You are on page 1of 12

Materials Science and Engineering A316 (2001) 115– 126

www.elsevier.com/locate/msea

Experimental study of porosity and its relation to fatigue


mechanisms of model Al–Si7–Mg0.3 cast Al alloys
J-Y. Buffière a,*, S. Savelli b, P.H. Jouneau a, E. Maire a, R. Fougères a
a
GEMPPM UMR, CNRS 5510, INSA, Lyon, 69621 Villeurbanne Cedex, France
b
Pechiney CRV Parc Economique, Centr’Alp, Voreppe BP 27 38340, France

Received 17 November 2000; received in revised form 14 February 2001

Abstract

The microstructure and fatigue properties of three model AS7G03 cast aluminium alloys containing artificial pores have been
studied. Synchrotron X-ray tomography has been used to characterise in three dimensions the pore population in the alloys. The
development of fatigue cracks in relation with local crystallography has been studied by means of electron back scattered
diffraction (EBSD). Both the average number of cycles to failure and the lifetime scatter depend on the pore content specially at
high stress level. The mechanism leading to the initiation of a crack from a pore has been identified. The crack propagation at
high stress level appears to be quite insensitive to microstructural barriers and can be reasonably well described by a Paris type
law. At low stresses, however, short cracks are often observed to be stopped at grain boundaries and the fatigue life is no longer
predicted by a simple propagation law. © 2001 Elsevier Science B.V. All rights reserved.

Keywords: Electron back scattered diffraction; Fatigue; Porosity; X-ray tomography; Crack

1. Introduction pore independently of the loading conditions and of the


stress [2–6]. Almost all pores leading to crack initiation
Cast aluminium alloys can be easily processed for are located at the surface of the material or just beneath
manufacturing objects with an intricate shape. How- it [6–9].
ever, because aluminium contracts when it solidifies, The number of cycles necessary to initiate a crack
and, also, because hydrogen dissolves very easily in the depends on the loading conditions; it generally ranges
molten metal, cast aluminium alloys always contain a from almost 0% (quasi immediate initiation) [6,9] to less
certain amount of porosity when no special casting than 10% of the fatigue life [10]. The detailed mecha-
technique is being used. This porosity has a detrimental nisms of the initiation of a crack from a pore are
influence on the mechanical properties of the material. however still not well elucidated.
For instance, in the case of the AS7G03 alloy studied After the initiation stage, the crack propagation
here, it has been shown [1] that 1% volume fraction chronology can be divided schematically in four steps:
porosity can lead to a reduction of 50% of the fatigue (i) interdendritic crack propagation alternatively in the
life and 20% of the endurance limit compared with the a matrix or along the Si particles/matrix interfaces
same alloy with a similar microstructure but showing [11 –13]. (ii) intradendritic crack propagation [5]; (iii)
no pores. As a result, a considerable amount of re- rapid interdendritic propagation with silicon particle
search has been carried out on the fatigue mechanisms damage ahead of the crack tip [14,15] and final rupture.
in cast aluminium alloys in relation with porosity. When the crack is small and still in the neighbourhood
Those mechanisms can be roughly summarised as of the pore it is sometimes observed to slow down or
follow: in almost all cases, fatigue cracks initiate on a stop with a spatial frequency close to the dendrite arm
spacing (DAS) [16]. For longer cracks (typically with a
* Corresponding author. Tel.: + 33-472-438854; fax: +33-472-
size above 200 mm [7]) the crack growth rate is found
438539. similar to that of long cracks observed for instance
E-mail address: buffiere@insa-lyon.fr (J.-Y. Buffière). during compact test (CT) fatigue tests.

0921-5093/01/$ - see front matter © 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 1 - 5 0 9 3 ( 0 1 ) 0 1 2 2 5 - 4
116 J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126

Various metallurgical factors others than porosity chemical composition, shown in Table 1, was common
have been observed to influence the fatigue properties of for the three alloys which were chill mould cast in a
cast Al alloys like the refinement of the DAS [2,8], the silicon carbide (SiC) mold. The artificial pores were
presence of intermetallic particles [4,17,18], the material obtained by introducing at 1093 K a mixture of hydrogen
heat treatment [19]. It must be pointed out that the effect (H2) and argon (Ar) gases in the liquid metal through a
of grain size on fatigue properties of cast alloys has been propeller. The solubility of hydrogen being an increasing
very poorly studied although it is now recognised that function of temperature, a part of the gas content was
grain boundaries play a major role on the crack growth rejected (degassing) during the solidification and formed
rate in wrought alloys [20– 23]. artificial gas pores. Three different H2/Ar ratio (2%, 7%
From this literature review, it appears that, in cast Al and 10%) were used giving different pore contents and
alloys, porosity is the key factor which controls the sizes in the alloys which will be referred to as alloys A,
fatigue properties. Thus, any attempt for modelling and B and C, respectively. After cooling, the alloys were
predicting the cyclic properties of such material requires solution treated during 10 h at 813 K, quenched in cold
a thorough characterisation of the pore population in the water, and aged during 6 h at 333 K (T6 heat treatment).
material. For the moment such a characterisation is The microstructure of the alloys before testing has been
carried out by means of two dimensional (2D) metallo- characterised using optical microscopes (OM) but also
graphic sections from which the actual three dimensional scanning and transmission electron microscopies (SEM
(3D) sizes and shapes of pores are very difficult to and TEM).
determine. Besides, only little attention has been paid, so
far, to the relation between fatigue mechanisms and the 2.2. X ray tomography and radiography
local microstructure such as grain boundaries, eutectic
particles or the local crystallographic orientation.
High resolution X-ray tomography was performed on
The recent development of electron back scattered
the ID 19 beamline at the European Synchrotron Radi-
diffraction (EBSD) technique has enabled the character-
ation Facility (ESRF), in Grenoble, France. It had been
isation of the effect of local texture on the short fatigue
shown previously that this non-destructive technique
crack growth [21]. Besides, high resolution X-ray tomog-
provides reliable 3D images of internal pores in the
raphy using synchrotron X-ray sources can now be used
studied material [25]. The principle and specificity of the
to obtain 3D images of internal defects such as pores or
tomography technique using synchrotron radiation and
cracks with a resolution in the micrometer range [24]. In
this paper, both techniques are combined in order to some recent applications to the characterisation of struc-
investigate the fatigue mechanisms of a model cast Al tural materials can be found elsewhere [26], and only the
alloy containing different porosity levels. From the 3D main experimental details are reported here. The chosen
images of the material obtained by tomography, exper- energy was 23 keV. The investigated samples were small
imental pore distributions are determined and correlated cylindrical samples (3 mm in diameter) with no polishing
with fatigue tests results. The initiation and propagation of the surface. For each sample, 600–900 2D projections
of fatigue cracks are studied in detail. A mechanism for were recorded on a 1024×1024 CCD camera developed
the initiation of a crack from a pore is described. at ESRF and set at 80 cm behind the sample. Those
Eventually, the retardation of cracks by grain boundaries projections were used to reconstruct numerical images of
is shown experimentally through a detailed analysis of the samples using a classical filtered back projected
the local texture by EBSD. algorithm. The resolution of the technique, i.e. the
volume of the smallest element (voxel) in the recon-
structed 3D images was in that case (6.65 mm)3. The
2. Experimental procedure volume of material investigated with this technique was
around 30 mm3.
2.1. Material A dedicated software performing 3D image analysis
developed at the LISA laboratory in Lyon was used to
Three model AS7G03 cast alloys with different con- study the pore population for the three alloys. The grey
tents of artificial pores have been used in this study. The level images obtained after reconstruction were

Table 1
Chemical composition of the studied alloys (wt.%)

Si Fe Mg Ti Sb Cr Others

6.6–7.0 0.09–0.14 0.29–0.34 0.11–0.14 0.12–0.14 0.2 B0.01


J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126 117

thresholded, in order to separate the pores from the rest to be around 40 nA. In these conditions, the probe
of the material. The resulting binary images were auto- diameter is about a few nanometers, with an interaction
matically analysed, giving the concentration of pores in volume underneath the specimen surface of the order of
the studied block as well as individual parameters such 1 m3. The spatial resolution for aluminium samples was
as volume, surface, aspect ratios, axis of inertia, etc. about 1 m, and the angular resolution for the crystallo-
For alloy A, plain numerical 2D radiographs were graphic orientations measurement was about 1°. The
also recorded (using the same specimen geometry) to FEG microscope provides the advantages of a more
investigate a larger volume of material and increase the stable probe current, and of a slightly better spatial
probability of encountering large pores in the specimen. resolution. A sample preparation was also needed for
The resolution of those numerical radiographs was 6.65 EBSD analysis in order to remove the plastically de-
mm. With this technique, the investigated volume in- formed layer induced by mechanical polishing. This was
creased to 1000 mm3. done by an electrochemical etching of the surface, using
a Struers Lectropol equipment and a Struers A2
2.3. Mechanical testing and characterisation solution.
Finally a smaller (9 mm2 section) fatigue specimen
Hour glass shaped specimens with a 24 mm2 cross has been characterised in tomography. It has been
section have been used to perform constant stress am- subsequently fatigued at 145 MPa during 233 000 cycles
plitude tests on an INSTRON 8516 servo-hydraulic and the initiation and propagation of fatigue cracks at
fatigue machine. The tests were carried out at room its surface has been studied by optical microscopy as
temperature using a frequency of 10 Hz and a load
described above.
ratio R= 0.1. Before testing, the samples flat faces were
mechanically polished using SiC sand paper (down to
2400 grit) and diamond paste (3 and 1 mm). The stress
3. Results
levels investigated ranged from 240 MPa down to 130
MPa. For each stress level either five or four specimens
3.1. Microstructural characterisation
were tested. After testing, the fracture surfaces of the
broken specimens were studied by SEM in the sec-
The microstructure common to the three alloys con-
ondary electron mode.
sist in aluminium a dendrites separated by eutectic
In order to investigate the crack initiation and propa-
zones containing silicon particles (cf. Fig. 1). The mean
gation mechanisms, the surface of some specimens was
value of the DAS for the three alloys is 30 mm. The
characterised by OM during the fatigue tests. This was
silicon particles have an average size (deduced from
done in the unloaded state by removing the sample
optical 2D measurements) of 3.1 mm. TEM observa-
from the tensile rig. The examination were carried out
tions show that those particles are polycrystalline and
after 1000 cycles, 10 000 cycles and then every 25 000
that the interface with the a matrix contains no sec-
cycles. Such tests were interrupted when a main crack
had attained a size sufficient to be in the long crack ondary phase. Some iron based intermetallic particles
regime (typically 400 mm) where it became quite insensi- are also observed; they appear to play no role in the
tive to microstructural effects. crack initiation or propagation because of the low iron
The local crystallographic structure around cracks content of the alloy.
was studied by EBSD in a scanning electron micro- The grain structure of the alloys has been studied
scope (see for example [27] for a review of this tech- from EBSD observations and also from optical micro-
nique). Two microscopes were used for this work. The graphs. In the latter case, the surface of the material
first one was a Philips XL30 fitted with a Schottky field was put into contact with gallium at room temperature.
emission gun (FEG). A silicon intensified target camera The mechanical effect of a further mechanical polishing
was used to record Kikuchi patterns on the phospho- resulted in some decohesions at the grain boundaries
rous screen, and the indexation was carried out with the which were therefore easy to observe. The grains were
‘Orientation Imaging Microscopy’ software from observed to be dendritic in nature with a random
TexSEM Laboratories [28,29]. The second microscope texture. Only a few fine pure eutectic grains (:3 mm)
was a JEOL 840A LGS fitted with a tungsten filament, could be observed; the majority of those grains growing
a CCD camera, and the ‘CHANNEL’ software from in epitaxy with the dendritic ones. The mean grain size
HKL Technology for the automatic acquisition and at the sample surface was found equal to 310 mm with
indexation of Kikuchi patterns. Orientation data have a standard deviation of 122 mm.
been analysed with a home build software [30]. As already mentioned, pores represent a major mi-
In both microscopes, an accelerating voltage of 30 crostructural feature of the material. On the basis of the
kV was used, and the probe current was usually chosen 3D tomographic images of the material, a quantitative
118 J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126

Fig. 1. Optical micrograph of the microstructure of alloy A.

Fig. 2. Classification of the pores in the three studied alloys ( , alloy A; , alloy B; × , alloy C) as a function of their equivalent size u and
their sphericity s (see the text for the definitions of these quantities). The 3D volume rendering of some typical pores have been added.

study of the pore sizes and shapes has been carried of volume fraction, for the three alloys. Those pores
out for the three alloys. Two geometric parameters, correspond very probably to micro-shrinkages which
the equivalent size and the sphericity have been appear at the end of the solidification. The second
defined to help classifying the pores. The equivalent population, containing pores with sizes larger than 50
size l corresponds to the diameter of a sphere having mm shows a strong correlation between size and
the same volume as the studied pore and the spheric- sphericity: the larger the size, the lower the sphericity.
ity s is a normalised ratio between the pore surface Their volume fraction increases with the amount of
and its volume (1\s \0 with s tending to 0 when the hydrogen introduced in the liquid metal. They corre-
pore shape gets tortuous) [25]. A classification of the spond to artificial pores appearing during the whole
pores using those two parameters is shown in Fig. 2. solidification process. Their growth is geometrically
This figure shows that two populations of pores hindered by the solidifying dendrites and this results,
co-exist in the material. The first one, containing a for the larger ones, in a very tortuous shape (low
majority of pores with an equivalent size lower than sphericity). The volume fraction of those pores in-
50 mm, exhibits a weak correlation between size and creases exponentially with the H2 content as already
sphericity and is almost equally represented, in terms observed in the literature [31].
J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126 119

Almost all the pores are intergranular; only a few


small ones are purely intragranular, at least at the
surface of the material. The number of grain boundaries
starting from the surface of a pore increases with the
pore size. Besides, the convex parts of the pores are in
all cases associated with Si eutectic particles (see Fig.
3a), and also with at least one grain boundary. This
typical arrangement of pores, Si particles and grain
boundaries, represented schematically in Fig. 3b, will be
shown to be of crucial importance with respect to the
crack initiation mechanisms.

Fig. 5. Optical micrograph of the surface of a sample of alloy C after


15 000 cycles at 160 MPa showing the initiation of a crack by
decohesions of Si/Al interfaces (arrows) at the vicinity of the convex
part of a pore (not entirely visible).

3.2. Fatigue tests

The pseudo Wöhler curves obtained for alloys A, B


and C are shown on Fig. 4. In all cases, a non linear
increase of the logarithm of the mean fatigue lives is
observed when the stress amplitude decreases. More
precisely, for a given stress level (between 240 and 130
MPa), the mean value of the number of cycles to failure
decreases when the amount of gas introduced in the
alloy increases. Regarding the scatter on the fatigue
lives, it increases when the stress level decreases, for all
alloys. Finally, for a given stress level, the scatter
increases when the gas content decreases. Those results
clearly indicate that the number of pores per unit
volume and the pore size distribution are major factors
Fig. 3. Typical spatial arrangements of Si eutectic particles with which control the fatigue life and its scatter.
respect to pores. (a) SEM micrograph of the interior of a pore
showing the Si eutectic particles (arrow) located in the convex parts
of the pores. (b) Schematic 2D illustration of this arrangement. Grain 3.3. Crack initiation
boundaries have been added.
The fractographic analysis of the broken samples
reveals in nearly all cases that the crack responsible for
the final fracture initiates from at least one pore located
at the surface of the specimen or just below it. In some
rare cases (high stress levels, 240 MPa) crack initiation
can occur by the decohesion of internal Si particles
probably located within a grain boundary. By studying
carefully the fracture surface near the pores responsible
for the rupture by SEM or by OM, it appears that crack
initiation (defined here as a visible crack in the a
matrix) is induced by a decohesion at the Si particle/ma-
trix interface. In most cases, this decohesion occurs on
Si particles located very close to one convex part of a
pore lying nearly perpendicular to the applied stress. A
Fig. 4. S – N curves of alloys A, B and C alloys (, alloy A; , alloy
typical example of such a decohesion and the resulting
B; , alloy C). Although the same stress levels have been used for the
three alloys the points on the graph have been slightly shifted along crack is shown on Fig. 5. At low and high stress levels
the stress axis for the sake of clarity. The slopes of the curves for the number of cycles necessary to initiate such cracks is
alloys A, B and C are respectively: − 0.131, − 0.168 and − 0.158. always negligible with respect to the fatigue life (B2%).
120 J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126

The presence of a grain boundary favours crack 10–20 mm and stop. Only a small proportion of those
initiation. This point is illustrated on Fig. 6 which cracks manage to grow further. When this happens the
shows a micrograph of a pore (called P1) which has crack growth is discontinuous with rather frequent
initiated a crack. By analysing the EBSD image of the stops which can last several thousand cycles. Between
same zone it is found that the crack initiated on the side those stops the crack growth rate at the sample’s sur-
of the pore which contains a grain boundary (almost face can be larger or smaller than that observed for
perpendicular to the stress direction). Once initiated, longer cracks with the same stress intensity factor DK.
this crack quickly propagated along the grain boundary The stops observed for short cracks can be correlated
(initiation and propagation within 10 000 cycles=3% with the presence of grain boundaries. This is illustrated
in Fig. 7 which shows an optical micrograph (Fig. 7a)
of the life time); then it stopped until the end of the test
of a crack observed at the surface of a sample of alloy
when the grain boundary orientation changed and be-
C during 360 000 cycles. At least three stops were
came less favourable with respect to the applied stress.
observed for this crack corresponding to nearly 100 000
A small intragranular pore not visible on the optical cycles (Fig. 7b). The EBSD technique was used to
micrograph (called P2) located nearby as well as the analyse the same zone as the one observed in optical
other — transgranular — side of pore P1, did not microscopy and revealed (Fig. 7c) that the stops corre-
initiate any crack during the test. sponded to grain boundaries. More precisely, the whole
propagation of this crack can be described as follow:
3.4. Crack propagation first the crack propagated in an intergranular mode and
remained stopped at 3% of the fatigue life (point W on
The observed propagation mechanisms of fatigue the micrograph). Between 3 and 8% of the fatigue life,
cracks can be divided in two different regimes corre- the crack bifurcated and propagated again in the trans-
sponding to high (240 MPa) and low (200 MPa and granular mode. This bifurcation at point W did not
lower) stress levels. occur from the very tip of the crack but from a point
At high stress, the cracks propagate in the material behind it, maybe because of the formation of a sub
with no visible stop at microstructural barriers and the surface transgranular crack. Then the crack stopped at
fatigue lives can be well predicted by a Paris law as it point X which corresponded to a grain boundary. After
will be shown later in this paper. 50 000 cycles the crack propagated again in a transgran-
At low stress levels, once initiated, the cracks propa- ular mode with a slight slow down apparently trans-
gate very rapidly along eutectic zones up to a size of granular at point Y. No noticeable stop was detected
when the crack met the next grain boundary, which was
in fact a triple grain boundary junction, but immedi-
ately after, the crack was stopped again at another
grain boundary (point Z) during 25 000 cycles. At that
stage, the crack propagation was screened by the prop-
agation of another crack in the sample which eventually
led to fracture. This behaviour was very typical of the
crack propagation in the studied material and has been
observed several times. It must be pointed out also that,
in many cases, a plastic zone ahead of the blocked
crack could be observed by optical microscopy in the
adjacent grain.

3.5. Correlation between pore characteristics and


fatigue mechanisms

In order to investigate a possible relation between the


pore geometrical characteristics and the initiation or
propagation of cracks, a sample of alloy C was charac-
terised by X-ray tomography. The crack initiation and
Fig. 6. (a) Optical micrograph of a crack (arrow) at the surface of a propagation at the surface was studied by optical mi-
sample of alloy C fatigue cycled at 160 MPa. (b) EBSD image of the croscopy during 233 000 cycles of fatigue at 145 MPa.
corresponding zone (note the different scales). The crack initiated However, those cracks were not visible in tomography.
from the pore labelled P1 is located in a well oriented grain boundary
The studied sample contained 34 pores located at the
with respect to the stress (nearly 90°). No crack could be seen on the
grain boundary almost parallel to the stress direction on the other surface or just below. The majority of the studied pores
side of P1 nor near the pore P2 (not visible on the micrograph) which (85%) initiated a crack. From the 3D data available for
is transgranular. each pore, a stress intensity factor K was calculated by
J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126 121

Fig. 7. (a) Optical micrograph of a crack at the surface of a sample of alloy C after 360 000 cycles at 160 MPa; (b) EBSD image showing the same
zone as in (a); (c) Length of the crack shown in (a) as a function of the number of cycles. Several stops indicated by letters (W, X,Y and Z) have
been observed during the propagation of the crack and could be correlated with the presence of grain boundaries (different grey levels).

considering the pore as a crack and by taking into to reach a size larger than (or comparable to) that of the
account its maximum width X measured at the surface pore from which they initiate.
and its maximum depth Y perpendicular to the surface
(K proportional to |local[(XY)1/2]1/2 cf. Fig. 8b).
Regarding the crack initiation, the results shown on 4. Discussion
Fig. 8a, indicate that the probability of initiating a crack
from a pore is not obviously correlated with the stress 4.1. Crack initiation
intensity factor defined above or with a stress criterion.
As a matter of fact, some pores located near the heads The mechanisms leading to the initiation of a crack
of the specimen with an applied stress as low as 45 MPa from a pore have been identified. For the alloys investi-
have been observed to initiate a crack. Finally, it was gated here those mechanisms are the same at high and
found that pores with an equivalent size lower than 50 low stresses. Namely, the application of a cyclic stress on
mm did not initiate any crack [32]. the material induces some decohesions between the Si
Regarding the propagation of cracks, Fig. 8a shows eutectic particles and the surrounding Al– a matrix.
that the ability of a crack to propagate away from a pore Those decohesions rapidly lead to a small crack in the
is also apparently not correlated with the stress level aluminium phase. At least two factors can account for
around the pore nor with the associated stress intensity the decohesions. First, the elastic and plastic incompat-
factor. 3D images also reveal (see Fig. 9) that non ibilities existing between the hard and brittle Si particle
propagating cracks always have a size (measured at the and the ductile aluminium matrix and, second, the
surface by OM) smaller than the pore from which they location of the Si particles near the convex parts of pores
initiate. In other words, only propagating cracks manage which act as stress concentrators (see Fig. 3).
122 J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126

Moreover, as observed experimentally, the presence 4.2. Crack propagation


of a suitabily orientated grain boundary (Fig. 6) is a
microstructural factor which promotes the initiation of At low stresses, once initiated, many cracks have
a crack from a pore, according to the mechanism been observed to stop rapidly with no further propaga-
described above. Indeed, a grain boundary induces a tion. The possible mechanisms which could account for
plastic incompatibility which is superimposed to that such stops are discussed in what follows.
existing between the matrix and the Si particle. Also, No correlation could be found between the ability of
some gain in surface energy can be obtained for the a crack to propagate and the stress levels around the
nucleation of a crack in a grain boundary. It must be pore or the stress intensity factors calculated from the
remembered that all large pores are intergranular thus pore size taking into account the depth of the pore.
the probability to initiate a crack from such defects is This indicates clearly that the classical approach which
high, in agreement with the experimental observations. treats pores as pre existing cracks of similar geometrical
The fact that small pores (with an equivalent size lower dimensions was not valid in our case.
than 50 mm) did not initiate any crack could be due A pore with a very complex shape (low sphericity)
either to a higher probability of being transgranular or could result in a high (or a low) value of the local stress
to a lower stress concentration effect. intensity factor which cannot obviously be calculated

Fig. 8. (a) Classification of the pores observed by tomography at the surface of a sample of alloy C fatigue tested during 233 000 cycles. For each
pore, the effective stress level has been indicated as well as an equivalent stress intensity factor K calculated from the X transversal and Y
longitudinal dimensions of the pore. For each pore the different symbols indicate if no crack initiation (), a non propagating crack ( ) or a
propagating crack () were observed. (b) Definitions of X and Y on a 3D volume rendering of one pore.
J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126 123

Fig. 9. (a) Classification of the pores observed by tomography at the surface of a sample of alloy C fatigue tested during 233 000 cycles. For each
pore, the ratio between the lateral size Lc of the crack at the surface (including the pore) and the maximum lateral dimension of the pore below
the surface Lp obtained by tomography has been indicated. The different symbols indicate if a non propagating crack ( ×) or a propagating crack
() were observed. (b) Definitions of Lp and Lc parameters. In black: the pore intersection with the sample surface. In grey a projection of the
3D shape of the pore below the surface.

precisely using the X and Y dimensions defined in Fig. nodular cast iron, for cracks which are initiated on
8. The measure of the 3D radius of curvature at the micro-shrinkage pores.
surface of such pores could in principle give more Using similar arguments, it is believed that, in the
accurate values of stress intensity factors. However, the case of the cast alloy studied here, almost all pores
resolution of the tomography images (6.65 mm)3 was too cutting the surface are likely to initiate a crack because
low to obtain reliable data. of the high value of the local stress intensity factor at
Following the same idea, complementary Finite Ele- the intersection between the pore and the surface. This
ment calculations have been used to investigate numeri- is in agreement with our experimental observations
cally a possible effect of the pore shape on the local showing that nearly all pores initiate a crack even at
stress concentration coefficients. It was found that a very low stress levels. However, for the reasons men-
pore with a complex shape leads to a stress intensity tioned above, most of the resulting cracks tend to stop
factor which is nearly the same (10% difference) as the rapidly when a 3D equilibrium shape is reached (de-
stress intensity factor of a simple equivalent ellipse crease of the stress intensity factor at the surface). It is
believed that this shape does not result in a stress
exhibiting the same projected lateral size and the same
intensity factor high enough for propagation unless if
local radius of curvature. Thus, a shape effect is quite
some microstructural factor can favour the propagation
unlikely to explain the observed crack propagation/
as discussed in the following paragraph. Unfortunately,
stop.
it was not possible to visualise the 3D shapes of the
Similar stops have already been reported for natu-
stopped cracks directly because they were not visible on
rally occurring cracks in nodular cast iron [33,34]. In the reconstructed 3D images. Further tomography ex-
this material, the nodules can be treated as holes be- periments at ESRF using a 1 mm resolution and a larger
cause their interfaces with the matrix are weak. Cracks (2000*2000 pixels) CCD camera are currently being
initiate at the intersection of the graphite nodules with planned to investigate this point.
the surface because the stress concentration is higher in Finally, as shown above, the local crystallography
this region [33]. Very often, those cracks are observed to around the pores plays an important role on the crack
stop in the vicinity of the nodules. It has been verified propagation. For the cracks which managed to escape
experimentally by serial polishing that the shape of the ‘their’ pore a careful investigation of the local crystal-
stopped crack corresponded to a 3D equilibrium shape lography by EBSD revealed that, from the beginning,
for which the stress intensity factor is roughly constant many of them were located in a well orientated grain
along the crack path [35] and lower than the propaga- boundary which has eased their propagation. If the
tion threshold for long cracks. According to the same grain boundary changed direction of if a triple grain
authors this ‘three dimensional effect’ can also occur, in boundary junction was reached the crack might stop.
124 J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126

The presence of a grain boundary not visible at the tion clearly underestimates the experimental values
surface might also cause the crack to slow down or and the gap between experiments and predictions in-
even to stop as already reported elsewhere [36]. creases when the stress decreases. In order to assess
In any case, it must be remembered that, even at the influence of the Paris law formulation on the fa-
low stresses, the fatigue life of a sample mostly con- tigue life calculation, a different expression suggested
sisted in a propagation stage of one main crack. recently [38] has been used. The parameters of this
Hence, an estimation of the fatigue life of the mate- law have been determined on the basis of the experi-
rial has been calculated through the fatigue crack mental fatigue lives measured at 240 MPa. The re-
growth rate da/dN using a simple Paris law: sults of this calculation are shown in Fig. 10. The
same trend as with Eq. (1) is observed for the calcu-
da DK m lated and experimental fatigue lives: good agreement
=C . (1)
dN (1−R)Kf −DK at high stress levels, maximum disagreement at low
In this equation R is the load ratio, R =0.1, m and stress levels.
C are two material dependent constants m = 2.94, Various factors including crack closure could ac-
C= 2.9 ×10 − 9 Kf is the critical stress intensity factor, count for this discrepancy. However, in our case, due
Kf = 41.5 MPa m1/2, and DK is the classical stress to both the value of the stress ratio used, and,
intensity range. This law has been used previously in mainly, to our experimental observations, it is be-
the literature to describe successfully the fatigue be- lieved that the crack stops at grain boundaries is the
haviour of a cast aluminium alloy similar to the one main factor which account for the underestimation of
studied here [37] and has been verified experimentally, the fatigue lives through Paris laws. As an example, it
on average, on our measurements. was observed [32] that for fatigue cracks initiating on
The initial size a0 of the crack leading to failure small pores (DK lower than the case shown on Fig. 7)
has been measured by SEM on the fracture surface of crack stops can account for 90% of the fatigue life at
130 MPa (nominal stress).
the broken samples. It is equal to the geometrical
As a matter of fact, the influence of grain
mean of the transversal and lateral dimensions of the
boundaries on crack propagation has already been re-
initiating pore with respect to the surface. The size of
ported in the literature for steels [36,39] and alu-
the critical crack ac leading to fracture is taken equal
minium [21,23,40]. The physical processes explaining
to 2 mm. This last value has little influence on the
how a crack can cross a grain boundary are still
calculated fatigue lives. The stress intensity factor
unclear. One possible mechanism is the activation of
range DK is calculated by assuming the initial pore/
dislocation sources in the adjacent grain and the cre-
crack to be a semi circular notch of radius a cutting ation of a new micro crack by irreversible accumula-
the surface of the sample. tion of dislocations in that grain. This is in agreement
The fatigue lives calculated from Eq. (1) are shown with our experimental observations of plastic zones in
in Fig. 10. They are in good agreement with the ex- the neighbouring grain ahead of the stopped crack. A
perimental data at high stresses (|=240 MPa) where dislocation based model derived from this mechanism
the crack propagation is quite insensitive to the mi- has been proposed [32]. The number of cycles neces-
crostructure. At lower stresses, however, the calcula- sary for a crack to overcome a grain boundary can
be estimated with this model and is in good agree-
ment with experimental observations. Details of this
model will be published in a forthcoming paper.

5. Conclusion

The microstructure and the fatigue properties of


three model AS7G03 cast aluminium alloys contain-
ing artificial pores have been studied. Much attention
has been paid to the 3D characterisation of porosity
by means of synchrotron X-ray tomography and also
to the study of fatigue crack development in relation
with local crystallography by means of EBSD. The
Fig. 10. Comparison between experimental (Nexp) and calculated major findings can be summarised as follow:
(Ncalc) number of cycles to failure (2, D|= 240 MPa; , D|= 200
Two types of pores are present in the material:
MPa; , D|= 160 MPa; , D|= 130 MPa). The straight line
indicates the locus of the points where Nc = Nexp. Full and hollow microshrinkage pores and artificial gas pores. The
symbols indicate the values calculated with Eq. (1) and with the Paris amount of gas introduced in the alloys during pro-
law expression given in [38], respectively. cessing has almost no influence on the volume
J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126 125

fraction of microshrinkages but increases exponentially [2] P.C. Inguanti, Overcoming Material Boundaries, Proceedings
of the 17th SAMPE Technical Conference, SAMPE, Covina
the volume fraction of gas pores. Most pores are inter-
CA, 1985, p. 61.
granular. Their convex parts correspond to interden- [3] C.M. Sonsino, K. Dieterich, Giessereiforschung 43 (1991) 119.
dritic zones containing brittle eutectic silicon particles. [4] S. Murali, T.S. Arvind, K.S. Raman, K.S. Smurthy, Mater.
Regarding the fatigue properties, the porosity level of Trans. JIM 38 (1997) 28.
the alloys was found to be a first order parameter which [5] G.W. Powell, Mater. Charact. 33 (1994) 275.
[6] B. Skallerud, Eng. Fract. Mech. 44 (1993) 857.
greatly alters both the average number of cycles to
[7] M.E. Seniw, M.E. Fine, E.Y. Chen, M. Meshii, J. Gray, in:
failure and the lifetime scatter, specially at high stress W.O. Soboyevo, T.S. Srivatsan (Eds.), Proceedings of the Con-
level. The initiation of fatigue cracks occurs by decohe- ference on High Cycle Fatigue of Structural Materials, TMS,
sions of Si particles located near the convex parts of Warrendale, PA, 1997, p. 371.
pores which are located at or below the surface of the [8] C.C. Wigant, R.I. Stephens, Report No. 870096 Society of
Automotive Engineers, Warrendale PA, 1987.
samples. For the stress levels investigated the initiation
[9] M.J. Couper, A.E. Neeson, J.R. Griffiths, Fatigue Fract. Eng.
stage is always a negligible part of the whole fatigue life. Mater. Struct. 13 (1990) 213.
Although nearly all pores give rise to a crack, most of [10] B. Skallerud, Fatigue Eng. Mater. Struct. 7 (1988) 183.
those cracks remain stopped near the pores. The crack [11] C.C. Wigant, R.I. Stephens, in: R.O. Ritchie, E.A. Starke
propagation at high stress level (|= 240 MPa) is quite (Eds.), Proceedings of Fatigue 87, Engineering Materials Advi-
sory Services, Warley, 1987, p. 49.
insensitive to microstructural barriers and can be reason-
[12] G.A. Hoskin, J.W. Provan, J.E. Gruzleski, Theor. Appl. Fract.
ably well described by a Paris type law. At low stresses, Mech. 10 (1988) 27.
however, cracks are often observed to stop at grain [13] C. Verdu, H. Cercueil, S. Communal, P. Sainfort, R.
boundaries and the fatigue life is no longer predicted by Fougères, Mater. Sci. Forum 217 (1996) 1449.
a simple propagation law. [14] A. Plumtree, S. Schafer, in: K.J. Miller, E.R. de los Rios
(Eds.), The Behaviour of Short Fatigue Cracks, Mechanical
Thus, the modelling of fatigue behaviour at low stress
Engineering Publications, London, 1986, p. 215.
requires a thorough understanding of the mechanisms [15] P.N. Crepeau, S.D. Antolovitch, J.A. Worden, in: H.A. Ernst,
which allow cracks to cross grain boundaries. Some A. Saxena, D.L. McDowell (Eds.), Proceedings of the 22nd
interesting experimental information on those mecha- Symposium on Fracture Mechanics, American Society for
nisms can be obtained by precise EBSD [21] but those Testing and Materials, Philadelphia PA, 1992, p. 707.
[16] W.J. Evans, H.V. Jones, J.A. Spittle, S.G.R. Brown, in: H.
experiments can only be performed at the sample sur- Oikawa (Ed.), Proceedings of the 10th Conference on the
face. A significant step ahead could be made if one were Strength of Materials, The Japan Institute of Metals, Sendai,
able to observe the progressive crack path in the bulk of 1994, p. 501.
the material. This work is currently being carried out by [17] Y. Tan, S. Lee, H. Wu, Int. J. Fatigue 18 (1985) 137.
means of synchrotron X-ray tomography. Preliminary [18] A. Wickberg, G. Gustafsson, L.E. Larsson, Report No.
840121 Society of Automotive Engineers, Warrendale, PA,
results have already shown that such images could be 1984.
obtained with a 1 mm resolution. This should provide an [19] J.A. Ødegard, H.J. Roven, K. Pedersen, in: M. Jono, T. Inoue
invaluable information for the study of crack propaga- (Eds.), Proceedings of the 6th International Conference on the
tion and for the modelling of fatigue properties. Mechanical Behaviour of Materials, Pergamon Press, Oxford,
1992, p. 439.
[20] K.J. Miller, E.R. de los Rios, The Behaviour of Short Fatigue
Cracks, Mechanical Engineering Publications, London, 1986,
Acknowledgements p. 191.
[21] T. Zhai, A.J. Wilkinson, J.W. Martin, Mater. Sci. Forum 331
The authors would like to thank J.C. Ehrström and F. (2000) 1549.
Cossé from the Pechiney Research Centre (Voreppe [22] A. Turnbull, E.R. De Los Rios, Fatigue Fract. Eng. Mater.
Struct. 18 (1995) 1355.
France) for the financial support during this study, for [23] J. Lankford, Fatigue Eng. Mater. Struct. 8 (1985) 161.
providing the material and also for stimulating discus- [24] J-Y. Buffière, E. Maire, P. Cloetens, G. Lormand, R.
sions. Thanks are also due to the ID 19 staff at ESRF Fougères, Acta Mater. 47 (1999) 1613.
and G. Peix from CNDRI Laboratories for their help [25] S. Savelli, J.Y. Buffière, R. Fougères, Mater. Sci. Forum 331
during the tomography experiments and the fruitful (2000) 197.
[26] J. Baruchel, J.Y. Buffière, E. Maire, P. Merle, G. Peix, X-Ray
discussions. Finally, the authors are grateful to Professor Tomography in Materials Science, Hermes Science, Paris,
P. Buffat, from CIME laboratory at EPFL Lausanne, 2000, p. 45.
for using the Philips XL 30 FEG microscope and to M. [27] R.A. Schwarzer, Micron 28 (1997) 249.
Jourlin for his help in analysing the 3D images. [28] B.L. Adams, Ultramicroscopy 67 (1997) 11.
[29] D.P. Field, Ultramicroscopy 67 (1997) 1.
[30] F. Cléton, P.H. Jouneau, S. Henry, M. Gäumann, P. Buffat,
Scanning 21 (1999) 232.
References [31] Q.G. Wang, D. Apelian, Report No. 97A-PR991,Worcester
Polytechnic Institute, Metals Processing Institute, Aluminium
[1] J.A. Ødegard, K. Pedersen, Report No. 940811, Society of Casting Research Laboratory, MA, 1998.
Automotive Engineers, Warrendale, PA, 1984. [32] S.Savelli, Doctorate thesis, INSA Lyon, 2000.
126 J.-Y. Buffière et al. / Materials Science and Engineering A316 (2001) 115–126

[33] A. Pineau, in: R.O. Ritchie, J. Lankford (Eds.), Proceedings of [37] L. Schwarmann, Material Data of High Strength Aluminium
the 2nd International Conference on Small Fatigue Cracks, The Alloys for Durability Evaluation of Structures, Aluminium Ver-
Metallurgical Society, Warrendale, PA, 1986, p. 191. lag, Düsseldorf, 1985, p. 110.
[34] J.P. Monchoux, Doctorate thesis, INSA Lyon, 2000. [38] A.J. McEvily, H. Bao, S. Ishihara, in: X.R. Wu, Z.G. Wang
[35] P. Clément, J.P. Angeli, A. Pineau, Fatigue Eng. Mater. Struct. (Eds.), Proceedings of Fatigue 99, Engineering Materials Advi-
7 (1984) 251. sory Services, Warley, 1999, p. 329.
[36] K. Tokaji, T. Ogawa, Y. Harada, Z. Ando, Fatigue Fract. Eng. [39] T. Magnin, J. Stolarz, J. Phys. IV (8) (1998) 105.
Mater. Struct. 9 (1986) 1. [40] J. Lankford, Fatigue Eng. Mater. Struct. 5 (1982) 233.

You might also like