-crystals subjected to a single
pass ECAP.
Parameters of the Basquin and Coffin–Manson relation derived from stress- and
strain-controlled tests, repsectively, are summarized in Table 2. The exponets c and b
remain nearly the same for all ECAPed samples. On the other hand, the differences
in the fatigue strength coefficient reflect the trends in HCF lives.
If, as it is commonly assumed, the LCF life is mediated by fatigue crack
propagation, the relatively low sensitivity of constitutive parameters entering the
Coffin–Manson relation should be linked to the ‘‘universal’’ development of surface
damage. Aiming to account for the seeming universality of the Coffin–Manson
exponent, Sornette el al. [74] proposed a statistical approach based on observations
of the ‘‘mesoscopic’’ structural scale intermediate between the dislocation scale and
the macroscopic crack scale, which plays a fundamental role in microcrack initiation
and propagation. The universality of the c exponent is explained from the relevance
of the mesoscopic scale at which specific details are washed out and ‘‘renormalized’’
at the grain scale into a random nucleation process. Hence, this mesoscopic scale can
14 A. Vinogradov et al.

be plausibly associated with the ‘‘universal’’ cell size, which is commonly of a sub-
micrometer order and which depends only weakly on processing and loading
conditions.
Comparing the results shown in Figures 5 and 6 for stress- and strain-controlled
testing, one can notice that the fatigue lives do not simply agree with each other if
one tries to translate the results of the strain-controlled tests into an S–N diagram.
For example, the stress level at half-life under strain-controlled cycling at
D"pl/2 ¼ 1  103 is 120 MPa (Figure 6 compared to Figure 5), while the specimen
tested under a constant stress amplitude of the same level lasts much longer. For
cyclically fully stable materials, which do not exhibit any hardening and/or softening
during cycling, the translation between fatigue life diagrams obtained either under
stress- or strain-controlled experiments is straightforward [75]. The agreement
Downloaded by [Monash University Library] at 14:56 13 November 2011

between the transformed and the directly measured fatigue life curves is expected to
be satisfactory for materials with a short stage of hardening and/or softening.
However, for longer and more pronounced hardening/softening, the agreement
between the transformed and the experimentally measured curves is significantly
worse. Hence, it can be argued that this discrepancy is caused by significant
instability of the hysteresis loop due to intensive cyclic softening at the onset of
loading. This effect is particularly pronounced in samples after one EACP pass, for
an obvious reason that the dislocations accumulated at modest strains of 1.15
attained during single pressing have not been fixed firmly in ‘‘optimal’’ low energy
configurations and, therefore, recovery is expected under cyclic deformation. The
easy cross-slip in copper facilitates dislocation annihilation processes associated with
fatigue-induced recovery and softening. The microstructure after fatigue and the slip
behaviour during cyclic deformation will be discussed in the next section.
In the initial part of the cyclic hardening/softening curve the stress amplitude
reaches a value higher than 200 MPa but then decreases quickly and significantly.
This initial load ‘‘overshooting’’ in a strain-controlled regime may induce substantial
stress-dependent changes in the microstructure and may significantly damage a
fatigued sample already in the early stage of cycling. In other words, it is evident that
the sensitivity of estimated fatigue life to cyclic softening is a clear function of the
degree of softening exhibited by a material. Although, it may not be significant at
extremely long life, in the intermediate strain/stress region, this phenomenon cannot
be disregarded. Hence, unlike materials showing stable hysteresis behaviour, it is
difficult (if not impossible) to translate the results of the present LCF test at constant
plastic strain amplitude to the total life diagram without substantial uncertainty
arising from variations in the stress (or strain) amplitudes if the stress amplitude is
evaluated at half life. This situation is not unique as the extreme influence of cyclic
softening on predicted fatigue life in the intermediate life regions (104–105 reversals)
has been reported previously [76].

3.4. Microstructure after HCF fatigue: structural instability and fatigue-induced


recrystallisation and grain growth
One question to be addressed is why the #100 single crystals show shorter fatigue
life at lower stress amplitudes? The EBSD and ECCI SEM observations of the
Philosophical Magazine 15

post-fatigue microstructures suggest that the degradation of HCF resistance in #100


samples can be plausibly associated with the instability of their deformation
microstructure. This microstructure, consisting primarily of LABs, was significantly
destroyed by cyclically induced recovery and recrystallisation processes (Figure 7).
Actually, some partly recrystallised regions were observed occasionally, even in the
as-ECAPed #100 crystal (Figure 2a). However, after fatigue testing, coarsened areas
separated by HABs are readily observed across the TD sample surface. As soon as
cyclically induced grain growth occurs, the coarsened regions with reduced
dislocation density become the preferential sites for localised cyclic plastic deforma-
tion. Figure 7d illustrates the formation of an ordinary dislocation vein structure in
the dynamically coarsened grains in the #100 specimen. It is well known that
recrystallisation may take place during or after SPD of pure metals, even at room
Downloaded by [Monash University Library] at 14:56 13 November 2011

temperature, for high purity Al [67] and for Cu [76,77]. Fatigue-induced dynamic
cell-coarsening, recrystallisation and grain growth were also frequently observed
during cyclic deformation (though mostly in the strain-controlled testing) of UFG
Cu, Al, and Ni polycrystals [43,71,78–81] (see also [82] for a comprehensive review).
The #112 Cu samples (Figure 8) were less prone to recrystallisation and grain growth

Figure 7. SEM ECCI images showing fine structure in the flow plane (TD) of the #100 Cu
single crystal after fatigue at 150 MPa (a,c) and around the fatigue limit at 80 MPa (b,d). (a,b)
Low magnification; (c,d) high magnification. Recrystallised areas are evident. The loading axis
is horizontal and parallel to the extrusion direction ED.
16 A. Vinogradov et al.

Figure 8. SEM ECCI images showing fine structure in the flow plane (TD) of the #211 Cu
Downloaded by [Monash University Library] at 14:56 13 November 2011

single crystal after fatigue at 150 MPa (a),and 80 MPa (b). Recrystallised regions are
occasionally seen only after HCF at 80 MPa.

during cyclic deformation; although a few recrystallised regions were also occasion-
ally found after long-term cycling (approximately 2 weeks) at low stress amplitude
(80 MPa) near the fatigue limit (Figure 8b). The #011 samples, however, appeared
immune from any recrystallisation; no signs of recrystallisation or structure
coarsening were observed in this type crystal after ECAP or after fatigue.
Figure 9 shows ECC (electron channelling contrast) and SEI (secondary electron
image) micrographs of the fatigue microstructure (Figure 9a and c, respectively) and
the fatigue surface morphology (Figure 9b and d) of the #110 specimen tested under
constant plastic strain amplitudes of D"pl/2 ¼ 1  103 (Figure 9a and b) and
D"pl/2 ¼ 5  103 (Figure 9c and d). The uniform cellular dislocation structure with a
network of dislocation walls having inter-boundary spacing less than 1 mm is seen in
Figure 9a and c, which agrees with TEM observations reported earlier [30–32] and
EBSD results discussed above. At lower plastic strain amplitudes, the fatigue
markings with visible intrusions and extrusions are confined to the primary slip
system and are uniformly distributed across the sample cross-section. At higher
plastic strain amplitudes, the single slip cannot accommodate the imposed plastic
strain and the secondary slip systems are activated. Multiple fatigue cracks are seen
along the coarse primary slip bands. The interaction between the secondary and
primary slip at larger D"pl/2 values promotes fatigue cracking which finally evolves
to a catastrophic failure.
Figure 10 further confirms that the fatigue damage morphology after stress-
controlled testing follows the crystallographic features of the microstructure
obtained during ECAP. The slip is heterogeneous on a micro- and macro-scale in
the #100 and #211 samples, being confined to deformation bands created during
ECAP (Figure 2). The density of fatigue-induced slip markings within the bands is
significantly higher than in the matrix. In the contrast, the slip is fairly uniform in the
#011 crystal. At low stress or strain amplitudes, only one primary slip system is
activated in this specimen, giving rise to typical fatigue morphology with extrusions
and intrusions on a cyclically deformed surface. Figure 10 provides further clear
evidence for the negative role of cyclically induced grain growth in fatigue life: slip
Philosophical Magazine 17
Downloaded by [Monash University Library] at 14:56 13 November 2011

Figure 9. SEM ECCI images showing the cellular structure in the ECAPed #110 single crystal
after testing at D"pl/2 ¼ 1  103 (a) and D"pl/2 ¼ 5  103 (c), and coarse fatigue slip markings
in the flow plane (TD) corresponding to the same testing conditions (b and d, respectively).
Arrows L.A. indicate loading axis.

lines are clearly visible to be localised primarily within the coarsened regions of #100
and #211 crystals (Figure 10a, b and d). No signs of recrystallisation were found in
the #011 crystal after fatigue at different D/2 magnitudes (Figure 10e and f).
Dynamic recrystallisation and coarsening occurs at the expense of lattice disloca-
tions. Hence, coarsened grains are virtually dislocation-free and yield at lower
stresses than the rest of the samples, thereby providing a natural pathway for early
strain localisation and premature failure either due to cracking at the interfaces
between the coarse grains and the surrounding fine-grain matrix or due to
transgranular surface crack initiation in a rough intrusion.
Note again that the #011 single crystal exhibits the longest HCF life among the
other single- or poly-crystals tested in the present work or referred to in the literature
[83]. This is despite its lower yield stress or ultimate tensile strength (Figure 4 and
Table 1) and, therefore, the extended fatigue life in this specimen should be related to
delayed strain localisation and crack initiation due to the uniformity of dislocation
glide and to the chance of interaction between slip lines and grain boundaries.
A compilation of the available experimental data, shown in Figure 5, verifies that
substantial grain refinement through multiple ECAP can exert an extra positive
18 A. Vinogradov et al.
Downloaded by [Monash University Library] at 14:56 13 November 2011

Figure 10. SEM images showing fatigue slip markings in the flow plane (TD) of single
crystals: (a,b) #100, (c,d) #211 and (e,f) #011 at different scale. Arrows L.A. indicate
loading axis.

effect on fatigue resistance, which is undoubtedly associated with formation of an


UFG microstructure with HABs. This demonstrates the significant potential offered
by SPD for tailoring microstructures, where combinations of different strengthening
mechanisms give rise to superb fatigue performance.
Overall, the fatigue performance and fatigue mechanism are greatly affected by
the stability/instability of the microstructure. When longer fatigue life is of concern,
one should benefit from a stable microstructure, as has been also emphasised by
Kunz et al. [84]. A common metallurgical strategy to stabilise the microstructure
employs dislocation pinning by alloying. In fact, it has long been recognised that
chemical composition has a major effect in fatigue [60]. In the present work,
Philosophical Magazine 19

we attempted to achieve a stabilised microstructure by ‘anchoring’ dislocations while


keeping solid solution strengthening as low as possible. We produced a dilute CuAg
solid solution with just 0.07 (wt%) addition of silver. Silver has low solubility and a
large size effect (rAg/rCu ¼ 1.13) in a copper matrix. This should induce strong
dislocation pinning. The single crystals of this CuAg alloy were grown from Cu seeds
of #100 orientation and then tested under the same conditions as reference copper
single crystals. The microstructure of the CuAg single crystal after ECAP was not
appreciably different from that of the Cu #100 crystal shown in Figure 2, except that
no recrystallised regions are observed across the CuAg specimen. Within the scatter
of experimental results, its ultimate tensile strength (Figure 5) was also the same as
that of the #100 Cu single crystals, indicating that the solution-hardening effect is
minor. However, the initial hardening rate and yield stress in the CuAg alloy were
Downloaded by [Monash University Library] at 14:56 13 November 2011

notably higher, while uniform elongation was shorter in the reference Cu #100
crystal. This can be plausibly linked to structure stabilisation and the absence of any
recrystallisation and grain coarsening during or after fabrication of this alloy.
Figure 5 shows that fatigue life benefits greatly from microalloying. The #100
CuAg single crystal lasted much longer than its pure copper counterpart or even
conventional polycrystals of the same purity under the same cyclic stress amplitude.
At relatively high cyclic stress amplitudes of 150 and 120 MPa, the number of cycles
to failure of this microalloyed crystal was as a high as that reported for 12 times
ECAPed UFG copper [82]. At the lower stress amplitude of 100 MPa, the fatigue life
of CuAg single crystal increased to an even higher number of cycles, exceeding 107
without failure. This is notably longer than the available reference data for copper
polycrystals subjected to single or multiple ECAPing [32,35,36,48,78]. SEM ECCI
images illustrating the microstructure of the #100 Cu-Ag single crystals after ECAP
and after fatigue are shown in Figure 11a and b, respectively. The SEM images
shown in Figure 11c display a slip pattern in the vicinity of the major fatigue crack.
No signs of recrystallisation or structure coarsening during fatigue were found in
these microalloyed specimens. Hence, the microstructure had been successfully
stabilised by impurity addition and this, in turn, caused a significant enhancement in
fatigue life. The value of the conventional fatigue limit at 110 MPa in the
microalloyed single crystalline specimens is quite remarkable and is higher than that
typically documented for commercial purity Cu polycrystals [49]. However, Kunz
et al. [84] have reported even higher values of 150 MPa for lower purity (99.5%)
UFG Cu produced by multipass ECAP and observed no structural change after
fatigue. Therefore, the improved fatigue life is associated with improved micro-
structural stability due to efficient pinning of dislocation and grain boundaries by
impurities.
Finally, take note that specific beneficial combinations of properties, such as high
strength and acceptable ductility, can be attained by grain boundary engineering
[84,85], where SPD can be also involved. For instance, low-angle boundaries are
reasonably resistant to intergranular fracture while HABs are the sites most prone to
crack nucleation and propagation, particularly in fatigue [86]. Therefore, in many
cases, a desired HCF resistance can be efficiently achieved through the formation of
low-angle boundaries alone. This is attainable by the deformation to moderately
large shear strains ¼ 2–8.
20 A. Vinogradov et al.
Downloaded by [Monash University Library] at 14:56 13 November 2011

Figure 11. SEM ECCI images showing the microstructure of the #100 dilute Cu–0.07 (wt%)
Ag single crystal after ECAP (a) and after fatigue (b). (c) SEM image showing the slip pattern
in the vicinity of a major fatigue crack. No signs of recrystallisation or structure coarsening are
evident.

4. Summary and conclusions


(1) In the present study, two strategic ideas have been explored experimentally
to clarify the role of dislocation strengthening in the fatigue of
Philosophical Magazine 21

materials manufactured by SPD. First, the employment of single crys-


tals hardened via ECAP minimised or completely eliminated the effect of
HABs on strengthening, fatigue behaviour and damage. Second,
dilute alloying of copper by silver stabilised a deformation microstructure
dominated by low-angle boundaries and considerably improved fatigue
properties.
(2) Dislocation hardening is a crucial mechanism influencing fatigue life of
metals manufactured by SPD techniques. Effective strengthening through a
single pass ECAP, whereby low-angle boundaries form predominantly,
notably improved HCF properties. Considering the processing cost, this may
a preferred mechanism compared to Hall–Petch strengthening achieved
through the formation of HABs at very high strains.
Downloaded by [Monash University Library] at 14:56 13 November 2011

(3) Single pass ECAP resulted in the formation of low angle dislocation
configurations with very homogeneous distribution in h110i-oriented sin-
gle crystals. This preserved the uniformity of slip during testing at low
and moderate stress amplitudes. Combined with the chance of interac-
tion between dislocation slip and grain boundaries, this improved the HCF
life of monocrystalline versus polycrystalline copper after the same
processing.

Acknowledgements
The authors are indebted to Professor H. Miyamoto and Dr. D. Orlov for their interest in the
present topic and fruitful discussions. Special thanks are due to Professor H. Mughrabi for
useful discussions and critical comments. The help of M. Ueno with LCF experiments is
appreciated. This study was financially supported by the Grant-in-Aid for Scientific Research
on Innovative Area, ‘‘Bulk Nanostructured Metals’’, through MEXT Japan (contract
No. 22102006), and this support is gratefully appreciated. One of the authors (AV) wishes to
thank the Russian Ministry of Education and Science for partial support of this research
through the Grant-in-Aid No. 11.G34.31.0031.

References

[1] R.Z. Valiev and T.G. Langdon, Prog. Mater. Sci. 51 (2006) p.881.
[2] A. Azushima, R. Kopp, A. Korhonen, D.Y. Yang, F. Micari, G.D. Lahoti, P. Groche,
J. Yanagimoto, N. Tsuji, A. Rosochowski and A. Yanagida, CIRP Ann. Manuf. Technol.
57 (2008) p.716.
[3] R.Z. Valiev, R.K. Islamgaliev and I.V. Alexandrov, Prog. Mater. Sci. 45 (2000) p.103.
[4] R.Z. Valiev, Nature 419 (2002) p.887.
[5] M. Furukawa, Z. Horita and T.G. Langdon, J. Mater. Sci. 40 (2005) p.909.
[6] E.V. Kozlov, A.N. Zdanov, N.A. Popova and N.A. Koneva, A composite grain model of
strengthening for SPD produced UFG materials, in Nanomaterials by Severe Plastic
Deformation, M. Zehetbauer and R.Z. Valiev, eds., Willey/VHC, New York, 2004, p.263.
[7] A.A. Popov, I.Y. Pyshmintsev, S.L. Demakov, A.G. Illarionov, T.C. Lowe,
A.V. Sergeyeva and R.Z. Valiev, Scripta Mater. 37 (1997) p.1089.
[8] M. Seefeldt, Rev. Adv. Mater. Sci. 2 (2001) p.44.
[9] G.A. Malygin, Phys. Solid State 46 (2004) p.2035.
[10] G.A. Malygin, Phys. Solid State 44 (2001) p.2072.
22 A. Vinogradov et al.

[11] Y. Estrin, L.S. Tóth, A. Molinari and Y. Bréchet, Acta Mater. 46 (1998) p.5509.
[12] Y. Estrin, L.S. Tóth, Y. Bréchet and H.S. Kim, Mater. Sci. Forum 503/504
(2006) p.657.
[13] R.J. Hellmig, S.C. Baik, J.R. Bowen, Y. Estrin, D.J. Jensen, H.S. Kim and M.H. Seo,
Evolution of mechanical and microstructural properties of ECAP deformed copper,
in Nanomaterials by Severe Plastic Deformation, M. Zehetbauer and R.Z. Valiev, eds.,
Willey/VHC, New York, 2004, p.257.
[14] N.Q. Chinh, J. Gubicza and T.G. Langdon, J. Mater. Sci. 42 (2007) p.1594.
[15] D.A. Hughes and N. Hansen, Acta Mater. 48 (2000) p.2985.
[16] N. Hansen, X. Huang and D.A. Hughes, High strain monotonic deformation-structure and
strength, in Ultrafine Grained Materials II, Y.T. Zhu, R.S. Mishra, S.L. Semiatin,
M.J. Saran and T.C. Lowe, eds., TMS, Warrendale, PA, 2002, p.3.
[17] N. Hansen and D.J. Jensen, Philos. Trans. R. Soc. A 357 (1999) p.1447.
Downloaded by [Monash University Library] at 14:56 13 November 2011

[18] Y. Estrin and H.S. Kim, J. Mater. Sci. 42 (2007) p.9092.


[19] J.A. Wert, X. Huang and F. Inoko, Proc. R. Soc. Lond. A 459 (2003) p.85.
[20] J.A. Wert, K. Kashihara, T. Okada and T. Huang, F. Inoko. Philos. Mag. 85 (2005)
p.1989.
[21] P. Wagner, O. Engler and K. Lucke, Acta Metall. 43 (1995) p.3799.
[22] M. Haberjahn, P. Klimanek and M. Motylenko, Mater. Sci. Eng. A 324 (2002) p.169.
[23] A. Malin, J. Huber and M. Hatherly, Z. Metallk. 72 (1981) p.310.
[24] G.D. Kohlhoff, A.S. Malin, K. Lucke and M. Hatherly, Acta Metall. 36 (1988) p.2841.
[25] M. Wrobel, S. Dymek and M. Blicharski, Scripta Mater. 35 (1996) p.417.
[26] M. Lewandowska, W. Wiatnicki, A. Piatkowski and Z. Jasiesk, J. Microsc. 223
(2006) p.275.
[27] Z. Jasieński and A. Piatkowski, Adv. Eng. Mater. 12 (2010) p.1068.
[28] A. Akef and J.H. Driver, Mater. Sci. Eng. A 132 (1991) p.245.
[29] F. Basson and J.H. Driver, Acta Mater. 48 (2000) p.2101.
[30] H. Miyamoto, U. Erb, T. Koyama, T. Mimaki, A. Vinogradov and S. Hashimoto, Philos.
Mag. Lett. 84 (2004) p.235.
[31] H. Miyamoto, J. Fushimi, T. Mimaki, A. Vinogradov and S. Hashimoto, Mater. Sci. Eng.
A 405 (2005) p.221.
[32] H. Miyamoto, J. Fushimi, T. Mimaki, A. Vinogradov and S. Hashimoto, Mater. Sci.
Forum 503/504 (2006) p.799.
[33] G. Wang, S.D. Wu, Q.W. Jiang, Y.D. Wang, Y.P. Zong, C. Esling and L. Zuo, Mater.
Sci. Forum 495–497 (2005) p.815.
[34] M. Furukawa, Y. Fukuda, K. Oh-ishi, Z. Horita and T.G. Langdon, Mater. Sci. Forum
503/504 (2006) p.113.
[35] Y. Fukuda, K. Oh-ishi, M. Furukawa, Z. Horita and T.G. Langdon, Acta Metall. 52
(2004) p.1387.
[36] Y. Fukuda, K. Oh-ishi, M. Furukawa, Z. Horita and T.G. Langdon, Mater. Sci. Eng. A
420 (2006) p.79.
[37] Y. Fukuda, K. Oh-ishi, M. Furukawa, Z. Horita and T.G. Langdon, J. Mater Sci. 42
(2006) p.1501.
[38] H.R.Z. Sandim, H.H. Bernardi, B. Verlinden and D. Raabe, Mater. Sci. Eng. A 467
(2007) p.44.
[39] W.Z. Han, H.J. Yang, X.H. An, R.Q. Yang, S.X. Li, S.D. Wu and Z.F. Zhang, Acta
Mater. 57 (2009) p.1132.
[40] A. Vinogradov, J. Mater. Sci. 42 (2007) p.1797.
[41] Y. Estrin and A. Vinogradov, Int. J. Fatigue 32 (2010) p.898.
[42] H.-W. Höppel, H. Mughrabi and A. Vinogradov, in Bulk Nanostructured Materials,
Chap. 22, M.J. Zehetbauer and Y.T. Zhu, eds., Wiley/VCH, New York, 2008, p.481.
Philosophical Magazine 23

[43] A. Vinogradov and S. Hashimoto, Mater. Trans. JIM 42 (2001) p.74.


[44] S.R. Agnew, A. Yu. Vinogradov, S. Hashimoto and J.R. Weertman, J. Electron. Mater.
28 (1999) p.1038.
[45] H. Mughrabi and H.W. Höppel, Mater. Res. Soc. Symp. Proc. 634 (2001) p.B2.1.1.
[46] H.J. Maier, P. Gabor, N. Gupta, I. Karaman and M. Haouaoui, Int. J. Fatigue 28 (2006)
p.243.
[47] F. Dalla Torre, R. Lapovok, J. Sandlin, P.F. Thomson, C.H.J. Davies and
E.V. Pereloma, Acta Mater, 52 (2004) p.4819.
[48] A. Vinogradov, S. Yasuoka and S. Hashimoto, Mater. Sci. Forum 584–586 (2008) p.797.
[49] M.C. Murphy, Fatigue Eng. Mater. Struct. 4 (1981) p.199.
[50] V.M. Segal, Mater. Sci. Eng. A 345 (2003) p.36.
[51] V.M. Segal, Mater. Sci. Eng. A 271 (1999) p.322.
[52] A. Vinogradov, T. Maruyama and S. Hashimoto, Scripta Mater. 61 (2009) p.817.
Downloaded by [Monash University Library] at 14:56 13 November 2011

[53] T. Grosdidier, J.-J. Fundenberger, D. Goran, E. Bouzy, S. Suwas, W. Skrotzki and


L.S. Tóth, Scripta Mater. 59 (2008) p.1087.
[54] W. Skrotzki, L.S. Tóth, B. Klöden, H.G. Brokmeier and R. Arruffat-Massion, Acta
Mater. 56 (2008) p.3439.
[55] W. Skrotzki, N. Scheerbaum, C.-G. Oertel, H.-G. Brokmeier, S. Suwas and L.S. Tóth,
Acta Mater. 55 (2007) p.2211.
[56] J.A. Wert, Q. Liu and N. Hansen, Acta Mater. 45 (1997) p.2565.
[57] K. Morii, H. Mecking and Y. Nakayama, Acta Metall. 35 (1987) p.1747.
[58] L. Llanes, A.D. Rollett, C. Laird and J.L. Bassani, Acta Metall. 41 (1993) p.2667.
[59] P. Peralta, A. Czapka, K. Obergfell, L. Llanes, C. Laird and T.E. Mitchell, Mater. Sci.
Eng. A 270 (1999) p.349.
[60] P. Lukáš and L. Kunz, Mater. Sci. Eng. A 189 (1994) p.1.
[61] P.O. Kettunen, Philos. Mag. 14 (1966) p.421.
[62] P.O. Kettunen, Philos. Mag. 16 (1967) p.253.
[63] G. Rudolph, E. Haasen, B. L. Mordike and P. Neumann, in Proceeding ICF-1,
T. Yokobori, T. Kawasaki and J.L. Swedlow, eds., Nihon Gakujutsu ShinkOkai, Sendai,
1965, Vol. 2, p. 501.
[64] X.W. Li, Y. Umakoshi, Z.G. Wang and S.X. Li, Philos. Mag. Lett. 81 (2001) p.465.
[65] P. Lukáš and L. Kunz, Mater. Sci. Eng. A 85 (1987) p.67.
[66] A.W. Thompson and W.A. Backofen, Acta Metall. 19 (1971) p.597.
[67] C.D. Liu, D.X. You and M.N. Bassim, Acta Metall. 42 (1994) p.1631.
[68] H. Mughrabi, Philos. Mag. 23 (1971) p.869.
[69] H. Mughrabi, Scripta Metall. 13 (1979) p.479.
[70] P. Neumann, Mater. Sci. Eng. 81 (1986) p.465.
[71] S.R. Agnew and J.R. Weertman, Mater. Sci. Eng. A 244 (1998) p.145.
[72] H. Mughrabi and R. Wang, Cyclic stress-strain response and high cycle behaviour of copper
polycrystals, in Basic Mechanisms in Fatigue of Metals, P. Lukáš and J. Polák, eds.,
Academia/Elsevier, Prague/Amsterdam, 1988, p.1.
[73] P. Lukáš, M. Klensil and J. Polák, Mater. Sci. Eng. 15 (1974) p.239.
[74] D. Sornette, T. Magnin and Y. Bréchet, Europhys. Lett. 20 (1992) p.433.
[75] P. Lukáš, L. Kunz and M. Svoboda, Kovove Mater. 47 (2009) p.1.
[76] H.R. Jhansale, J. Test. Eval. 3 (1975) p.348.
[77] H. Miyamoto, T. Mimaki, A. Vinogradov and S. Hashimoto, Annales de Chimie 27
(2002) p.S197.
[78] X. Molodova, G. Gottstein, M. Winning and R.J. Hellmig, Mater. Sci. Eng. A 460/461
(2007) p.204.
[79] E. Thiele, C. Holste and R. Klemm, Z Metallkd. 93 (2002) p.730.
24 A. Vinogradov et al.

[80] H.W. Höppel, Z.M. Zhou, H. Mughrabi and R.Z. Valiev, Philos. Mag. A 82 (2002)
p.1781.
[81] H. Mughrabi, H.W. Höppel and M. Kautz, Scripta Mater. 51 (2004) p.807.
[82] H. Mughrabi and H.W. Höppel, Int. J. Fatigue 32 (2010) p.1413.
[83] H. Mughrabi and H.W. Höppel, Mater. Res. Soc. Symp. Proc. 634 (2001) p.B2.1.1.
[84] L. Kunz, P. Lukáš and M. Svoboda, Mater. Sci. Eng. A 424 (2006) p.97.
[85] T. Watanabe, Res Mechanica 11 (1984) p.47.
[86] S. Kobayashi, T. Inomata, H. Kobayashi, S. Tsurekawa and T. Watanabe, J. Mater. Sci.
43 (2008) p.3792.
Downloaded by [Monash University Library] at 14:56 13 November 2011

View publication stats

You might also like