Sie sind auf Seite 1von 24

Accepted Manuscript

Comparative analysis of the fatigue short crack growth on Al 6061-T6 alloy by


the exponential crack growth equation and a proposed empirical model

Jan Mayén, Arturo Abúndez, Isa Pereyra, Jorge Colín, Andres Blanco, S. Serna

PII: S0013-7944(16)30354-X
DOI: http://dx.doi.org/10.1016/j.engfracmech.2017.03.036
Reference: EFM 5461

To appear in: Engineering Fracture Mechanics

Received Date: 2 September 2016


Revised Date: 25 March 2017
Accepted Date: 26 March 2017

Please cite this article as: Mayén, J., Abúndez, A., Pereyra, I., Colín, J., Blanco, A., Serna, S., Comparative analysis
of the fatigue short crack growth on Al 6061-T6 alloy by the exponential crack growth equation and a proposed
empirical model, Engineering Fracture Mechanics (2017), doi: http://dx.doi.org/10.1016/j.engfracmech.
2017.03.036

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Comparative analysis of the fatigue short crack growth on Al 6061-T6 alloy by the exponential
crack growth equation and a proposed empirical model

1
Jan Mayén*, 1Arturo Abúndez, 2Isa Pereyra, 1Jorge Colín, 1Andres Blanco, 2S.Serna.
1
Tecnológico Nacional de México-Centro Nacional de Investigación y Desarrollo Tecnológico
(CENIDET),
Prolongación Palmira s/n esq Apatzingan, Col. Palmira. Cuernavaca, Morelos, México
2
CIICAp-FCQeI-Universidad Autónoma del Estado de Morelos Av. Universidad 1001 Col. Chamilpa
C.P. 62209-Cuernavaca, Mor. México.
*E-mail: jmayen@cenidet.edu.mx, Phone/fax: Tel./fax: +52 777 362-7770 ext. 1408

Abstract
In this work, the short crack propagation and growth rate nonlinear behavior of an Al 6061 alloy in
T6 condition, were assessed. The specimens were fatigue tested at 20 Hz in a rotating bending fatigue
machine with constant amplitude loading of 88%, 61% and 38% of yield strength, and the short crack
initiation and propagation were followed up by optical microscopy. Interaction between the surface
cracks and aluminum microstructure were identified: persistent slip marks (PSMs) can be observed
on the surface specimen, where persistent slip bands (PSBs) emerge from the surface as a result of
accumulated damage by intrusion-extrusion mechanism, which eventually leads to the accumulation
of dislocations within the grain as a consequence of grain border barriers and stress concentration at
second phase particles. The recorded length crack paths were later analyzed to study the crack growth
by Frost and Dugdale exponential crack growth equation (ECG) and its comparison with a proposed
model (PM) developed by multiple linear regression. The results showed that the crack growth rate is
affected by microstructural features in the early growth stage, which is assumed to be from the
nucleation up to 150 µm of crack length. The ECG is reliable only for the stage where the crack
growth is mainly a consequence of stress concentration at crack tip; in contrast, the PM do properly
describes the nonlinearity behavior between short and long crack growth.

Keywords: Fatigue crack propagation; aluminum alloy; exponential crack growth equation; multiple
linear regression.
1. Introduction
The aluminium alloys are the preferred material choice for numerous structural applications,
specifically for the aerospace and automotive industries, since they exhibit benefits such as the ease
of manufacture, corrosion resistance, weldability, along with lightweight over those of structural
steels. The Al 6061 is well known as heat treatable alloy that exhibits high strength–weight ratio
coupled with excellent corrosion resistance. Its extrudability is a unique feature that make it suitable
for a broad usage in welded structural parts for vessels, aircrafts, and automotive body panels and
pipelines. All these structural parts are subjected to high cycle loading which causes fatigue. The
fatigue damage of materials has been studied since the discovery of the fatigue occurrence. Ewing et
al. [1] studied the Swedish iron during rotating bending fatigue surface and focused on the surface
damage. By the Usage of optical microscope, they found and studied the formation of surface
markings that occurred during cyclic loading, where fatigue cracks developed later in fatigue life.
Today, problems on crack propagation in aluminium and its alloys, are of great interest to the
scientific community, where the vast accumulated experimental results and theoretical methods have
been used to acquire a deep knowledge of crack propagation over the past few years. However,
additional experimental and theoretical work is required in order to achieve a deeper understanding
related to crack initiation (known as stage I) of fatigue on structural parts, which is a crucial
constraint for structural design. Some researchers as Zhuang et al. [2] and Jones et al. [3], have
commented on the lesser effect of the final crack size on fatigue crack life, mainly because the crack
growth rate is much faster during the later stage of growth, where the crack growth rate is relatively
slow in the early stage. This makes the total life strongly dependent on the initial crack size [2].

The fatigue crack propagation has three well-known stages: the stage I occurs when the crack has
been nucleated and initiates its propagation, which is related to several microstructural features. Stage
II, corresponds to an increase of the stress intensity factor as a consequence of an increase of the
crack length, along with the activation of deformation systems. Stage III is related to a critical crack
length and an unsteady behavior [4].

There are several studies reported in literature [5], [6] and [7], which are related to the crack
propagation in stage I. For the aluminium alloys, the crack propagation in Stage I acquired more
importance when two Comet jet airplanes crashed in 1954. These airplanes, were manufactured with
7075 aluminium components; the examination on the airplane remains showed that the crash was
related to a fatigue failure initiated in the fuselage, along with a stress concentration at the windows
and hatches [8]. However, the key research on fatigue crack propagation in the past decades have
focused on Linear Elastic Fracture Mechanics approach (LEFM) rather than short crack or small
crack issues [4], where it was originally thought that, for any given material and thickness, da/dN
versus ΔK relationship was unique. Pearson [9], conducted a research related to the behavior of short
cracks in commercial aluminium alloys, and he observed an atypical propagation phenomenon: short
cracks propagate below the ΔKth threshold. Also, Meggiolaro et al. [10] and Nalla et al. [11] stated
that short cracks propagate below the effective stress intensity threshold ΔKth,eff. It has been verified
that the crack nucleation and crack propagation in Al-Si cast alloys occur through the α-aluminum
matrix, and is due to the stress intensity increase at the crack tip; when the crack is long enough, the
stress intensity factor at the crack tip becomes the mechanical driving force for the crack propagation
[12].

Some techniques, which are focused on the crack propagation characterization have been proposed
over the years, where the most common technique used is the Linear Elastic Fracture Mechanics
(LEFM). This technique has been effectively used to calculate the crack propagation from a flaw,
defect or initial crack size a, related to fatigue cycles N. The crack propagation behavior usually
follows the Paris equation [4]; therefore, the Paris equation continues to be the accepted standardized
method worldwide to investigate the fatigue crack growth. However, several scientists [13], [14] and
[15], have found that short fatigue cracks showed irregular behavior, which has been characterized as
atypical when compared to longer cracks under nominally elastic cycling, for a diversity of
commercial alloys employed in structural parts [16], whose behavior cannot be described by the
stress intensity factor threshold (Kth) from the LEFM [17], suggesting that the crack growth and
evaluation of failure requires that the crack tip stress field has to be evaluated at a length scale r0 in
front of the crack. The characteristic length scale r0 has been widely used in the assessment of fatigue
crack growth by the fracture mechanics [17], [18], [19], [20], [21]. Therefore, the evaluation of long
cracks propagated under Mode I uniaxial loading, where the ratio r0/a have a tendency to be quite
small and, as a consequence, the error in the usage of ΔK and Kmax to characterize the crack tip stress
field is also quite small. However, for short cracks with length scales of the order of 10 µm, the ratio
r0/a will not be small [22], consequently increasing the error in the usage of ΔK and Kmax.
Subsequently, it has been confirmed by several authors [22] and [23] that the conclusions reached by
[17], were correctly assessed. Therefore, it can be concluded that K dominance is lost for short
cracks. Consequently, the similitude hypothesis is also invalid for short cracks. In this context, it was
clear that the generalized Frost and Dugdale [24] model was a potential alternative. They studied the
crack growth of fatigue cracks on huge thin panels of steel and light alloys, including aluminium
plates with a central notch as crack starters; they related the fatigue crack growth rate to the crack
length and to the third power of the stress range in a proportional way. Later, Ritchie [25] studied the
nucleation and growth of fatigue cracks on aluminium and found that the propagation behavior was
crack length dependent, and that the crack growth was dependent to native crystallographic
characteristics as the direction of neighboring grains and grain border structure. This was also
demonstrated by Molent et al. [26], Zhai et al. [27]and Polák et al. [28].

In this work, the crack length a, and fatigue cycles N values obtained by rotating bending fatigue
testing of an aluminium 6061-T6 temper using a slightly modified notched specimen, were assessed.
The crack length and fatigue cycles values were the input parameters for the exponential crack
growth equation (ECG) proposed by Frost and Dugdale, in order to characterize the short crack
propagation behavior. In addition, a proposed robust multiple linear regression (PM) on crack length
and fatigue cycles values was done in order to describe the correlation among stress amplitude, crack
length and number of cycles during the fatigue tests. The results obtained by the PM showed an
initial extrapolated crack ai value similar to those values obtained by the ECG. Furthermore, the PM
allows to directly estimate the crack growth rate as a function of the stress amplitude for aluminium
alloys. Finally, a microstructure relation to the crack growth initiation and early propagation of the
cracks was studied to identify the possible microstructural features affecting the crack growth
behavior.

2. Experimental Procedure
The specimens were obtained from an aluminium 6061-T6 alloy plate with 12.5 mm thickness and
machined to obtain longitudinal specimens along the extruded axis direction (T-L), as depicted in
Fig. 1. The reported commercial chemical composition in wt% is: Al-Balance, Cr-0.1%max, Cu-
0.1%max, Fe-0.35%max, Mg-0.45-0.9%, Mn-0.1%max, Si-0.2-0.6%, Zn-0.1%max, Ti-0.1%max.
The mechanical properties, experimentally obtained, were: 249 MPa for yield strength (YS), 69 GPa
for Young’s modulus (E), 281 MPa for ultimate tensile strength (UTS), and a 13% for ultimate strain
(ε).

Fig. 1. Extruded Axis Direction.

Jan Mayén et al. [29], following the research reported by Narasaiah and Ray [30], developed a new
subsized specimen geometry, where regions “A” and “B” were modified in order to induce a crack
initiation, as depicted in Fig. 2. The specimens used for crack growth tests were prepared as
described elsewhere. [29]. The fatigue crack growth experiments were performed using a rotating
bending fatigue machine by Fatigue Dynamics, Model RBF 200.
Fig. 2. Specimen proposed geometry (dimensions in mm).

The loading conditions for fatigue testing were: 242 MPa, 169 MPa and 106 MPa of stress amplitude,
corresponding to 88%, 61% and 38% of yield strength of the aluminium alloy, respectively. These
loads were obtained according to the Finite Element Analysis reported elsewhere [29], where the
higher stress values in the specimen subjected to bending loads are located at regions A and B,
including the notch effect. The experimental conditions for the crack growth by fatigue were: a) 20
Hz at room temperature, b) an initial fatigue test was performed in order to induce a 15-50 µm short
crack at regions A and B, c) the crack paths were monitored by an optical microscope along the test,
d) the tests were arrested once crack length was approximately 1x10 3 µm long. The recorded crack
paths were posteriorly analyzed in order to obtain the short crack length, which were fitted by Eq. (1)
[24].

(1)

Where, N is the number of accumulated cycles, kg is a coefficient that characterizes the crack growth
rate, which is directly related to the geometry specimen, material and load [24]; a is the crack length
and ai is the initial crack length related to a materials defect, a second phase particle or other
microstructural features.

In order to observe the PSB’s density along the crack path, the specimens were analyzed by optical
microscope. One specimen, for each of the loading conditions was etched by immersion in Kellers
etch (190 ml of distilled water, 5 ml of nitric acid, 3 ml of hydrochloric acid and 2 ml of hydrofluoric
acid) for 10 seconds lapse of time, rinsed in distilled water followed by immersion in acetone and
rapidly dried by hot air, in order to allow a slight revealed of the microstructure to observe the PSB’s
by scanning electron microscope (SEM JSM 5900-LV) at 20 kV and 6 nA for the beam current.
2.1 Multiple linear regression methodology for crack length, stress amplitude and fatigue cycles
number correlation.

The main aim of this work, was the development of an empirical model that correlates the stress
amplitude during fatigue cyclic loading with the crack length. The empirical model was developed by
Multiple Linear Regression (MLR) technique, since there are more than two independent variables
(predictors) in the experimental conditions. The model, using population information, is described by
Eq. 2 [31],

(2)

This method requires the usage of some assumptions. These are necessary for the mathematical
process, in order to properly find the optimal empirical equation that better fits the experimental data.
The predictors x1=N1, x2= σ2, …xkt represents the control parameters for the experimental conditions,
and are: a) number of fatigue cycles N, and b) stress amplitude σ. These assumptions are:
1. The error terms εi are normally distributed and equal to 0.
2. The error terms are independent of past error terms, which is

(3)

3. All the population have equal variances, S1= S2=S3=···


4. The predictors are not correlated with each other.
The Eq. (2) can be represented as the matrix form, where:

(4)

Then,

(5)

The least squares method is then applied to Eq. (5) in order to find the estimates of β that minimizes
the sum of squares of the residual,

(6)

Since ε=Y-Xβ. Expanding S(β) it is then obtained,


(7)

Subsequently,

(8)

As it can be done in calculus to find the minimum of a function, the least squares estimate β can be
obtained by solving the Eq. (8) based on the first derivative of S(β),

(9)

Then,

(10)

And thus β is obtained,

(11)

3. Experimental results and discussion

3.1 Relation between short crack propagation and microstructural features.


In the Fig. 3, the final crack path on the specimen surface, which is located in the notch at region A of
the specimen according to FEA results [29], which corresponds to the loading conditions at 88%,
61% and 38% of yield strength of the aluminium alloy.

a) 106 MPa.
b) 168 MPa

c) 242 MPa.
Fig. 3 Crack initiation for the three different loading conditions.

The predominant crack growth mechanism observed at a detail by SEM observations (see Fig. 5) was
crack propagation due to the intrusion-extrusion mechanism, which is described in Fig. 4.

Fig. 4 Intrusion – extrusion mechanism [32], [33].

This intrusion-extrusion mechanism, has been considered as the earliest evidence of metal fatigue,
which is related to the development of surface deformation in the form of PSB’s that emerge inside
the grains of the metal matrix [34]. These microstructural features are visible by their profile, as can
be seen in the Fig. 5 (a). These PSB’s are the evidence of a severe plastic deformation localized
within the grain that leads to the formation of the fatigue crack and its propagation.
The SEM micrographs depicted in Fig. 5, show evidence of PSB’s (a), as the main crack growth
mechanism. In Fig. 5 I, it can be observed that the crack is propagating from the PSB’s (a). Also,
decohesion of second phase particles (b) can be observed (see Fig. II). Another important observation
is that second phase particles act as crack starters (c) (see Fig. 5 II) or propagation barriers (d) (see
Fig. 5 I and II).

I)106 MPa II) 168 MPa

III) 242 MPa


Fig. 5 Microstructural features at the crack tip for the corresponding experimental loading conditions.

The aluminium 6061 alloy, contains intermetallic compounds of a very complex structure; these
compounds are nucleated by the casting process and may act as crack starters or crack propagation
barriers, depending on its chemical composition and morphology. Furthermore, the intermetallic
compounds reported for this aluminum alloy are (Fe, Mn, Cu) 3SiAl12 and Mg2Si, being the latest the
main phase reported [35]. In addition, the intrusion-extrusion mechanism leads to the accumulation
of dislocations within the grain, which is due to grain border barrier; this mechanism allows the crack
to propagate within the material matrix.

3.2 Determination of fatigue short crack growth by Frost and Dugdale exponential crack
growth equation (ECG).
The short crack length was inspected during the fatigue tests, which was carried out in batches, by
acquiring images by optical microscopy from notched area, at each test stop. Once the tests were
over, principal crack lengths were assessed as a function of loading cycles for the three loading
conditions mentioned above. The short crack path usually follows the slip plane or grain boundaries
(see Fig. 5), which corresponds to the crystallographic orientation with the highest Schmidt factor
[32].

The crack growth is depicted in Fig. 6 for the three cyclic loading conditions. As it can be observed,
the crack for the specimens subjected to 242 MPa nucleated at a lower cycling than those specimens
subjected to 168 MPa and 106 MPa, this is due to high localized plastic deformation at the notched
region. As can be observed in Fig. 6, the slope of the crack growth curves represent the crack growth
rate, which is related to the loading test condition, which implies that propagation rate is directly
proportional to the stress amplitude applied. At the early stage of fatigue tests, for 242 MPa and 106
MPa a crack arrest can be observed (see enclosed points); in consequence, this instability of the crack
growth has to be taken into account when estimating the crack growth rate.

Fig. 6. Crack growth for three different loading conditions.

In Fig. 6, the crack length as a function of fatigue cycles for the three experimental loading
conditions is depicted. As can be observed, the cracks of the specimens subjected to 242 MPa and
106 MPa were nucleated earlier than those of the specimens subjected to 169 MPa, this maybe is due
to the microstructural features acting as propagation barriers in the notched zone for the 106 MPa
condition. The crack growth curve of the specimens subjected to 242 MPa depicts a linear behavior,
which is related to a faster crack propagation due to a lack of interaction between crack and
microstructural features due to the high localized plastic deformation at the crack tip. As can be seen
in Fig. 6, the crack growth for the specimen subjected to 106 MPa is arrested from 103 up to 105
cycles, around to 33 µm long; after that, the crack grows at a constant rate during the test remaining.
This behavior can be related to the interaction between the crack path and the microstructural
features, as stated by Polák et al. [28].

The Eq. (1) [24] relates the crack length a and the fatigue cycles N during fatigue tests, and indicates

that the crack growth rate directly depends on the crack length a, which represents the cumulative

crack length during the experimental fatigue test. The equivalent crack growth for the three
experimental loading conditions were obtained according with Eq. (1), which can be rewritten as:

(12)
And in the matrix form:

(13)

Where ai is the initial crack length or the initial flaw size, and kg characterizes the crack growth rate
at the applied stress amplitude, n is the amount of experimental data. The input information of the Eq.
(13) are obtained from the experimental data depicted in Fig. 6 and, when solved, the values of kg and
ai are obtained and, therefore the exponential equation that describes the behavior of crack growth is
obtained. The Table 1 contains the values of the crack length obtained by solving Eq. (13), as a
function of the testing loading conditions and the number of fatigue cycles.

Table 1. Crack growth as a function of loading condition.


loading Crack length (µm) R-square
242 MPa Test 1 0.98
242 MPa Test 2 0.97
169 MPa Test 1 0.89
169 MPa Test 2 0.93
106 MPa Test 1 0.89
106 MPa Test 1 0.87

The extrapolated crack values for the initial crack length at 0 cycles is in the range between 82 µm
and 34 µm for 242 MPa and 106 MPa, respectively. This extrapolated crack is assumed as the initial
defect that propagates the crack during the fatigue testing.

For the three tested conditions, an accelerated growth can be observed up to 100 µm, and the crack
growth rate changes approximately between 100 µm and 200 µm, this is due to the influence of the
microstructure at this early stage (indicated by the lines in Fig. 7, where two slopes can be
approximated). The influence of microstructural features decreases inversely to crack length, and
stress concentration at the crack tip, becomes the main mechanical driving force for crack growing,
as depicted in Fig. 7. The crack growth becomes steady above 150 µm length; this behavior may be
explained by plastic strain field localized in front of crack tip.

Fig. 7. Unsteady behavior below the 150 µm length.

Following Eq. (14) [24], an analysis of the crack growth rate was performed for the three different
loading conditions. The results of the analysis are depicted in Fig. 8, where is showed that for 242
MPa loading condition, the crack growth rate is higher than the two other loading conditions.

(14)

Fig. 8. Crack growth rate exponential equation approach.

The identification of crack nucleation and short crack growth depends strongly on detection method
employed, whether it is optical or scanning electron microscopy. With the first, the detection can be
performed about 20% of fatigue life, while with the second one, about 5% of fatigue life [36].
However, for low cost and simplicity testing, optical microscopy monitoring method by [30] for
crack growth was used. The size of the initial crack depends on the microstructural dimensions of the
specimen. Thus, if crack is nucleated by the development of PSB’s, the initial crack will be the same
size as its containing grain; in contrast, if neighboring grains have same crystallographic orientation
among them, it is possible that the cracks nucleate [37] throughout more than one neighboring grains;
crack nucleation is also possible for second phase particles, as can be seen in Fig. 5 (c). Additionally,
the linkage between the principal and secondary cracks as a crack growth mechanism, has been
observed and this behavior contributes to an accelerated crack growth, as observed for 242 MPa
loading condition. It is well known that the crack growth rate decreases as the crack tip crosses grain
boundaries, including matrix/precipitate interfaces [37].

3.3 Proposed model (PM) for the estimation of the short crack growth.
A similar short crack growth behavior analysis as described in the section above was studied, for the
Al 6061-T6 alloy cycled with constant stress amplitude, using the multiple linear regression approach
as described in section 2.1. A correlation was established between the measured crack lengths a and
their corresponding given cycles N, as a function of the stress amplitude σa during the tests. The
empirical model obtained, which describes the behavior of the crack growth as a function of stress
amplitude and number of cycles, is given in Eq. (15):

(15)

Where N is the number of fatigue cycles, σa is the stress amplitude during the test and aσaN is the
crack length as a function of stress amplitude, and P1 and P2 are defined by.

(16)
(17)

The corresponding coefficients βi are shown in Table 2.

Table 2. Coefficient βi values obtained for the PM.


Coefficients Values
-56.5927
4.7347x10-4
3.8157x10-11
0.7897
-4.3638x10-6
-2.2725x10-3
-5.0491x10-7
4.7080x10-13
-1.1745x10-2
1.2736x10-9
3.1364x10-4
0.98

It is of great importance to mention that the PM takes the stress amplitude conditions into account,
which indicates that the interpolation in the tested range is possible for crack length prediction at
other stress amplitude levels, in contrast to the ECG, where the stress amplitude variable is implicit in
the equation, therefore an equation is needed for each condition to correlate the Kg as a function of
stress amplitude, as can be seen in Table 1.

The ECG equation proposed by Frost and Dugdale [24] in the late 60’s, and employed more recently
by some researchers as Polák et al. [28] and [37], characterizes accurately the long crack growth
behavior for a given stress amplitude; however, the relation of crack length with different stress
amplitude ranges cannot be described by one single equation, and the ECG shows R2=< 0.9 for some
crack propagation data, which is due to high data dispersion during the early crack propagation,
a<150 µm, depicted in Fig. 8.

The standardized residuals values obtained from the multiple linear regression are plotted in Fig. 9.
The distribution of the residuals showed in the plot confirms that: a) the data values are
symmetrically distributed, tending to cluster towards the middle of the plot (horizontal line), b) the
data values are clustered around the lower single digits of the y-axis and, c) the data values plotted
exhibits an unpattern behavior, which indicates that the Eq. (15) describes accurately the crack
growth behavior.

Fig. 9. Standardized residuals of “a”.

In Fig. 10, the estimated values of crack length versus the experimental values are plotted, where
observable behavior between the experimental and estimated crack length values are linearly
correlated.
Fig. 10. Linear correlation between experimental and predicted crack length data with a R2=0.98.

The length of the cracks for the experimental loading conditions can be approximately estimated by
Eq. (15) for a valid interval from zero to Nf cycles, where is expected that an hypothetical crack
grows from ai up to af. Then, the period where crack initiation occurred is substituted by the length ai
of an imaginary crack at zero cycles and zero stress amplitude, which has a hypothetical length of 0
µm for the loading conditions according to Eq. (15).

The confidence intervals, obtained by the multiple linear regression methodology from the crack
length of the specimens subjected to rotating bending fatigue at 242 MPa, 169 MPa and 106 MPa, are
depicted in Fig. 11, Fig. 12, and Fig. 13, respectively. This estimated range of values, which is likely
to include the unknown population for the crack length was calculated for a 95% confidence interval.

Fig. 11. 106 MPa 95% Confidence Interval. Fig. 12. 169 MPa 95% confidence interval.
Fig. 13. 242 MPa 95% confidence interval.

In the figures above, it is observable that the proposed model takes into account the high dispersion
data below 150 µm, due to the strong interactions between the fatigue crack path and the
microstructural features. Additionally, the 106 MPa stress amplitude condition exhibits the higher
data dispersion due to the low stress concentration in regions A and B in the fatigue specimen, which
causes a chaotic behavior due to a slower crack growth that increases the interaction between the
crack and the microstructural features of the Al 6061 alloy.

3.4 Comparison between the ECG and the PM for short crack growth estimation.

The PM obtained by multiple linear regression has been compared with the ECG [24]; the results for
the Kg coefficient for the Al 6061 alloy obtained by the ECG are in good agreement with
contemporary researchers as Polák et al. [28]. In the Fig. 14, a comparison is made among the
experimental data (ED), the exponential equation (ECG) and the proposed empirical model (PM); the
curves depicted in the Fig. 14 shows that the ECG equations describe the general behavior for the
short crack, however it fails to describe the early stage of the crack, below 150 µm in length, where
the highly data dispersion is due to the strong interactions between the crack path and the
microstructural features. In contrast, the PM accurately describes the short crack growth and its
transition to long crack even for the early stage of the crack, below to 150 µm length, which allows to
estimate the fatigue life Nf.
Fig. 14. Model comparison for the fatigue crack length.

In Table 3, the R2 for the ECG, and the PM are shown. It is clear that the correlation coefficient is
directly proportional to the stress amplitude, this indicates that for lower stress amplitude values, the
data dispersion increases, and the ECG is unable to predict the behavior accurately, while the PM has
the same R2 for the stress amplitude tested ranges.

Table 3. Correlation coefficients for the models comparison.


Model Stress Amplitude R2
ECG 106 0.88
ECG 169 0.91
ECG 242 0.97
PM Any stress amplitude 0.98

In the Fig. 15, the crack growth rate is depicted for the three experimental loading conditions, where
it is observable that the ECG generally describes the crack growth rate by the Kg coefficient, as
described by Eq. (14), with a straight line for each stress amplitude during testing. In contrast, the PM
involves the interaction between the crack and the microstructural features and takes into account the
stress amplitude conditions during the fatigue testing. Hence, the PM can be used for crack length
and crack growth rate prediction by interpolation in the tested stress amplitude range for the studied
alloy, which is the range between the yield stress and the apparent fatigue limit of the Al 6061 alloy.
2x10-1
106 MPa ED
1x10-1 106 MPa ECG
106 MPa PM
5x10-2 169 MPa ED
169 MPa ECG

da/dN (m/cycle)
2x10-2 169 MPa PM
242 MPa ED
1x10-2 242 MPa ECG
242 MPa PM
5x10-3

2x10-3
1x10-3
5x10-4

2x10-4
1x10-4
5x10-5
10 20 30 40 50 60 7080 100 200 300 400 500 700 1000
Crack length "a" (m)
Fig. 15. Crack growth rate models comparison.

The correlation between the experimental data and the estimated values by ECG proposed by Frost
and Dugdale, which has been used by several contemporary authors, and the PM is compared by the
R2. The results of the correlation are plotted In Fig. 16 and Fig. 17 for the ECG and the PM,
respectively. The ECG shows an R2=0.94 and the PM a R2=0.98, these results shows that the PM has
competitive advantages against the ECG, as a more accurate estimation of the crack growth in early
stages and, therefore, estimate the remaining life of the material.

Fig. 16. Experimental crack length vs. Fig. 17. Experimental crack length vs

predicted values by ECG. predicted values by PM.

4. Conclusions

The modification of the specimen geometry allows the observation of the short crack growth by both
optical and electron microscopy, since the crack grows over a flat surface, during the rotating
bending fatigue test. Therefore, specimen geometry employed in this work has a lot of potential
benefits for the study of initiation and propagation of short cracks, i.e. low cost test or simplified
crack growth monitoring.

Experimental data obtained from fatigue testing showed that crack nucleation takes the highest
fraction of the fatigue lifetime, where this period took about 6x105 and 3x105 cycles for 106 MPa and
164 MPa loading condition, respectively. It is assumed that the crack nucleation stage involves a
short crack length up to 103 µm, which means that the highest fraction of the fatigue lifetime follows
the behavior described by the PM, instead of the fracture mechanics approach.

The responsible mechanisms for crack nucleation and propagation found in this work were: grain
boundaries as crack barriers, second phase particles as crack nucleation sites and barriers, and the
PSB’s over the specimen surface as the primary mechanism for crack growth due to extrusion-
intrusion process.

The crack growth a was fitted by the ECG, and a crack growth coefficient kg for each experimental
loading condition was obtained from the experimental data, which represents the crack growth
resistance and is inversely proportional to the crack growth rate. Also, the crack growth a was fitted
by multiple linear regression and an empirical equation was proposed, which showed good
correlation between crack growth rate and crack length at any given loading condition and cycling.
The PM describes the behavior for the stress amplitude conditions tested, which indicates the
possibility to interpolate between the tested range of stress amplitude, in order to predict the crack
length and the crack growth rate at any other stress amplitudes. On the other hand, the Frost and
Dugdale ECG describes the whole crack growth by a single slope Kg; however, this equation has not
the sensitivity required to properly describe the slope change on the transition between short and long
crack growth.

Acknowledgements

Funding supported by CONACyT through agreement Number 291113-CENIDET


We acknowledge to Dr. Iván Puente-Lee for the assistance on the scanning electron microscopy.

References

[1] Ewing J a., Humfrey JCW. The Fracture of Metals under Repeated Alternations of Stress.
Philos Trans R Soc A 1903;200:241–50. doi:10.1098/rsta.1903.0006.
[2] Zhuang W, Barter S, Molent L. Flight-by-flight fatigue crack growth life assessment. Int J
Fatigue 2007;29:1647–57. doi:10.1016/j.ijfatigue.2007.01.029.
[3] Huang P, Peng D, Jones R. The USAF characteristic K approach for cracks growing from
small material discontinuities under combat aircraft and civil aircraft load spectra. Eng Fail
Anal 2017. doi:10.1016/j.engfailanal.2017.03.008.
[4] Paris PC, Erdogan F. A Critical Analysis of Crack Propagation Laws. J Basic Eng
1963;85:528. doi:10.1115/1.3656900.
[5] Künkler B, Düber O, Köster P, Krupp U, Fritzen CP, Christ HJ. Modelling of short crack
propagation - Transition from stage I to stage II. Eng Fract Mech 2008;75:715–25.
doi:10.1016/j.engfracmech.2007.02.018.
[6] Provan JW, Zhai ZH. Fatigue crack initiation and stage-I propagation in polycrystalline
materials. II: Modelling. Int J Fatigue 1991;13:110–6. doi:10.1016/0142-1123(91)90002-G.
[7] Kim WH, Laird C. Crack nucleation and stage I propagation in high strain fatigue-I.
Microscopic and interferometric observations. Acta Metall 1978;26:777–87.
doi:10.1016/0001-6160(78)90028-7.
[8] Wanhill RJH. 1.04 - Milestone Case Histories in Aircraft Structural Integrity. Compr. Struct.
Integr., vol. 1, 2007, p. 61–72. doi:10.1016/B0-08-043749-4/01002-8.
[9] Pearson S. Initiation of fatigue cracks in commercial aluminium alloys and the subsequent
propagation of very short cracks. Eng Fract Mech 1975;7. doi:10.1016/0013-7944(75)90004-1.
[10] Meggiolaro MA, Miranda ACO, Castro JTP, Martha LF. Crack retardation equations for the
propagation of branched fatigue cracks. Int. J. Fatigue, vol. 27, 2005, p. 1398–407.
doi:10.1016/j.ijfatigue.2005.07.016.
[11] Nalla RK, Boyce BL, Campbell JP, Peters JO, Ritchie RO. Influence of microstructure on
high-cycle fatigue of Ti-6Al-4V: Bimodal vs. lamellar structures. Metall Mater Trans A
2002;33:899–918. doi:10.1007/s11661-002-1023-3.
[12] Caton M., Jones J., Allison J. The influence of heat treatment and solidification time on the
behavior of small-fatigue-cracks in a cast aluminum alloy. Mater Sci Eng A 2001;314:81–5.
doi:10.1016/S0921-5093(00)01916-X.
[13] Tanaka K, Nakai Y. PROPAGATION AND NON-PROPAGATION OF SHORT FATIGUE
CRACKS AT A SHARP NOTCH. Fatigue Fract Eng Mater Struct 1983;6:315–27.
doi:10.1111/j.1460-2695.1983.tb00347.x.
[14] Suresh S, Ritchie RO. Propagation of short fatigue cracks. Int Met Rev 1984;29:445–76.
doi:10.1179/imtr.1984.29.1.445.
[15] Hornbogen E, Gahr KH Zum. Microstructure and fatigue crack growth in a ??-Fe-Ni-Al alloy.
Acta Metall 1976;24:581–92. doi:10.1016/0001-6160(76)90104-8.
[16] Peralta P, Laird C. Fatigue of Metals. Phys. Metall. Fifth Ed., vol. 1, 2014, p. 1765–880.
doi:10.1016/B978-0-444-53770-6.00018-6.
[17] Eftis J, Jones DL, Liebowitz H. Load biaxiality and fracture: Synthesis and summary. Eng
Fract Mech 1990;36:537–74. doi:10.1016/0013-7944(90)90112-T.
[18] Chan KS. Roles of microstructure in fatigue crack initiation. Int J Fatigue 2010;32:1428–47.
doi:10.1016/j.ijfatigue.2009.10.005.
[19] Taylor D. Geometrical effects in fatigue: a unifying theoretical model. Int J Fatigue
2000;21:413–20. doi:10.1016/S0142-1123(99)00007-9.
[20] Bellett D, Taylor D. The effect of crack shape on the fatigue limit of three-dimensional stress
concentrations. Int J Fatigue 2006;28:114–23. doi:10.1016/j.ijfatigue.2005.04.010.
[21] Ritchie RO, Knott JF, Rice JR. On the relationship between critical tensile stress and fracture
toughness in mild steel. J Mech Phys Solids 1973;21:395–410. doi:10.1016/0022-
5096(73)90008-2.
[22] Livne A, Bouchbinder E, Fineberg J. Breakdown of linear elastic fracture mechanics near the
tip of a rapid crack. Phys Rev Lett 2008;101. doi:10.1103/PhysRevLett.101.264301.
[23] Riemelmoser FO, Pippan R. Consideration of the mechanical behaviour of small fatigue
cracks. Int J Fract 2002;118:251–70. doi:10.1023/A:1022928722764.
[24] Frost NE, Dugdale DS. The propagation of fatigue cracks in sheet specimens. J Mech Phys
Solids 1958;6:92–110. doi:10.1016/0022-5096(58)90018-8.
[25] Ritchie ROO. Influence of microstructure on near-threshold fatigue-crack propagation in ultra-
high strength steel. Met Sci 1977;11:368–81. doi:10.1179/msc.1977.11.8-9.368.
[26] Molent L, Jones R, Barter S, Pitt S. Recent developments in fatigue crack growth assessment.
Int J Fatigue 2006;28:1759–68. doi:10.1016/j.ijfatigue.2006.01.004.
[27] Zhai T, Wilkinson AJ, Martin JW. Crystallographic mechanism for fatigue crack propagation
through grain boundaries. Acta Mater 2000;48:4917–27. doi:10.1016/S1359-6454(00)00214-
7.
[28] Pol??k J, Kruml T, Obrt??k K, Man J, Petrenec M. Short crack growth in polycrystalline
materials. Procedia Eng., vol. 2, 2010, p. 883–92. doi:10.1016/j.proeng.2010.03.095.
[29] Mayén J, Serna SA, Campillo B, Flores O. Short crack initiation and growth kinetics analysis
in a microalloyed steel plate using rotating bending fatigue modified notched specimens.
Mater Sci Eng A 2013;582:22–8. doi:10.1016/j.msea.2013.06.039.
[30] Narasaiah N, Ray KK. Initiation and growth of micro-cracks under cyclic loading. Mater Sci
Eng A 2008;474:48–59. doi:10.1016/j.msea.2007.05.026.
[31] Aiken LS, West SG, Pitts SC. Multiple Linear Regression. Handb Psychol 2003:481–507.
doi:10.1051/eas/1466005.
[32] Man J, Obrtlík K, Polák J. Extrusions and intrusions in fatigued metals. Part 1. State of the art
and history†. Philos Mag 2009;89:1295–336. doi:10.1080/14786430902917616.
[33] Karuskevich M, Karuskevich O, Maslak T, Schepak S. Extrusion/intrusion structures as
quantitative indicators of accumulated fatigue damage. Int J Fatigue 2012;39:116–21.
doi:10.1016/j.ijfatigue.2011.02.007.
[34] Forsyth PJE. The physical basis of metal fatigue. BLACKIE SON LTD, LONDON, 1969, 200
P 1969.
[35] Hirth SM, Marshall GJ, Court SA, Lloyd DJ. Effects of Si on the aging behaviour and
formability of aluminium alloys based on AA6016. Mater Sci Eng A 2001;319–321:452–6.
doi:10.1016/S0921-5093(01)00969-8.
[36] Marinelli MC, Krupp U, Kübbeler M, Hereñú S, Alvarez-Armas I. The effect of the
embrittlement on the fatigue limit and crack propagation in a duplex stainless steel during high
cycle fatigue. Eng Fract Mech 2013;110:421–9. doi:10.1016/j.engfracmech.2013.03.034.
[37] Polák J, Man J, Obrtlík K. AFM evidence of surface relief formation and models of fatigue
crack nucleation. Int. J. Fatigue, vol. 25, 2003, p. 1027–36. doi:10.1016/S0142-
1123(03)00114-2.
August, 2016

Highlights

 Interaction between the surface cracks and aluminum microstructure were identified.

 Recorded length crack paths were later analyzed to study the crack growth by Frost and Dugdale exponential
crack growth equation (ECG) and its comparison with a proposed model (PM) developed by multiple linear
regression.

 The ECG is reliable only for the stage where the crack growth is mainly a consequence of stress concentration at
crack tip; in contrast, the PM do properly describes the nonlinearity behavior between short and long crack
growth.

Das könnte Ihnen auch gefallen