Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Acta mater.

49 (2001) 3755–3765
www.elsevier.com/locate/actamat

AFM CHARACTERIZATION OF THE EVOLUTION OF SURFACE


DEFORMATION DURING FATIGUE IN POLYCRYSTALLINE
COPPER
L. CRETEGNY1‡ and A. SAXENA2†
1
GE Corporate R&D Center, Schenectady, NY, USA and 2School of Materials Science and Engineering,
Georgia Institute of Technology, Atlanta, GA 30332-0245, USA

( Received 27 December 2000; received in revised form 6 July 2001; accepted 6 July 2001 )

Abstract—Atomic force microscopy (AFM) is a relatively new tool that readily provides high resolution
digitized images of surface features. AFM is used here to study the development of slip bands and protrusions
in strain controlled fatigue tests on polycrystalline copper at 0.161 and 0.255% strain amplitudes. The average
slip band heights at failure for both strain amplitudes conditions are comparable, implying that the growth
of slip bands saturates at a specific height. A parameter, γirrev, is defined that is a measure of the local slip
irreversibility at the surface and is applicable to any type of surface deformation feature, independently of
the size of the fields of view. Thus, estimates of surface deformation developed in regions where fatigue
crack nucleation is likely to occur can be obtained, from which a fatigue crack nucleation criterion is defined.
 2001 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.

Résumé—La microscopie à force atomique (AFM) est un outil qui permet l’obtention d’images digitales à
haute résolution d’éléments de déformation de la surface et leur analyse quantitative. Cette technique est
utilisée ici pour étudier le dévelopement de bandes de glissements et de protrusions lors d’essais de fatigue
en déformation contrôlée sur du cuivre polycristallin à des amplitudes de déformation relative de 0.161 et
0.255%. Pour les deux conditions d’essais, la hauteur moyenne des bandes de glissement après rupture était
comparable, ce qui signifie que la croissance des bandes de glissements sature à une certaine hauteur. Un
paramètre, γirrev, est défini comme mesure de glissements locaux irréversibles en surface. Ce paramètre est
indépendant de la taille du champ de vision et fournit des informations sur la distribution de la déformation
de surface le long de la section calibrée, y compris des régions susceptibles de développer des fissures. 
2001 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.

Keywords: Atomic force microscopy (AFM); Copper; Fatigue; Slip bands; Crack nucleation

1. INTRODUCTION fatigue mechanisms were obtained from the study of


dislocation arrangements in copper single crystals.
Fatigue crack initiation and crack growth are the two
Since fatigue crack nucleation is a surface phenom-
main stages in the life of cyclically loaded structures.
enon, fatigue damage is most likely better charac-
Following the formation of a macro-crack, crack
growth and the occurrence of failure can be predicted terized by changes on the surface rather than by alter-
by fracture mechanics. Even though engineering ations in the interior of the material. Various concepts
models are available for predicting crack initiation, have been introduced to explain the formation of
damage evolution that leads to the formation of a extrusions and intrusions based on either the glide of
macro-crack is difficult to predict, because no easily dislocations out of the crystal or the accumulation of
measurable parameter uniquely describes the state of point defects in the bulk [1–6]. When large clusters
damage during this stage. of extrusions and intrusions form, they are referred to
Most significant advances in the understanding of as macro-PSBs and form either positive or negative
protrusions, which have also been called encroach-
ments.
† To whom all correspondence should be addressed. Tel.: In polycrystalline copper, dislocation arrangements
+1-404-894-2888; fax: +1-404-894-9140. are similar to those that develop in single crystals.
E-mail address: ashok.saxena@mse.gatech.edu (A. However, due to the constraint between grains, slip
Saxena)
‡ Formerly Graduate Research Assistant in the School of
readily occurs on multiple slip systems and results in
Materials Science and Engineering at Georgia Institute of three-dimensional networks of dislocations even at
Technology low strain amplitudes. The detailed analysis of the

1359-6454/01/$20.00  2001 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
PII: S 1 3 5 9 - 6 4 5 4 ( 0 1 ) 0 0 2 7 1 - 3
3756 CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE

mechanisms involved in fatigue deformation and a gage diameter of 8.9 mm (0.35 in) and the surface
crack nucleation is more difficult for polycrystalline was prepared by a combination of mechanical and
materials than for single crystals. Theories by chemical polish. The fatigue tests were performed at
Essmann and colleagues [4, 5] and Tanaka and Mura a constant strain rate of 0.005 s⫺1 and a stress ratio
[8] provide some insight on the interaction between R = ⫺1, with strain amplitudes of 0.161 and 0.255%
PSB and grain boundaries leading to crack nucleation. that yielded fatigue lives of 75 900 cycles and 6900
These models predict that high applied strain ampli- cycles with total cumulative plastic strains of 184 and
tudes and large grain sizes favor the formation of 79, respectively. Several tests were performed at both
intergranular cracks, and low strain amplitudes and strain amplitudes with interruptions at fractions of life
small grain sizes promote the development of trans- of 0.5 and 0.9 for the 0.161% amplitude and at 0.25
granular cracks in copper. However, when an for the 0.255% strain amplitude. Mechanical tests
environment is factored, intergranular cracking can were followed by sectioning of the specimens for
also be rationalized at smaller grain sizes and low post-test observations by SEM and AFM. The AFM
strain amplitudes because of the quasi-brittleness of used in this study was an Aris-3500 with a long range
the process associated with oxidation of copper. In scanning module METRIS-3070. Due to the curva-
fact, intergranular cracks have been observed in cop- ture of the fatigue specimens and the limited vertical
per for a wide range of applied strain amplitude and range of the AFM (about 7 µm), the fields of view
grain sizes [5, 9–12]. (FOV) were limited to 30×30 µm2. About twenty
The potential of high resolution devices capable of scans were analyzed along the length and at several
quantitatively describing the surface topography, such positions around the circumference of each specimen.
as scanning tunneling microscopy (STM) and atomic The collection of AFM images was submitted to a
force microscopy (AFM), was discovered in the mid verification procedure described elsewhere [20] to
1990s and used for accurately measuring the height ensure that the area covered by the AFM scans was
of surface features in various materials [13–19]. representative of the whole surface. If the statistical
Quantitative parameters were defined from these requirements were not met, that is, the area covered
measurements to describe the state of surface defor- was deemed too small, additional AFM scans were
mation, such as the average slip distance [13] or the obtained.
ratio of the average height to the average spacing
between the surface features [14] and root mean
3. RESULTS AND DISCUSSION
square (RMS) height of the surface [16]. A direct
relationship exists between these parameters and the In the following, the various surface features of
average accumulation of surface plastic strain, which fatigued specimens of copper are first identified by
was exploited in these studies to determine the irre- SEM and broad set of AFM measurements are also
versibility of slip at the surface. Although, the use described. This allows us to place our results in con-
of average values of surface parameters is correct to text of previously published results on copper.
describe the general state of surface damage, the Detailed AFM results follow this preliminary dis-
extension of their use to the determination of the cussion and lead to the development of a formalism
onset of the nucleation of fatigue cracks, which is by to quantitatively characterize the surface deformation.
nature a heterogeneous process, is beyond the capa-
3.1. Observation of surface damage
bilities of such parameters. The objective of this study
is to exploit the capabilities of AFM to accurately The most common occurrence of surface defor-
perform an analysis which takes into account length, mation in cyclically loaded polycrystalline copper is
height, count and orientation of the surface features in the form of slip bands regularly distributed within
over a significant portion of cyclically loaded copper grains. An example of a typical arrangement of slip
specimens. This will provide the basis necessary for bands covering the entire width of a large surface
the development of a model describing the evolution grain in copper is shown in Fig. 1, taken from a speci-
of the topography that can be used for determining the men tested at 0.161% strain amplitude. In general, it
onset of fatigue crack nucleation based on physical was observed that the majority of the slip bands are
observations of the surface. The subsequent appli- extrusions and only a few intrusions are visible. From
cation of this model to cyclically loaded components AFM measurements, the height of slip bands was
has the potential to evaluate, from actual surface found to vary between 30 and 900 nm, with an aver-
measurements, how much of their fatigue life is spent. age height at about 200 nm. The highest slip bands
were found exclusively at the latest stages of life,
while small bands were present at all life fractions.
2. EXPERIMENTAL PROCEDURE
This suggests that both the creation of new slip bands
Strain controlled fatigue tests were performed on and their growth occur throughout the fatigue life.
high purity C101 grade polycrystalline copper The measured heights of the extrusions in polycrys-
(OFHC) with an average grain size of about 40 µm talline copper do not compare with those observed in
(estimated by the mean intercept length method). single crystals, which were on average between 3 and
Standard axial fatigue specimens were machined with 4 µm high [5, 7, 21, 22]. The smaller slip band sizes
CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE 3757

Fig. 1. Typical arrangement of slip bands at the surface of a large grain in cyclically loaded polycrystalline
copper at 0.161% strain amplitude. (b) The magnified region within the square shown in (a).

in polycrystalline material is, however, not contradic- damage with considerable bulging of the surface
tory to the findings on single crystals, because the (about 1 µm) (Fig. 3). The width of the measured
amount of material available below the surface to cre- protrusions was between 10 and 20 µm, creating a
ate the upset is far less than the entire specimen thick- height to width ratio smaller than measured in single
ness and width that is available for single crystals. crystals [7, 9]. This is consistent with findings on cop-
As previously mentioned, the development of slip per single crystals by Hunsche and Neumann [7] that
bands in a given grain depends on the orientation of showed a decrease in the height of the protrusions
the active slip systems within that grain. However, if with a reduction of the thickness of the single crys-
no active slip system is oriented towards the surface tals, explained by availability of less material to pro-
of the material or if the interaction of the several slip duce the upset. In the present study, the term pro-
systems creates a three-dimensional arrangement of trusion is used when PSBs occupy the greater part of
dislocations (e.g. cellular of labyrinth structure) a grain and the density of the slip bands forming the
which prevents the development of large scale slip PSB has reached a high enough level that no band of
along a single band, surface upsets cannot form and matrix is visible between the slip bands.
a smooth grain surface is observed. Fig. 2 clearly
3.2. AFM analysis of surface features
illustrates this phenomenon, where some grains have
fully developed sets of slip bands while the neighbor- As mentioned earlier, slip bands are the primary
ing grains have little or no surface deformation. form of surface damage and towards the end of life
In this study, slip bands were observed at both at the higher strain amplitude (0.255%) protrusions
applied strain amplitudes, although towards the end are also observed. In this section, these features are
of the fatigue life in 0.255% strain amplitude tests, quantitatively analyzed and, since they all evolve
protrusions became the dominant form of surface from the same surface deformation mechanism, no
distinction is made between them. For instance, a pro-
trusion consisting of ten slip bands is treated as a
group and individual dimensions of slip bands are
computed as the average value for the protrusion.
Three measures of surface damage evolution were
considered that include: (a) the length of slip bands;

Fig. 2. Distribution of slip bands of various orientations and Fig. 3. AFM image of a protrusion in polycrystalline copper
densities among the different grains, which shows regions (⌬⑀/2 = 0.255%). The vertical scale is indicated on the right of
where no surface deformation has occurred (⌬⑀/2 = 0.255%). the image in nanometers.
3758 CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE

(b) the number of slip bands; and (c) the height of bands per unit area is not subject to the same draw-
slip bands. backs and is therefore a suitable surface damage para-
Figure 4(a) provides the average length of slip meter, as shown in Fig. 4(b). For additional infor-
bands at a strain amplitude of 0.161% at various frac- mation on the evolution of surface deformation, the
tions of life and shown as a function of the angle average height of the slip bands as a function of their
that the slip bands form with the loading axis. This orientation with the loading axis is also computed,
representation was chosen to highlight the possible Fig. 4(c). The behavior for the strain amplitude of
influence that the orientation of the trace of the slip 0.255% was similar and was therefore not included
systems with the loading axis might have on the sur- in the figures.
face features. One notices from Fig. 4(a) that the aver- From Fig. 4(b), one notices a strong influence of
age length of slip bands does not vary substantially the angle of orientation on the number of slip bands.
with orientation or with the number of cycles. This At both strain amplitudes, the number of slip bands
correlates with the earlier observation that slip bands is much greater at angles between 30° and 80° with
cover the entire width of a grain and hence cannot the loading axis. Although these angles do not corre-
grow further. This, along with the fact that slip bands spond to the actual orientations of the active slip sys-
may be truncated by the edge of the FOV and appear tems, they are consistent with favorable slip orien-
shorter than they actually are, restricts the capacity tations for the fulfillment of the Schmid law. Indeed,
of the average length of slip bands to be used as a the trace of a plane oriented at 45° with the longitudi-
representative parameter of surface deformation in nal axis of a cylinder assumes a range of orientations
polycrystals. On the other hand, the number of slip comprised between 45° and 90° around the circumfer-
ence. This explains the orientations of slip bands
⬎45°. On the other hand, slip bands oriented closer
than 45° to the loading axis do not belong to a favor-
ably oriented slip plane and must therefore be the
result of constraint from neighboring grains that
locally disturbs the local strain and stress fields.
At 0.161% strain amplitude, the number of slip
bands is not measurably affected by the number of
cycles between N/Nf = 0.5 and 0.9, but increases
between N/Nf = 0.9 and 1. The increase in the number
of slip bands between half-life and failure is only
30%, which may imply that the multiplication of slip
bands is not the leading mechanism of fatigue damage
evolution. On the other hand, the height of the slip
bands is affected more significantly by the increase
in number of cycles, Fig. 4(c). At half-life, an average
height of about 180 nm is measured for all orien-
tations. It is noted from Fig. 4(c) that slip bands with
orientations between 45° and 90° with the loading
axis reach heights of about 250 nm at N/Nf = 0.9 and
up to 400 nm at failure. This tends to show that the
height of the slip bands is a good indicator of fatigue
damage evolution. This is justified by the theory that
slip bands create sharp discontinuities at the surface
that cause the stress concentration responsible for
crack nucleation and the higher the upset, the larger
the stress concentration. However, the transgranular
cracks that one would expect from these stress con-
centrations, and are observed in single crystals at the
edges of PSBs, were not present here. On the other
hand, based on surface observations, it is believed
that critical conditions that lead to crack nucleation
arise from the development of a large number of slip
bands with significant height in a grain adjacent to a
grain with little surface strain. It was shown above
how important the orientation of the active slip sys-
tem relative to the load axis is on the development of
Fig. 4. Distribution of (a) the average slip band length, (b) the slip bands and adjacent grains may have significantly
number of slip bands and (c) the average height of slip bands different lattice orientations, leading to a severe mis-
in polycrystalline copper tested at 0.161%. match in the surface strains between the neighboring
CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE 3759

grains, for example, Fig. 2, that is accompanied by


residual stresses. These, in turn, favor crack
nucleation along the grain boundary according to the
mechanisms proposed by Essmann et al. [4, 5] or
Tanaka and Mura [8]. This issue is further dis-
cussed below.
At 0.255% strain amplitude, the height of the slip
bands was not significantly affected by the orientation
of the slip bands relative to the loading axis for angles
of 30° and beyond. The average height at
N/Nf = 0.25 is 108 nm, which shows that a significant
growth occurs early in the fatigue life, and the height
then increases to 247 nm at failure. The latter com-
pares well to the average height of 271 nm obtained
at failure for the 0.161% strain amplitude. That value
is significantly smaller than the several µm measured
on single crystals, a fact that actually supports the
theory by Essmann et al. [4] that relates the surface Fig. 5. Comparison of the characteristics of protrusions and
stand-alone slip bands at (a) quarter-life and (b) failure
roughness to point defect formation in the bulk, which (⌬⑀/2 = 0.255%). The charts on the left indicate the relative
is limited by the size of the grains in polycrystals. As distribution of protrusions and slip bands as a function of the
a result, the growth of slip bands in polycrystalline angle with loading axis and, on the right, the fraction of the
copper with grains on average 40 µm in diameter total count and length of each feature is shown.
seems to saturate at a height of about 250 nm.
The number of slip bands per unit area at the covered by protrusions is hence overall smaller than
0.255% strain amplitude on the other hand is affected the surface covered by an equivalent number of slip
by the slip band orientation and is about three times bands. By combining the results of the proportion of
larger than at 0.161% strain amplitude. The signifi- slip bands that are not part of protrusions and the total
cantly larger number of slip bands at the higher strain number of all slip bands per unit area, one derives a
amplitude is due to two factors that are inter-related. count of 0.9×0.035 = 0.032 slip bands per µm2 at
First, because of the higher applied stress and the dis- N/Nf = 0.25 and only 0.5×0.050 = 0.025 slip bands
ruption of the stress field by adjacent grains that are per µm2 at failure for the slip bands that are not part
undergoing deformation, the resolved shear stress of of protrusions, while the total number of all slip bands
slip systems not ideally oriented may reach levels per unit area actually increases from 0.035 to 0.050
high enough to cause plastic deformation, hence µm⫺2. This decrease in the number of slip bands that
increasing the number of active slip planes. In are not part of protrusions shows that protrusion for-
addition, the formation of protrusions as shown in mation is an out-growth of a continuous process of
Fig. 3 occurs at this strain amplitude and their high multiplication of active slip planes within existing
slip band density increases the number of slip bands. slip bands and not the growth of closely packed slip
The characteristics of protrusions are discussed next. bands due to cyclic deformation in a grain. In
addition, the fact that the distribution of the slip band
3.3. Protrusions versus slip bands
length in and out of protrusions follows closely the
Figure 5 compares total count and length of stand- count distribution, Fig. 5, indicates that protrusions
alone slip bands versus those that are part of pro- are formed in grains of any size. Indeed, if protrusion
trusions at a strain amplitude of 0.255%. A total of formation was favored in larger grains and the num-
638 and 808 slip bands at N/Nf = 0.25 and 1, respect- ber of slip bands in and out of protrusions was the
ively, were measured. Early in the fatigue life, surface same, the fraction of the total length of slip bands
upset occurs principally in the shape of slip bands, within protrusions would be proportionally larger,
with protrusions covering only 10% of the total sur- which is not the case.
face upset in count and length, Fig. 5(a). The presence The orientations assumed by the protrusions are
of protrusions at this stage shows that their formation within the standard range of 30–80° that has been
is uniquely related to the magnitude of the strain observed for surface features for fatigued copper in
amplitude and not, for instance, to the cumulative this study. Very few protrusions have orientations
plastic strain or the height of the slip bands them- ⬍45° with the loading axis, whereas an appreciable
selves that are actually larger at the lower strain quantity of slip bands assumed orientations as low as
amplitude which did not produce any protrusions. 30°. This implies that the formation of slip bands may
At failure, the proportion of protrusions to slip be enhanced by the effect of constraint from neighb-
bands is 1:1 in terms of both number and length of oring grains and occur at various orientations with
slip bands, Fig. 5(b). However, protrusions within the loading axis, but protrusions develop only on slip
them contain a much higher density of slip bands than planes that have a favorable orientation relative to the
regular arrangements of slip bands and the area loading axis to satisfy the Schmid law.
3760 CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE

3.4. Slip band spacing crack nucleation at the grain boundaries as shown in
Fig. 7. In almost every case, cracks initiate at the
The analysis of the spacing between slip bands
boundary between one grain with considerable sur-
reveals an influence of the applied strain amplitude
face upset and another grain that does not show much
on the distribution of slip bands at the surface, Fig.
trace of surface deformation. Intergranular cracks at
6(a). Slip bands that are not part of protrusions are
various strain amplitudes and grain sizes were already
separated on average by about 1.6 µm at the higher
reported by other authors mentioned earlier [5, 9–11].
strain amplitude and 2.3 µm at the lower strain ampli-
Although not explicitly mentioned in those articles,
tude. This difference of about 25% does not change
most of the micrographs of interest showed that inter-
throughout the fatigue life, as the average slip band
granular crack nucleation occurred at the boundary
spacing remains constant at both strain amplitudes.
between two grains with significantly different
One has to note that these values correspond to aver-
amounts of surface deformation, as is the case in the
ages and that spacings vary significantly from one
present study. Incidentally, a study by Lin et al. [23]
region to another.
showed that the propagation of slip bands from one
At the 0.255% strain amplitude, the surface fea-
grain to another offers an alternative to crack
tures consist of either stand-alone slip bands or slip
nucleation along grain boundaries by relieving
bands that form protrusions. Obviously, the spacing
excessive stress at that location. They demonstrated
between them is strongly affected by this factor as it
that this type of stress relief occurs only when the
drops from over 1.6 µm outside to less than 1 µm
adjacent crystals have favorable relative orientations.
inside protrusions, Fig. 6(b). Here again, the number
These findings support the results from the current
of cycles does not influence the spacing between slip
study in which cracks form along grain boundaries
bands, as only a slight decline is measured throughout
between grains that have a strong mismatch in their
the fatigue life. The smaller spacing between slip
respective amount of surface deformation. In con-
bands illustrates the strong contribution of protrusions
clusion, the mismatch in the deformation of two
towards the accumulation of dense surface defor-
neighboring grains seems to be the driving force for
mation.
the nucleation of an intergranular crack.
3.5. Crack nucleation
3.6. Derivation of the irreversible surface defor-
Copper specimens cyclically loaded at 0.161 and mation parameter
0.255% applied strain amplitudes show systematic
The above characterization by AFM of the features
that developed at the surface of copper fatigued speci-
mens consisted of count per unit area and average
length and height. The emphasis of this section is on
developing a single surface parameter that
encompasses the diversity of the above measurements
and can be used to determine a criterion for crack
nucleation.
For general applicability, a parameter that charac-
terizes surface deformation must be independent of
the type of surface features (e.g. slip bands,
extrusions, protrusions, streaks, etc.) and the size of
FOV. This is essential because the dimensions of the
features of fatigue damage can range from a few nan-
ometers up to several micrometers and different sizes
of FOVs are then needed to obtain statistically sig-
nificant results. Let us first consider the simple case
of a single crystal oriented for single slip, Fig. 8. If
pulled monotonically in tension, the crystal experi-
ences an irreversible deformation, δ, that can be mea-
sured by the surface step height and is directly related
to the applied plastic strain amplitude, γ, by the fol-
lowing relationship

d
g⫽ (1)
L
Fig. 6. Comparison of (a) the spacing between slip bands that
are not part of protrusions for 0.161% and 0.255% strain ampli-
tude and (b) the spacing between slip bands at the 0.255% where L is the gauge length or length of the crystal
strain amplitude that stand alone (not part of protrusions) and in this case. The magnitude of the strain may vary
that are part of protrusions. if a different set of coordinates is chosen, or if the
CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE 3761

Fig. 7. Typical intergranular cracks that are formed in polycrystalline copper cyclically loaded at various strain
amplitudes. Cracks are usually formed at the boundary between one with grain with considerable surface upset
and another with much less visible deformation.


n

2|di|
i⫽1
girrev ⫽ . (2)
L

It is important to note that γirrev is an average of the


measurable irreversible deformation experienced by
the material over the gauge length L and not the
amount of strain within each slip band. Here, slip is
assumed to occur in the shape of slip bands, but the
same concept is applicable to other types of features
Fig. 8. Final crystal shape of a single crystal oriented for single
that create a surface step. The theory applies to all
slip after (a) one half cycle and (b) one cycle if slip occurs forms of surface deformation features, because its
on an adjacent crystallographic plane in the reverse loading concept is simply based on the permanent defor-
direction. mation of the surface that is the result of irreversible
plastic strain. The above definition of normalized sur-
face deformation is different from the one used by
emergence angle of the slip band changes. The above Harvey et al. [14], which was given by the ratio of
definition of strain based on d, the normal displace- the average height to the average spacing between
ment of the surface deformation, was chosen because slip bands. The latter is actually a measure of the
it is readily measurable by AFM. The use of an average strain contained within the slip bands, while
adequate geometric factor may improve the accuracy equation (2) is a measure of normalized surface defor-
of g, but the factor will most likely remain constant mation. When regular arrays of slip bands develop
throughout the fatigue life. Therefore, its omission, as at the surface both definitions yield similar results.
implied by equation (1), will not affect monitoring of However, the definition by Harvey et al. cannot be
the relative changes of irreversible surface defor- applied to a case where isolated slip bands form,
mation with fatigue cycles. because it requires a measure of spacing between sur-
In the same crystal, if slip occurs in the reverse face features, while a measure of surface deformation
loading direction on a different but parallel slip plane, is still possible with the definition given by equation
the final crystal shape may resemble that of Fig. 8(b) (2). Consequently, the advantage of the latter defi-
and the residual plastic strain is then twice the amount nition is its potential to analyze regions with little or
given in equation (1), even though the crystal seems no surface deformation, which permits the analysis of
to have returned to its original shape at the macro- the distribution of deformation over the entire sample
scopic scale. According to Mura [24], slip occurs on surface and not only of selected FOVs that contain
adjacent layers of materials in the forward and surface deformation.
reversed loading directions due to accumulation of One significant difference between single crystals
back stress, which contributes to the formation of sur- and polycrystalline materials is that a slip band does
face upset (e.g. slip bands) separated by regions with not cross the full width of a specimen and, therefore,
no apparent slip. Subsequently, if n slip bands are its contribution to the total surface deformation not
created by cyclic loading, the irreversible plastic only depends on its height but also on its length,
strain is given by which must be included in the measure of surface
3762 CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE

deformation. For example, Fig. 9 shows a FOV of



m

width W and length L where a single slip band of gnorm


j

height δ developed across the width a of a grain at girrev ⫽


j⫽1
(5)
an angle α with the loading axis. To account for the

m

fact that the slip band does not always cover the com- Lj
j⫽1
plete width of the crystal, it is proposed to normalize
the slip band length by the width of the FOV in the
direction parallel to the slip band, or where the denominator represents the total length of
the FOVs, which is effectively the gauge length.
a The above definition of γirrev is not affected by the

冉 冊
anorm (3) size of the FOVs, because the parameter is nor-
W L
Min , malized relative to both the width W and the length
sin a cos a
L of the FOV. However, an appropriate choice of size
for the FOVs is critical to reach the best combination
of resolution and range necessary for accurate
where α is the angle between the slip band and the measurements of the surface features over a surface
loading axis and Min takes the smallest value listed area that is as large as possible. Indeed, if a small
as arguments. Now, the amount of the normalized size is selected, more FOVs will be necessary to meet
surface upset in one FOV containing n slip bands is the aforementioned validation criterion that ensures
obtained by the summation of the contribution of each that the measurements are representative of the entire
slip band, that is: surface. Conversely, the resolution is lowered on
large FOVs and small surface features may not be


n
included in the measurements.
FOV ⫽
dnorm 2|di|anorm
i (4)
i⫽1 3.7. Distribution of irreversible surface deformation
Due to the high level of heterogeneity of the sur-
face roughening process, average values of the irre-
If a constant normalizing factor had been chosen in versible surface deformation are not representative of
equation (3) instead of the width of the FOV, such the regions with high levels of deformation. There-
as the grain size, all FOVs would have to be the same fore, it is virtually impossible to predict experimen-
size to compare results from equation (4) but, by nor- tally the location of these regions and monitor
malizing relative to the width of the FOV, it is poss- changes in topography at the precise location where
ible to perform the analysis with different size FOVs. the fatal crack will nucleate. However, even though
This is useful if different surface features develop at the local maximum of γirrev cannot be directly meas-
various length scale (e.g. [20]) and different magnifi- ured, the statistical method of describing the hetero-
cations are required to fully analyze the surface defor- geneity of surface slip irreversibility described below
mation. To convert the surface upset in equation (4) is capable of estimating its magnitude and thus has
to a normalized surface deformation (strain), one the potential for detecting the onset of fatigue crack
needs to divide the amount of normalized surface nucleation. This concept, which is substantiated by
upset by the entire gauge length. In other words, when experimental data, assumes that surface damage
considering m FOVs, it becomes resulting from cyclic deformation occurs in a random
fashion, which implies that only a few regions have
extreme amounts of surface upset (high or low) and
the majority of the surface has a local amount of dam-
age that is close to the sample average. As a result, by
statistically characterizing the distribution of surface
deformation in a large number of FOVs, it is possible
to estimate the amount of damage in the region with
maximum local surface damage, even though that
specific region is not directly included in the FOVs.
Figure 10 is an example of the distribution of γirrev
at a quarter of the fatigue life and failure for the speci-
mens tested at 0.255% strain amplitude. The horizon-
tal axis represents ranges of γirrev and the respective
fractions of the surface that have reached these ranges
are indicated on the vertical axis (i.e. the sum of all
Fig. 9. Example of a FOV containing several grains and a sin- the bars for a given condition amounts to 100% of
gle slip band of height d and length a. The normalization of
the slip band length is done relative to the width of the FOV the surface). The distribution of the surface defor-
in the direction parallel to the slip band, which can be determ- mation of all specimens resembles those shown in
ined from the angle a with the loading axis. Fig. 10, which are analogous to a Gaussian (normal)
CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE 3763

under the curve equal to one, corresponding to 100%


of the specimen’s surface. Other statistical functions
may also be used to characterize the distribution of
surface deformation. For example, applying a Weibull
distribution to the current set of data did not notice-
ably affect the results from this analysis. Additional
measurements would be necessary to determine with
certainty the type of distribution that surface strains
assume during uniaxial cyclic loading of a polycrys-
talline material.
From the plots of Fig. 11(a), two main observations
are made on the evolution of the distribution of the
irreversible surface deformation. First, the average
Fig. 10. Distribution of the irreversible surface deformation on value of the normal distribution curves increases with
the surface of fatigue specimens tested at a strain amplitude
of 0.255%. These plots are based on the irreversible surface
the number of cycles, which is consistent with the
deformation measurements collected from the multiple fields results from the previous section on the average
of view analyzed on each specimen. values of irreversible surface deformation and, sec-
ondly, the spread of the curves becomes larger as the
distribution about the average value of the irreversible number of cycles increases. The latter is due to the
surface deformation. Consequently, the data can be development of substantial surface upset in some
equivalently represented with a continuous normal regions, while others do not experience much pertur-
distribution function, Fig. 11(a), which has the advan- bation of their surface topography due to unfavorable
tage that it can be easily integrated and has an area local crystallographic orientations of the slip planes
and/or constraint effects between grains. The regions
with significant increase of the surface upset eventu-
ally become zones of instability where crack
nucleation is possible.
Since the normal distribution curves of Fig. 11(a)
reveal the extreme values of irreversible surface
deformation reached in some regions, these plots
directly provide information on the advancement of
surface upset in the highly deformed regions. Thus,
this data can be used to predict when the material
reaches a critical state that will trigger the nucleation
of a fatal crack. This analysis is however better per-
formed on the cumulative version of these plots
shown in Fig. 11(b), which is simply the integral of
the normal distribution curves. The vertical axis on
the left indicates the portion of the surface that con-
tains amounts of irreversible surface deformation
between zero and the amounts indicated on the hori-
zontal axis of the graph. Conversely, the vertical axis
on the right provides the portion of the surface that
has developed amounts of irreversible surface defor-
mation larger than the amount indicated on the hori-
zontal axis. With the information provided in Fig.
11(b), one can assess the evolution of the deformation
distribution over the specimen surface, including in
the regions where extreme values of surface defor-
mation have developed.
3.8. Crack nucleation criterion
The previous AFM and SEM observation showed
that fatigue cracks nucleate in locations where the
Fig. 11. Representation of the distribution of the irreversible surface deformation is high. Since γirrev is capable of
surface deformation at the surface of fatigue specimens using determining the maximum local values of surface
(a) a normal (Gaussian) distribution function and (b) a normal deformation, it is ideally suited as an indicator of the
cumulative distribution of the irreversible surface deformation.
In (b), the right axis corresponds to one minus the left axis, probability of fatigue crack nucleation. One may
and the vertical dotted lines represent the critical values of the argue that the magnitude of surface upset is not
irreversible surface deformation. always the driving force for crack nucleation. For
3764 CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE

example, Polák et al. [21] argue that, in copper single mately 30%. Ideally, these experiments should be
crystals, crack nucleation is driven by the creation of repeated with in situ AFM measurements where the
point defects in the bulk and occurs through the link- progression of surface damage can be followed on the
ing of intrusions that are formed at the edges of same specimen at several N/Nf values.
extrusions. In this mechanism, γirrev is not a direct
measure of damage that leads to crack nucleation but,
because the forms of damage during fatigue (e.g. 4. CONCLUSIONS
point defect creation and development of surface From the SEM and AFM analyses of the surface
upset) occur concurrently, monitoring one provides a deformation and crack nucleation behavior of
good indicator of the other, and thus, the overall state polycrystalline copper tested at 0.161% and 0.255%
of fatigue damage. Therefore, γirrev is a relevant para- strain amplitude, the following conclusions were
meter for monitoring all types of fatigue damage and drawn.
a criterion for crack nucleation may be defined in A strong influence of the orientation of the slip
terms of a critical local value of γirrev necessary for bands was observed at both strain amplitudes, as the
crack nucleation. This critical value of γirrev can be number of slip bands that have an angle with the load
determined from AFM measurements by measuring axis between 30° and 80° was much larger than for
the value of γirrev of FOVs that contain a crack other orientations. Both strain amplitudes yield simi-
nucleus. Once known, this value can be traced on the lar average slip band height at failure of about 250
normal distribution curves of the irreversible surface nm, which indicates that the growth of slip bands
deformation, as shown by the vertical lines shown on saturates at this height in polycrystalline copper with
Fig. 11(b). It was found that the levels of local surface a grain size of 40 µm. Protrusions were present at the
deformation necessary to trigger the nucleation of a surface of copper specimens tested at 0.255% strain
fatigue crack are different between the high and the amplitude and developed in grains of various sizes,
low strain amplitude tests. from a few to several hundred µm. They develop
The criterion described above does not act like a throughout the fatigue life, with the fraction of slip
“failure or no failure” switch that would imply that a bands that are part of protrusion increasing from 10%
crack is always nucleated once the critical value of at N/Nf = 0.25–50% at failure. Protrusions appear to
surface deformation is achieved in a specific location. form by multiplication of active slip planes within
On the contrary, Fig. 11(b) clearly shows that a cer- existing slip bands.
tain portion of the surface may develop levels of sur- Intergranular crack nucleation occurs at both strain
face deformation well beyond the critical value with- amplitudes. Crack nuclei systematically develop at
out nucleating a fatal crack, as observed in all failed grain boundaries between a highly deformed grain
specimens and in some specimens tested to 90% of and one with little evidence of surface upset. The
the life. The actual fraction of the surfaces that has strain mismatch between the two grains seems to be
reached or exceeded the crack nucleation criterion is the driving force for intergranular crack nucleation.
indicated by the ordinate of the intersection of the The definition of the surface deformation para-
distribution curves with the vertical line that specifies meter, γirrev, was formulated to examine the distri-
the criterion. Now, if one postulates that the likeli- bution of surface deformation over the surface of any
hood of nucleating a crack increases proportionally material, independently of the type of surface upset.
with the fraction of the surface over which the critical This approach provides a means to estimate the
level of surface deformation is exceeded, the distri- maximum levels of surface deformation developed by
bution curve of the irreversible surface deformation critical regions where fatigue crack nucleation is
provides an estimate of the probability for crack likely to occur. A criterion for crack nucleation was
nucleation. For example, when several identical defined as follows: fatigue crack nucleation is poss-
fatigue experiments are performed, failure does not ible once the material develops, on a local scale, sur-
necessarily occur after the same number of cycles face deformation of a magnitude that exceeds a criti-
because of the individuality of each specimen and the cal value. This critical value can be determined from
ability to predict failure at that point becomes an exer- AFM images that contain crack nuclei.
cise of probability. According to the criterion defined
here, the probability of failure is null as long as γirrev
has locally not exceeded its critical value at least in Acknowledgements—The authors are grateful for the financial
some regions. On the other hand, as regions with fav- support of the Office of Naval Research under the M-URI Pro-
gram “Integrated Diagnostics” (Grant No. N00014-95).
orable lattice orientations for formation of surface
upset develop levels of surface deformation beyond
that critical amount, the nucleation of a fatal crack REFERENCES
becomes possible, and a measure of the probability
that the specimen has nucleated a fatal crack may be 1. May, A. N., Nature, 1960, 186, 573.
2. Basinski, Z. S. and Basinski, S. J., Acta metall., 1985,
directly read from the plots of Fig. 11(b). 33, 1319.
According to the above theory, the probability of 3. Rosenbloom, S. N. and Laird, C., Acta metall mater., 1993,
failure at N/Nf values of 1 in this study was approxi- 41, 3473.
CRETEGNY and SAXENA: SURFACE DEFORMATION DURING FATIGUE 3765

4. Essmann, U., Gösele, U. and Mughrabi, H., Phil. Mag. A, 214C, 1151.
1981, 44, 405. 17. Saxena, A., Yang, F. and Cretegny, L., in Proc. Seventh
5. Mughrabi, H., Wang, R., Differt, K. and Essmann, U., in Int. Fatigue Congress, ed. X. R. Wu and Z. G. Wang, Beij-
Fatigue Mechanisms: Advances in Quantitative Measure- ing, P.R. China, 1999, Vol. 4, p. 2777.
ments of Physical Damage, ASTM STP 811, ed. J. Lank- 18. Jono, M., in Proc. Seventh Int. Fatigue Congress, ed. X.
ford, D. L. Davidson, W. L. Morris and R. P. Wei. Amer- R. Wu and Z. G. Wnag, Beijing, P.R. China, 1999, Vol.
ican Society for Testing and Materials, USA, 1983, p. 5. 4, p. 57.
6. Ma, B. -T. and Laird, C., Acta metall., 1989, 37, 325. 19. Man, J., Obrtlı́k, K., Lopour, F., Blochwitz, C. and Polák,
7. Hunsche, A. and Neumann, P., Acta metall., 1986, 34, 207. J., in Proc. Seventh Int. Fatigue Congress, ed. X. R. Wu
8. Tanaka, K. and Mura, T., J. Appl. Mech., 1981, 48, 97. and Z. G. Wnag, Beijing, P.R. China, 1999, Vol. 4, p.
9. Kim, W. H. and Laird, C., Acta metall., 1978, 26, 777. 157.
10. Kim, W. H. and Laird, C., Acta metall., 1978, 26, 789. 20. Cretegny, L. and Saxena, A., Fatigue Fracture Eng. Mater.
11. Figueroa, J. C. and Laird, C., Mater. Sci. Eng., 1983, Structures, 2001 (submitted for publication).
60, 45. 21. Polák, J., Lepistö, T. and Kettunen, P., Mater. Sci. Eng.,
12. Christ, H. -J., Mater. Sci. Eng., 1989, A117, L25. 1985, 74, 85.
13. Sriram, T. S., Ke, C. -M. and Chung, Y. W., Acta metall. 22. Basinski, Z. S. and Basinski, S. J., Acta metall., 1985,
mater., 1993, 41, 2515. 33, 1307.
14. Harvey, S. E., Marsh, P. G. and Gerberich, W. W., Acta 23. Lin, T. H., Wong, K. K. F., Teng, N. J. and Lin, S. R.,
metall. mater., 1994, 42, 3493. Mater. Sci. Eng. A: Struct. Mater.: Properties, Microstruct.
15. Gerberich, W. W., Harvey, S. E., Kramer, D. E. and Process., 1998, A246, 169.
Hoehn, J. W., Acta mater., 1998, 46, 5007. 24. Mura, T., Mater. Sci. Eng. A: Struct. Mater.: Properties,
16. Yang, F. and Saxena, A., Proc. Inst. Mech. Engrs, 2000, Microstruct. Process., 1994, A176, 61.

You might also like