You are on page 1of 35

Accepted Manuscript

Fatigue crack growth simulation under cyclic non-proportional mixed mode


loading

Ying Yang, Michael Vormwald

PII: S0142-1123(17)30195-0
DOI: http://dx.doi.org/10.1016/j.ijfatigue.2017.04.014
Reference: JIJF 4324

To appear in: International Journal of Fatigue

Received Date: 21 February 2017


Revised Date: 26 April 2017
Accepted Date: 27 April 2017

Please cite this article as: Yang, Y., Vormwald, M., Fatigue crack growth simulation under cyclic non-proportional
mixed mode loading, International Journal of Fatigue (2017), doi: http://dx.doi.org/10.1016/j.ijfatigue.2017.04.014

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.
Fatigue crack growth simulation under cyclic
non-proportional mixed mode loading

Ying Yang a,b, c*, Michael Vormwald b


a
School of Automotive Engineering, Wuhan University of Technology, Wuhan, Hubei,
430070, China
b
Materials Mechanics Group, Technische Universitt Darmstadt, Franziska-Braun-Str.
3, D-64287 Darmstadt, Germany
c
Hubei Key Laboratory of Advanced Technology for Automotive Components,
Wuhan, Hubei, 430070, China

Abstract
An algorithm based on linear elastic fracture mechanics for three-dimensional
fatigue crack growth simulation under non-proportional mixed-mode loading is
proposed in the present paper. The crack growth behaviour in thin-walled, hollow
cylinders with a notch under combined non-proportional cyclic tension and torsion
loadings are investigated with the finite element program ABAQUS and a 3D
fracture analysis software FRANC3D. Different mixed-mode crack path prediction
criteria are evaluated to estimate the crack growth direction. Fracture mode
transition is observed, some influence factors are discussed. Crack growth cycles
calculated based on the effective stress intensity factors satisfactorily match the
experimental data for specimens tested under lower loading levels.

Keywords
Fatigue crack growth, Linear elastic fracture mechanics, Non-proportional loading,
Mixed mode

*Corresponding author: Ying Yang Tel: +86 18670810648

E-mail address: yycx2626@163.com


1. Introduction

When a crack encounters a mixed mode loading, the crack growth path differs
from the path under a normal tension loading case, and a deviation will occur.
Research in this topic has encountered many problems and challenges in both the
theoretical and the experimental fields. In recent years, a series of criteria have been
proposed for predictions of the crack propagation direction under mixed mode
loading. This paper attempts to investigate and explain fatigue crack growth
behavior under non-proportional mixed mode loadings, including the crack growth
path and fatigue life assessment.
Since Erdogan and Sih [3] first proposed the maximum tangential stress (MTS)
criterion for mixed mode fracture prediction in the year of 1963, research in the field
of mixed mode crack growth has attracted more and more attention. The minimum
strain energy density criterion [4] (S-Criterion, Sih 1974), the maximum energy
release rate criterion [5] (Nuismer, 1975), the maximum tangential strain criterion [6]
(Chambers, 1991) and so on were suggested subsequently. The various criteria do not
provide largely different results for the predicted crack propagation paths under
monotonic mixed mode loading. This conclusion is supported by a large quantity of
experimental data and theoretical analyses [7] [8] (Maiti, 1983; Maiti, 1984), if the
cracks growth is controlled by mode I dominated (tension mode) fracture. However,
large distinction between crack propagation direction and the predicted crack kink
angle were also observed in some recent experiments. In these cases, the crack
growth occurred in the plane where KII values are maximum. As a result the
maximum shear stress (MSS) criterion [9] was proposed subsequently (Maccagno,
1992). MSS criterion is available for the predictions of crack growth driven by mode
II dominated fracture (shear mode). Although all the above mentioned mixed mode
fracture criteria have been deduced from monotonic loading case, they can be
implemented straightforward to cyclic proportional mixed mode loading [10]
(Highsmith, 2009). According to Bold et al. [11] (1992), this is due to the maximum
mechanical parameter ranges (stress or strain) under proportional loading is in direct
proportion to the maximum value of these parameters in the crack tip region.
Fatigue crack growth under non-proportional mixed mode loading is a more
complex situation and there are insufficient hypotheses for application. The very
limited researches in this field generally focus on experimental analysis. Seven
relevant factors which influence the crack growth behavior under non-proportional
loading have been identified in a recent review by Vormwald and Zerres [12] (2012).
In the present paper, some of these factors (such as mode-mixity, crack closure and
different phase angle) are taken into account by an approach based on linear elastic
fracture mechanics.

2. Experiment and material information

The current paper focuses on the numerical simulation of 3-dimensional fatigue


crack growth under non-proportional mixed mode loading. The corresponding
experimental data were obtained by Brning [13] (2008). They have been discussed
by Zerres et. al [14] [15] (2010, 2011). The aluminium alloy AlMg4.5Mn specimens
and steel S460N specimens were tested with the geometry as shown in figure 1. The
material properties are shown in the following Table 1[13] [16] [21] (Brning, 2008;
Vormwald,1986; Vormwald,2011). Combined non-proportional cyclic tension and
torsion loading with phase angles of 45 and 90 were applied to the specimens with
an R ratio of -1. The results of 2 aluminium specimens marked as A7 and A8 and 2
steel specimens marked as S13 and S7 are discussed.

3. Simulation procedure

Figure 2 illustrates the procedure of this algorithm. Three main modules can be
considered as: (a) determining the crack initiation position; (b) calculating of the
maximum equivalent stress intensity factor Keq in one load cycle, which is taken as
the crack driving force parameter; (c) crack growth process. Module (b) and (c) are
repeated until the crack growth paths can be presented clearly.

3.1 Location of crack


The crack initiation will occur in the notch root at the position where the locally
uniaxial notch stress is maximum in one loading cycle. A 3-dimensional finite
element model was created using ABAQUS for stress analysis. The mesh was refined
in the crack growth region as shown in Fig. 3 (a). Along the notch root, there are four
potential crack initiation sites numbered as 1, 2, 3 and 4 clockwise, shown in Fig.3 (b).
The analysis indicated that the maximum tangential stress in one loading cycle along
the notch root on the inside edges is almost 2 times larger than the tangential stress
on the outside edge. Therefore a corner crack is considered to be a reasonable shape
to initial crack simulation. As can be expected, the initial corner crack will propagate
through the thickness of the notch and extend to the outside surface of the specimen
in the crack growth process.
An incipient corner crack was inserted into the inside edge of the notch root
where the tangential stress is maximum by program Fracture Analysis Code 3D
(FRANC3D), see figure 4. This program package is designed to simulate crack
growth in engineering structures or components with arbitrary crack geometry
under different loading and boundary conditions. FRANC3D adaptively remeshes
an existing finite element model with the inserted initial crack to generate a cracked
model. Furthermore, FRANC3D is also used to calculate stress intensity factor (SIF)
for an entire crack front.

3.2 Crack driving force calculation


The uncracked finite element model in the current simulation was created in
ABAQUS. The typical working flow for ABAQUS/FRANC3D analysis is illustrated
in figure 5. In ABAQUS the whole model is divided as the local model and the global
model. Boundary conditions and loadings are applied on the global model. The
initial cracks are inserted in the local model and cracks only grow in the local model,
so the local model is also called as the crack growth region. The local model is
inputted to the FRANC3D to insert the initial cracks in the first step and to calculate
SIFs, extend the cracks in the following steps.The cracked local model needs to be
merged with the global model, so that the stress analysis can be performed in
ABAQUS. Stress analysis results for the cracked model from ABAQUS are used to
calculate stress intensity factors in FRANC3D.

In the present paper, mode I and mode II SIFs, KI and KII in one cycle under
non-proportional tension combined with torsion loading are computed for the sake
of obtaining the equivalent SIF, Keq. Erdogan and Sih proposed that the equivalent
stress intensity factor according to the maximum tangential stress criterion can be
described by:

1 * 3 * 3 * 3 *
Keq 3cos cos
K sin sin K (1)
4 2 2 4 2 2
* is a function of KI and KII, representing the direction where the tangential stress
in the vicinity of the crack tip is maximum. And also the shear stress r vanishes
along this direction on the basis of the MTS criterion.

2
1 K 1 K
2arctan
*
8
4 K 4 K
(2)


The stress intensity factors for mode I and mode II are calculated by FRANC3D
software for a unit load in tension F=1 and in torsion MT=1. The combined tension
and torsion mixed mode stress intensity factors KI and KII are obtained from the
superposition according to equation (3):

K I t K I,F F t K I,M M T t
K II t K II,F F t K II,M M T t (3)

The combined SIFs computed from equation (3) are inserted to equation (2) to
find the angle *, which is put into equation (1). The equivalent stress intensity factor
Keq is calculated for a full loading cycle. Among all the equivalent stress intensity
factors, the peak value Keq,max is taken as the crack driving force, which means the
loads at this instant tmax, F(tmax) and MT(tmax) are used for crack propagation
simulation.

3.3 Crack growth process


The load corresponding to the Keq,max, F(tmax) and MT(tmax) are applied to the
cracked finite element model. Stress analysis is carried out in ABAQUS. The SIF
solutions for the cracks are again obtained from FRANC3D based on the stress
analysis. Although the solutions for KI, KII and Keq are already at hand for t= tmax, the
re-evaluation of this loading case combination can now be used to let the FRANC3D
software create the crack propagation increment. In FRANC3D, a 3-dimensional
crack front is a space curve comprising a series of extrapolated nodes. For each node,
the corresponding increment anode i is a value in a given proportion to a user defined
crack growth increment auser. This proportion is related to the stress intensity factors
along the crack front. A new node is extended in front of the old node with a distance
of anode i as shown in figure 6 calculated based on equation (4).

n
Knode i
anode i = auser (4)
Keq,mean

The ABAQUS/FRANC3D interface was implemented to perform the crack
growth analysis. The finite element modeling and the stress analysis are performed
in ABAQUS, FRANC3D is used to calculate crack growth parameters and updates
the crack geometry and mesh. This process is continued until the crack has grown to
a certain length.
The KI and KII of a node in the crack front and the KI and KII for the whole crack
front under F(tmax) and MT(tmax) for the initial crack of specimen A8 are shown in
figure 7.

For the intial crack of specimen A8, the maximum mode I stress intensity factor
component is more than 10 times larger than the maximum mode II values. This
indicates that the mode II fracture will not affect crack growth behavior in the full
cycle, mode I fracture is determined. For this reason, the MTS criterion is used as the
primary rule for estimating the crack propagation direction in the crack growth
simulation. The MSS criterion is performed as an alternative formula applicable to
mode II controlled crack growth.
In some cases, the simulated crack growth path deviated from the experimental
crack path no matter what criterion was employed. In order to investigate the crack
growth path behavior under these complicated loading cases, a modification is
designed to keep the deviation between experimental results and simulation in an
acceptable extent. The modification is implemented manually to adjust the position
of the newly generated crack front in order to make the simulated crack increment
direction coincident with the experimental results. As a result, although the
simulated crack growth path is not perfectly identical with the experimental data, a
close match is enforced.

4. Crack growth paths


4.1 Non-proportional loading with phase angle of 45
An aluminum specimen labeled A8 (AlMg4.5Mn; Fmax = 6.5 kN; MTmax = 72 Nm)
was tested under non-proportional mixed mode loading with a phase angle of 45.
Figure 8(a) shows the crack propagation paths on the outer surface of the specimen
in the experiment and in the simulation.
As mentioned before, four potential crack initiation sites at the notch root were
numbered 1, 2, 3, and 4 clockwise (see figure 3). The corresponding cracks are called
crack 1, 2, 3, and 4, respectively. Two crack paths were observed in specimen A8.
Crack 1 is shown on the left of figure 8(a), and crack 3 is on the right. Cracks initiated
at sites 1 and 3 because the maximum tangential stresses at these two sites were
larger than those at sites 2 and 4 in one loading cycle. In the simulation, the initiation
angle was approximately 38.4 (-38.4). As can be seen in the figure, the crack
propagation trajectories in the simulation were similar to those in the experiment.
Cracks propagated along an approximate angle of 40 to the longitudinal axis of the
specimen initially and then deflected gradually until the paths transformed into
planes, which were perpendicular to the longitudinal axis.
Because the two crack growth trajectories showed antisymmetry in both the
experiment and the simulation, the behaviors of these two cracks should coincide.
Taking crack 3 to analyze the crack path behavior, the simulated crack 3 can be seen
in figure 9. Three stages of the crack path are shown in this figure: The crack path
from notch root to point A was predicted by the MTS criterion in FRANC3D; that
from point B to the end was predicted by the MSS criterion in FRANC3D; and the
path AB was fitted from the experimental result. To provide a clear comparison
between the paths predicted by the different criteria, two estimated results are also
shown in figure 9. The path indicated by squares is the estimated result calculated
based on the MTS criterion, and the path indicated by triangles is that based on the
MSS criterion. Until point A, the crack propagation trajectory was satisfactorily
estimated by the MTS criterion. When the MTS criterion was applied continuously,
the estimated solution showed a downward crack path propagating in nearly the
original direction with slight deflection. However, from point A, the real crack path
deviated from the prediction based on the MTS criterion. This implies that the crack
propagation direction was no longer normal to the maximum tangential stress in the
vicinity of the crack tip. Another criterion was therefore used to attempt to calculate
the crack kink angle from this point. The MSS criterion was applied here as an
alternative rule to predict the crack path. The crack path starting from point A clearly
did not follow the route estimated by either the MTS criterion or the MSS criterion
based on the estimated solutions. The path according to the MSS criterion showed an
upward trend from the original crack growth direction. A path somewhere between
the predictions of the MTS and MSS criteria that gradually deviated from the original
direction appears to be more reasonable. At present, no criteria are available that can
be used to calculate crack path AB. As a result, path AB in FRANC3D is a fitting
curve based on the experimental data. From point B, the crack path flattened
(perpendicular to the longitudinal axis of the specimen), and the MSS criterion
provided a successful prediction. The crack path calculated in FRANC3D matched
the real crack path well. In contrast, large divergence was observed when the MTS
criterion was used to predict the crack path, as shown in figure 9, from the estimated
solution.
Generally, the MTS criterion predicted that the crack path maintained a nearly
straight line with a constant angle to the longitudinal axis, with slight deflection. The
flat crack path was satisfactorily described by the MSS criterion. As we already
know, based on the LEFM analysis, the MTS criterion is valid for tensile mode
fractures, and the MSS criterion is used for shear mode problems. Both of these
criteria are supported by a large quantity of corresponding experiments. Based on
the crack path behavior analysis, two distinct fracture modes and fracture mode
transition were observed in the experiment.

A steel specimen labeled S13 (S460N; Fmax = 22.5 kN; MTmax = 272 Nm) was
subjected to cyclic tension combined with torsion loading with an out-of-phase angle
of 45. The crack paths in the experiment and those in the simulation are shown in
figure 8(b). Crack 1 is on the left of this figure, and crack 3 is on the right. The crack
propagation trajectories were similar to the paths in specimen A8. Cracks initiated at
sites 1 and 3. The angle calculated based on finite element stress analysis was
approximately 40.6 (-40.6). Initially, both cracks propagated along an angle of
approximately 40 (-40) to the longitudinal axis. The final paths flattened and were
perpendicular to the longitudinal axis of the specimen. Fracture mode transition
from tensile to shear was also observed in this specimen.

4.2 Non-proportional loading with phase angle of 90


The other non-proportional loading tested in the experiment was with an
out-of-phase loading angle of 90. Figure 10(a) shows photographs of crack paths in
an aluminum specimen labeled A7 (AlMg4.5Mn; Fmax = 8.0 kN; MTmax = 96 Nm) and
the simulated crack paths. Two cracks were observed: crack 1 (on the left in the
figure) and crack 3 (on the right of the figure). From the photographs, it was found
that both crack paths were very curvilinear, and crack branching was observed in the
experiment. When this test was performed, the simulation algorithm did not cover
crack branching, and therefore, the simulation stopped before any crack branching
could occur. According to the stress analysis, the crack initiation angle was 51.9.
Taking crack 1 to analyze the crack path behavior, the path first propagated upward
following a smooth curve, and as the crack grew, the extension direction changed to
downward, which was opposite to the original orientation.

The crack paths in this specimen were different from those in specimens under
loading with a phase angle of 45. No flat crack path was observed in the experiment.
The result of analysis of crack path behavior is shown in figure 11. From the notch
root to point A, the smooth crack path can be described well by the MTS criterion.
The corresponding fracture was tensile mode fracture. Deviations between the
simulated path and the experimental path became evident as the crack extended.
Starting from point A, the MSS criterion was used as the alternative rule to predict
crack growth direction. Estimation shows that both the prediction from the MTS
criterion and that from the MSS criterion did not provide acceptable results, because
the real crack path was somewhere between these two estimations. When the crack
length reached point B, the downward crack path was fully controlled by the MSS
criterion. From this moment, the crack was controlled under shear mode fracture.
Path AB was considered to signify transition mode and was fitted from the
experimental data.
Steel specimen S7 was tested under tension combined with torsion loading with
Fmax = 27 kN and MTmax = 408 Nm. Four cracks were observed in the experiment. In
specimen S7, four crack paths were observed to be nearly symmetrical in both the
experiment and the simulation, as shown in figure 10(b). According to the FEA, the
maximum tangential stresses along the notch root at sites 1to 4 were nearly equal.
The calculated crack initiation angle was 53.1 (-53.1). Because the four cracks
initiated at nearly the same loading cycles in the experiment, initial cracks of the
same size were inserted into the corresponding positions in FRANC3D. The entire
simulated crack path was calculated from FRANC3D based on the MTS criterion,
which showed that crack propagation complied with tension controlled fracture, and
no fracture mode transition took place.

5. Crack growth cycles


Fatigue life includes the loading cycles expended in crack initiation and those
expended during crack propagation. In the present paper, the number of cycles
expended from the initial crack to that of the final crack is calculated based on Paris
law, the number of loading cycles consumed in crack initiation is obtained from
experimental data.

5.1 Crack closure effect


The maximum equivalent stress intensity factor Keq,max in one loading cycle is
used to calculate the instant forces corresponding to crack propagation. Under cyclic
loading, the stress intensity factor range K is associated with the crack deflection
angle as well as fatigue life. When loading ratio R = -1, crack closure must affect
fatigue life significantly. Usually, K is substituted for the effective stress intensity
factor range Keff to take account of the loading ratio effect. Here, Keff is defined as
Kmax Kop, where the subscript op denotes a crack opening stress intensity factor.
Newmans crack opening stress equation [18] is one of the most widely used
methods for estimating opening stress and is given by:
op
= A0 + A1 R + A2 R2 + A3 R3 for R > 0
max
(5)
op
= A0 + A1 R for - 2 R < 0
max

Where
1

A0 = 0.825 - 0.34 + 0.052 cos max (6)


20

max
A1 = (0.415 - 0.071) (7)
0

A2 = 1- A0 - A1 - A3 (8)
A3 = 2 A0 + A1 - 1 (9)

Here, 0 is the material flow stress, which Newman proposed to be


Rp0.2 + Rm
0 = (10)
2
where Rp0.2 is the monotonic yield stress and Rm is the material ultimate tensile
strength. In this paper, max is defined as the maximum tangential stress at the notch
root for the uncracked specimen, and 0 is calculated based on equation (10). The
effect of three-dimensional constraint is described via a constraint factor ,
corresponding to = 1 under plane stress conditions and = 3 under plane strain
conditions. An approximate equation describing the constraint factor based on
FEA by Newman [20] is written as:
1.5
= 1.15 + e- Kn (11)


where K n K max 0 B , = 1.25, and = 0.85. Here ,B represents specimen

thickness, which is considered to be 2.5 mm for the current specimen.

5.2 Aluminum specimen


Table 2 shows details of the loading and stress for the three aluminum
specimens investigated in this paper, where max represents the maximum tangential
stress along the notch root in uncracked specimens, and 'y is the material cyclic
yielding stress. The ratio max/'y which is used to assess the plasticity level for
specimens A8 and A7, was 0.378, 0.453, respectively. The maximum tangential
stresses in the specimens were all lower than the material cyclic yielding stress and
met small scale yielding requirements.

The number of cycles was computed on the basis of equation (12), which can be
integrated straightforwardly as follows:
N af 1
0
dN =
ai C (Keff )m
da

(12)
The calculation result for specimen A8 is shown in figure 12(a), which shows
loading cycles versus crack length c in the notch thickness direction and loading
cycles versus crack length a on the outer surface. From the crack path behavior
analysis, three fracture modestensile, transition, and shearwere determined for
this specific crack path. The calculated number of cycles agrees satisfactorily with the
experimental data. In the simulation, cracking in the thickness direction arrested at
c=2.2mm, corresponding to 182570 cycles. The number of loading cycles expended
along this direction in the experiment was 175010. Crack growth occurred on the
outer surface, as shown in figure 12(a) by the two dashed lines, which highlight the
crack length corresponding to the different fracture modes. Tensile mode occurred
from crack length a = 0 mm to a = 16.72 mm (point A). From a = 16.72 mm to a = 26.49
mm (point B), the crack was controlled by transition mode fracture. When the crack
was longer than 26.49 mm, shear mode fracture occurred.
The slope of loading cycles versus crack length curve can be used to evaluate
the crack growth rate. In the case of tensile mode fracture, the slope of the simulated
curve increased slightly with crack propagation. Generally, crack growth rates in the
simulations were similar to those observed in the experiments. The cycle numbers
under tensile mode fracture in the simulation and those observed in the experiment
were 387766 and 353000, respectively. When entering transition mode, the crack
growth rate increased continually as shown by the simulated curve, whereas the
slope of the experimental curve appeared to remain the same as the slope in the
tensile mode stage. The numbers of cycles at the end of the transition mode was
418082 in simulation and 445000 in the experiment.

Figure 12(b) shows the number of cycles versus crack length curves for specimen
A7. Along the thickness direction, the simulated crack growth curve exactly matched
the experimental data. Crack length reached c = 2.11 mm after 80541 cycles. In the
experiment, 77000 cycles were consumed in the thickness direction, corresponding to
c = 2.5 mm. As discussed above, crack extension conforms to tensile mode fracture in
the notch thickness direction. When the crack extrapolated to the outer surface of the
specimen, the tensile mode controlled crack grew steadily until a = 11.29 mm. The
slope of the simulated curve and that of the experimental curve are almost the same,
with only a slight difference with increasing crack growth. The number of cycles
corresponding to tensile mode fracture was 193134 in the simulation and 175010 in
the experiment. The transition mode was shorter than that in specimen A8, from a =
11.27 mm (point A) to a = 15.26 mm (point B).The crack growth rate in the simulation
was slightly lower than the real crack propagation rate. Loading cycles increased to
212517 from tensile mode to transition mode based on the calculation, and 195010
cycles were observed during the real crack growth, corresponding to a = 15.5 mm.
From both the simulation and the experimental data, the crack growth rate increased
when it appeared to be controlled by shear mode fracture.
The predicted fatigue life was 234381 cycles at final crack length a = 23.49 mm.
The real fatigue life in the experiment was 242721 cycles, at a = 51.0 mm. Crack
length under tensile mode fracture occupied only approximately 20% of the total
crack length, but more than 70% of loading cycles were consumed under tensile
mode fracture. This result is the same as that for specimen A8 and implies that mode
I fracture is the dominant fracture mode even in such complex mixed mode loading
cases.

5.3 Steel specimen


Loading and stress information is shown in table 3, including the maximum
tangential stress along the notch root and material cyclic yielding stress for
specimens S13 and S7, indicating that these two specimens are tested under
relatively high loading levels. The ratio max/'y was 1.17 for specimen S13 and 1.57
for specimen S7. Both values were above 1. Small scale yielding requirements for
LEFM may not be satisfied, particularly for specimen S7.

Figure 13(a) shows the simulated crack growth curve and the experimental data
for specimen S13. An obvious distinction between these results and those obtained
for the aluminum specimens discussed above is that large deviations exist between
the calculated crack growth cycles and the experimentally obtained data along the
thickness direction. The experimental curve is sharper than the simulated curve,
which means that the crack grew faster in the experiment than in the simulation.
From crack initiation to crack growth through the notch thickness, 14000 cycles were
expended in the experiment. In contrast, the number of calculated loading cycles was
35015 (corresponding to crack length c = 2.13 mm), which is double the real number
of cycles. Plasticity, which was not taken into account in the simulation, may be the
main reason for the acceleration of crack growth observed in the notch region. The
residual stresses caused by inhomogeneous plastic deformations, which is not
considered in this simulation may also contribute to this phenomenon.

As the crack extended to the outer surface and propagated continually, the notch
effect weakened. The crack was not affected by the notch effect when the crack
length was approximately equal to the notch root radius, which was 2 mm in the
present specimen. It can be seen from the crack growth curve for crack length a that
the calculated curve appears parallel to the experimental curve from a = 2 mm. These
parallel curves indicate that the region near the crack front can be treated as linearly
elastic when the crack reached a particular length even under relatively high loading
cases. From the experimental data, approximately 69% of fatigue life was spent
during tensile mode fracture growth. This supports the conclusion that mode I
fracture was the dominant fracture mode.
A similar phenomenon was observed in specimen S7. Larger deviations were
found between the simulated crack growth curve and the experimental crack growth
curve in the notch thickness direction. Parallel curves were observed in crack growth
along the a direction. The number of calculated crack growth cycles consumed in the
c direction was 33288, corresponding to c = 2.11 mm, which is nearly six times more
than the 5515 cycles observed in the experiment. Considering that specimen S7 was
subjected to a higher loading level than that of specimen S13, the effect of plasticity
and residual stresses were expected to be intensified. The acceleration of crack
propagation in the notch region was more obvious in this specimen. Because of the
plasticity and residual stresses, the simulation overestimated the fatigue life as 73544
cycles, compared with 27400 cycles in the experiment. During the crack growth,
including crack propagation in the thickness direction and on the outer surface,
mode I stress intensity factors played a dominant role in crack extension because
they were approximately 10 times larger than the mode II values throughout crack
growth. Tensile mode fracture therefore appropriately describes the entire crack
path. No fracture mode transition was detected.

6. Fracture mode transition


When discussing crack growth under mixed mode loading, fracture mode
transition must be considered. The transition from shear to tension is published in
literature[17], while for the transition from tension to shear, few literatures can be
found and the criterion or the theory to describe this behavior is absent.
As described in the above analysis and discussion, the employed MTS criterion
or the MSS criterion cannot completely reproduce the whole crack growth
mechanism for specimen A8 and A7, due to the fracture mode transition from
tension to shear. Many factors influence this transition behavior, for example,
material properties, the value of tension and torsion loadings, the loading level, the
phase angle of loadings and the material properties. An key factor for the current
specimen is the geometry. A special geometry that had an increasing thickness was
introduced into the current specimen as shown in figure 14. The increasing thickness
meant that there was more material that restricted crack growth in a relatively
narrow region. The crack paths were forced to become perpendicular to the
longitudinal axis to reduce the propagation resistance and make the fracture mode
transition from tension to shear.

7. Conclusion
In the present paper, 3-dimensional crack growth simulation is implemented by
using the LEFM-based algorithm. The simulation results including two out phase
angles loading of 45 and 90. Fracture mode transition from tensile to shear was
observed in the crack path behavior analysis. Crack paths belonging to the transition
stage cannot be predicted by either the MTS criterion or the MSS criterion. The actual
crack path is located somewhere between the predictions of these two criteria. In
other specimens, crack growth is only controlled by the MTS criterion, no fracture
mode transition is observed.
The estimated fatigue lives were computed using Paris law and the effective
stress intensity factors. For the small scale yielding specimens (tested under low
loading levels), the assessment of loading cycles were in reasonable agreement with
the experimental results. For the specimens subjected to higher loading levels, the
calculated cycle numbers overestimated the fatigue lives. In these cases, crack
growth rates in the notch region were much faster in the experiment than in the
simulation. When the crack a on the outer surface of the specimen reached
approximately a notch radius (2mm) length, the crack propagation rate in the
simulation began to match that in the experiment.

ACKNOWLEDGMENT
The authors express their gratitude to the Deutsche Forschungsgemeinschaft
(DFG) for providing financial support under grant VO729/13-1 and the Fundamental
Research Funds for the Central Universities (WUT: 2015IVA022).

REFERENCES
[1] Abaqus Analysis Users Manual, Version 6.12-EF.
[2] FRANC3D, Fracture Analysis Consultants, Inc., Ithaca, NY, USA; Cornel Fracture
Group, Cornell University, USA.
[3] Erdogan, F., Sih, G.C., On the Crack Extension in Plates Under Plane Loading and
Transverse Shear. J. Basic Engineering, 85D (1963) 519-527.
[4] Sih, G.C., Strain-energy-density factor applied to mixed mode crack problems. Int.
J. Fracture, 10 (1974) 305-321.
[5] R. J. Nuismer, An energy release rate criterion for mixed mode fracture. Int. J.
Fracture, 11 (1975) 245-250.
[6] Chambers. A. C., Hyde. T. H., Webster. J. J., Mixed mode fatigue crack growth at
550C under plane stress conditions in Jethete M152. Engng. Frac. Mech, 39 (1991)
603-619.
[7] S.K. Maiti, R.A. Smith, Comparison of the criteria for mixed mode brittle fracture
based on the preinstability stress-strain field, Part I: Slit and elliptical cracks under
uniaxial tensile loading.Int. J. Fracture, 23 (1983) 281-295.
[8] S.K. Maiti, R.A. Smith, Comparison of the criteria for mixed mode brittle fracture
based on the preinstability stress-strain field, Part II: Pure shear and uniaxial
compressive loading. Int. J. Fracture, 24 (1984) 5-22.
[9] Maccagno, T.M., Knott, J.F., The mixed mode I/II fracture behaviour of lightly
tempered HY 130 steel at room temperature. Engng. Frac. Mech. 41 (1992) 805-820.
[10] Highsmith Jr., S., Crack path determination for non-proportional mixed-mode
fatigue. PhD Thesis, Georgia Institute of Technology, USA, (2009).
[11] Bold, P.E., Brown, M.W. and Allen, R.J. A review of fatigue crack growth in
steels under mixed mode I and II loading. Fatigue Fract. Engng Mater. Struct.
150(1992) 965-977.
[12] P. Zerres, M. Vormwald, Review of fatigue crack growth under
non-proportional mixed-mode. Int. J. Fatigue, 58 pp. 75-83. 2014.
[13] Brning, J., Untersuchungen zum Rissfortschrittsverhalten unter
nichtproportionaler Belastung bei elastisch-plastischem
Materialverhalten-Experimente und Theorie. Institut fr Stahlbau und
Werkstoffmechanik der Technischen Universitt Darmstadt, report 85 ISBN
978-3-939195-14-6. (2008)
[14] P. Zerres, J.Brning, M. Vormwald, Fatigue crack growth behavior of
fine-grained steel S460N under proportional and non-proportional loading. Eng
Fract Mech. 77 (2010) 1822-1834.
[15] P. Zerres, J. Brning, M. Vormwald, Risswachstumsverhalten der
Aluminiumlegierung AlMg4.5M unter proportionaler und nicht-proportionaler
Schwingbelastung. Mat. Testing, 03 (2011) 109-117.
[16] M. Vormwald, T.Seeger, Riausbreitungsverhalten der Werkstoffe StE460 und
AlMg4.5Mn. Report FD-1/1986, Fachgebiet Werkstoffmechanik,TU Darmstadt (1986).
[17] D. Haboussa, T. Elguedj, B. Lebl, A. Combescure, Simulation of the
shear-tensile mode transition on dynamic crack propagations. Int. J. Fracture, 178
(2012) 195-213.
[18] J.C. Newman, Jr., A crack opening stress equation for fatigue crack geowth. Int.
J. Fracture, 24 (1984) 131-135.
[19] G. Savaidis, M. Dankert, T. Seeger, An analytical procedure for predicting
opening loads of cracks at notchs. Fatigue Fract. Engng Mater. Struct. 18 (1995)
425-442.
[20] Newman, J. C., Jr, Crews, J. H., Jr, Biglew, C. A., Dawicke, D.S. Variations of a
global constraint factor in cracked bodies under tension and bending loads. ASTM
STP 1244 (1995) 21-42.
[21] M. Vormwald, Ermdungslebensdauer von Baustahl unter komplexen
Beanspruchungsablufen am Beispiel des S460. Materials Testing, 53 (2011) pp.
98-108.
[22] Hos, Yigiter, Freire, Jos L F, Vormwald, Michael, Measurements of strain fields
around crack tips under proportional and non-proportional mixed-mode fatigue
loading. Int. J. Fatigue, 89 (2016) pp. 87-98.
[23] Hos, Yigiter, Vormwald, Michael, Experimental study of crack growth under
non-proportional loading along with first modeling attempts. Int. J. Fatigue, 92 (2016)
pp. 426-433.
Figure captions:
Fig.1. Specimen and Notch geometry (all dimensions in mm).
Fig. 2. Simulation algorithm procedure.
Fig. 3. Uncracked finite element model and crack number.
Fig. 4. Crack insertion and remeshed local model.
Fig. 5. ABAQUS/FRANC3D working flow.
Fig. 6. New crack front fit curve
Fig. 7. (a) KI and KII of a node in a full loading cycle (b) KI and KII for the whole
crack front under F(tmax) and MT(tmax)
Fig. 8. (a)A8 Crack paths in experiment (top) and simulation (bottom) (b) S13 Crack
paths in experiment (top) and simulation (bottom).
Fig. 9. Crack path 3 in A8.
Fig. 10. (a)A7 Crack paths in experiment (top) and simulation (bottom) (b) S7 Crack
paths in experiment (top) and simulation (bottom).
Fig. 11. Crack path 1 in A7.
Fig. 12. (a) Crack growth curve (A8, crack 3) (b) Crack growth curve (A7, crack 1).
Fig.13. (a) Crack growth curve (S13, crack 1) (b) Crack growth curve (S7, crack 3).
Fig.14. Specimen geometry and crack growth region
Fig. 1. Specimen and Notch geometry (all dimensions in mm)
Fig.2. Simulation algorithm procedure
reference point

(a) (b)

Fig. 3. Uncracked finite element model and crack number


Fig. 4. Crack insertion and remeshed local model
Fig. 5. ABAQUS/FRANC3D working flow
anode i

Fig. 6. New crack front fit curve


(a) (b)
Fig. 7. (a) KI and KII of a node in a full loading cycle (b) KI and KII for the
whole crack front under F(tmax) and MT(tmax)
z

x
y

(a) (b)

Fig. 8 (a)A8 Crack paths in experiment (top) and simulation (bottom) (b) S13 Crack paths in
experiment (top) and simulation (bottom)
Fig .9. Crack path 3 in A8
(a) (b)

Fig. 10 (a)A7 Crack paths in experiment (top) and simulation (bottom) (b) S7 Crack paths in
experiment (top) and simulation (bottom)
Fig .11. Crack path 1 in A7
(a) (b)

Fig. 12. (a) Crack growth curve (A8, crack 3) (b) Crack growth curve (A7, crack 1)
(a) (b)

Fig. 13. (a) Crack growth curve (S13, crack 1) (b) Crack growth curve (S7, crack 3)
crack growth region

Fig .14. Specimen geometry and crack growth region


Table captions:
Table 1 Mechanical properties.
Table 2 Loading and stress for aluminum specimens.
Table 3 Loading and stress information for steel specimens.
Table 1 Mechanical properties

E Rp0.2 Rm A5 y ' C m
Material
[MPa] [-] [MPa] [MPa] [%] [MPa] [-] [-]
AlMg4.5Mn 68000 0.33 169 340 20.2 341 3.510-7 2.9
S460N 208500 0.3 500 643 26.2 410 8.810-9 3.15

da/dN in [mm/cycle], Keff in [MPam0.5]


Table 2 Loading and stress for aluminum specimens
Fmax MTmax Phase angle max 'y
Specimen max/'y
(kN) (Nm) () (MPa) (MPa)
A8 6.5 72 45 128.8 341 0.378
A7 8.0 96 90 154.6 341 0.453
Table 3 Loading and stress information for steel specimens

Fmax MTmax Phase angle max 'y


Specimen max/'y
(kN) (Nm) () (MPa) (MPa)
S13 22.5 272 45 479 410 1.17
S7 27.0 408 90 644.6 410 1.57
Highlights
Crack growth simulation under non-proportional mixed-mode loading is
performed.
Fracture mode transition from tensile to shear is observed in some specimens.
Crack paths of transition stage cannot be predicted by either MTS or MSS
criterion.

You might also like