Vous êtes sur la page 1sur 15

Fatigue & Fracture of Engineering Materials & Structures 1998; 21: 1159–1173

THREE-DIMENSIONAL FINITE ELEMENT SIMULATION OF CRACK


EXTENSION IN ALUMINIUM ALLOY 2024FC
G. L,* A. C and K.-H. S
GKSS-Forschungszentrum Geesthacht GmbH, Max-Planck-Str., 21502 Geesthacht, Germany

Abstract—A simulation of 3D crack extension using a cohesive zone model (CZM) has been carried out
for a side-grooved compact tension specimen and a surface-crack tension specimen of aluminium 2024FC.
Detailed finite element calculations were conducted by assuming crack extension only along the crack
plane (mode I). For comparison, a 2D plane strain simulation is also presented. Load, displacement and
crack extension histories are predicted and compared with the experiment. It is shown that the 2D
approximation appears to agree reasonably well with experimental results, and that the 3D calculation
gives very good agreement with test data. The determination of the CZM parameters is also discussed.
Numerical results show that the CZM is a workable computational model which involves only a few
microstructurally motivated phenomenological parameters for crack extension simulation.
Keywords—Cohesive zone model; Surface crack; 3D crack extension; Finite element method.

NOMENCLATURE

a =initial crack length


0
Da=crack extension
B=specimen thickness
B =net section thickness for the side-grooved specimen
net
CMOD=crack mouth opening displacement
CT=compact tension specimen
CTOD=crack tip opening displacement
E=Young’s modulus
J=J-integral
LSY=large scale yielding
R =reference length for the size of plastic zone defined as 1/3p (K /s )2
0 0 0
SCT=surface cracked tension specimen
SSY=small scale yielding
T =elastic T -stress
T =cohesive stress
0
d =displacement at the crack tip over a gauge length of 5 mm
5
C=cohesive zone energy
C =critical cohesive zone energy (the work of fracture process)
0
n=Poisson’s ratio
s =yield stress
0
t=normalized elastic T -stress, t=T /s
0
v =reference length for the size of the cohesive zone defined as EC /T 2
0 0 0

INTRODUCTION

In two previous studies [1,2], a cohesive zone model (CZM) has been applied to simulate
mode I crack extension under conditions of small scale yielding (SSY) and large scale yielding
(LSY). The finite element calculations were conducted in the context of 2D plane strain and plane
stress. Systematic parametric studies have been carried out to investigate the dependence of the
fracture resistance on the solid properties, cohesive zone parameters, crack geometries and mode

* Current address: Center for Simulation of Advanced Rockets, Computational Science and Engineering, University of
Illinois at Urbana-Champaign, 1304 W. Springfield Avenue, Urbana, IL 61801, USA.

© 1998 Blackwell Science Ltd 1159


1160 G. L et al.

of loading. In this paper, the CZM has been extended to simulate 3D crack extension. Two-
dimensional interface elements have been developed for this purpose.
The initiation of mode I crack extension in ductile materials can be described by a critical value
of the J-integral [3]. Subsequent to initiation, increasing values of J may accompany crack
extension, and the consequent relation between values of J and the crack extension, Da, is called
the crack growth resistance curve, J (Da). In the past, the function J (Da) was often tacitly regarded
R R
as a material property. However, a strong geometry dependence of crack growth resistance, which
is associated with different levels of crack tip constraint under large scale yielding conditions, has
been observed by a number of investigators [4,5]. A numerical study by Brocks and Olschewski
[6] has shown that a strong interaction between in-plane and out-of-plane constraint effects on
crack-front fields in a 3D configuration cannot be captured in plane strain analyses. Several
proposals [7–12] have been suggested to quantify the crack tip constraint. However, they are still
not suitable to quantify the constraint effect on crack growth resistance. Therefore, a full 3D crack
extension simulation is desired to study the constraint effect on fracture behaviour.
The concept of the cohesive zone model (CZM) originates from the works of Dugdale [13] and
Barenblatt [14] to describe the non-linear material behaviour at a crack tip by means of cohesive
forces in the so-called process zone. This work has been re-considered later with some modifications
by other authors [1,2,15–27]. In the CZM, the fracture process is treated as a microscale strip
(cohesive zone) and is represented by a continuum traction-separation law. With increasing loading,
the cohesive zone grows and separates to form new crack surfaces, thereby advancing the crack.
Several traction-separation laws have been proposed by different authors [16–18,21,22,24].
Needleman [16] used an exponential type of traction-separation law to simulate the particle
debonding in metal matrices. Tvergaard and Hutchinson [21] applied a trapezoidal shape of the
traction-separation law to calculate the crack growth resistance. Yuan et al. [20] used a rectilinear
shape of the traction-separation law to study the ductile fracture in an aluminium alloy by using
the experimental resistance curves for simulating crack extension. Subsequently, Lin and co-workers
[1,2,26,27] extended this approach to model the crack growth resistance in homogeneous as well
as heterogeneous materials under small scale yielding (SSY) and large scale yielding (LSY)
conditions.
The CZM employed in this paper has been used in a previous study by Lin [27] on the ductile
fracture in homogeneous solids and heterogeneous material systems. Following the description of
the traction-separation law in Ref. [27], the cohesive zone is characterized by the normal traction
which is assumed to be constant, i.e. T (d)=T , where d is the normal separation across the cohesive
0
zone. A local fracture criterion is introduced by assuming the cohesive energy, C , to be constant
0
for both crack initiation and extension. The model is embedded within a conventional elastic–
plastic continuum described by material parameters, e.g. Young’s modulus, Poisson’s ratio, yield
stress and strain hardening exponent. Once the CZM parameters are calibrated by a set of
experimental crack extension data [25,27], the approach permits prediction of the macroscopic
behaviour, e.g. load, displacement and crack growth resistance [1,2], including the states where
stability is lost [26]. It was shown that the CZM is a phenomenological and computationally
efficient model involving only two parameters. Moreover, the model is also applicable to model
3D crack extension. This development will be discussed in detail in this paper.
The CZM is firstly applied to simulate the fracture in a round notched tensile bar (RNTB). The
specimens are made of aluminium 2024FC. Finite element calculations are performed under
axisymmetric conditions. The determination of the CZM parameters is discussed. Detailed finite
element analyses (FEA) based on the CZM are then carried out for simulating crack extension in
a side-grooved compact tension (CTsg) specimen and a semi-elliptical surface crack panel (SCT)

Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173 © 1998 Blackwell Science Ltd
Crack extension in aluminium alloy 2024FC 1161

under tension made of the same material. The numerical analyses are conducted under mode I
loading conditions, and it is assumed that the crack grows in the crack plane only, which is also
experimentally observed. Three-dimensional finite element analyses have been performed for both
specimens. A 2D plane strain calculation for the CTsg specimen has also been carried out to study
the difference between 2D and 3D predictions. The global mechanical performance, e.g. load versus
CMOD and crack extension history, etc. has been predicted. It is demonstrated that the 2D
approximation appears to agree reasonably well with experimental results and that the 3D
calculation gives a very good agreement with test data. The crack front behaviour is also predicted
by the model: the phenomenon of the canoe-effect in the surface crack tension panel observed in
the experiment is well reproduced by the simulation.

THE COHESIVE ZONE MODEL

Figure 1(a) defines the general mechanics problem: a band consisting of the cohesive material
with inelastic deformation emanates from a crack/notch in a component. The band is localized in
that its thickness is much smaller than any other length characterizing the component. The
inelasticity in the band is modelled as an array of springs with a non-linear traction-separation
law:
T =T (d) (1)
The cohesive energy is then given by:

P
df
C=
T (d) dd (2)
0
where d is the maximum separation. In the present study, a constant traction-separation law is
f
used, i.e.
T (d)=T (3)
0
where T is called the cohesive stress and may be interpreted as a ‘micro’ material strength to
0
resist the local material separation. The fracture mechanism is represented by a local critical

Fig. 1. Schematic representation of the CZM. (a) An inelastic band is regarded as an array of continuously
distributed springs. ( b) Graphical interpretation of fracture resistance.

© 1998 Blackwell Science Ltd Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173
1162 G. L et al.

fracture energy, C ,
0
C=C (4)
0
C is the work of separation per unit area of crack surfaces which may be understood as local
0
fracture energy per unit area of crack extension required for fracture without plastic deformation
in the surrounding solid of the cohesive material. It is noted that the above equation is assumed
to be valid for a given fracture mechanism. Furthermore, C can be understood as a ‘micro’ material
0
toughness. Both C and T are regarded as fracture mechanism-based material properties. C and
0 0 0
T can be determined by experiments [25,27].
0
Figure 1( b) shows schematically the basic idea of the model. The total work of fracture per unit
area of crack extension is approximately decomposed into two parts: a constant cohesive zone
energy, C , required for fracture; and plastic dissipation, C , in the surrounding solid required for
0 p
the constant cohesive zone energy, C .
0
C =C +C (5)
R 0 p
By definition of Eq. (4), the work required to rupture a cohesive element is the work of the fracture
process, C . At crack initiation, C is much smaller than C , so that the fracture resistance at crack
0 p 0
initiation, denoted by C , is about equal to C . By contrast, C can be much greater than C for the
i 0 p 0
subsequent crack extension. As has been noted, C is assumed to be constant at crack initiation
0
and subsequent crack extension, while C depends on the constraint situation of the cracked body.
p
If one assumes that the J-integral is valid for crack extension, then C is identical to J .
R R
An important feature of the CZM is that the traction-separation law (3) introduces a reference
stress, T , and a reference length, d =C /T , the maximum separation in the cohesive zone. Together
0 f 0 0
with Young’s modulus, E, of the surrounding solid, they form a material length,
CE
v = 0 (6)
0 T2
0
varying from a few nanometers for atomic decohesion to a few centimeters for metal voiding. This
material length determines the size of the inelastic deformation region. To be definite, the model
is now interpreted in the context of ductile metals, even though conclusions may be applicable to
other inelastic mechanisms. Thus, the traction-separation law for the ductile material system should
be regarded as a phenomenological characterization of the zone where the separation takes place.
The cohesive zone model described above is implemented into the finite element program
FECGS [28] using an augmented Lagrangian treatment, which was proposed by Landers and
Taylor [29] to solve the contact problem. Compared to traditional Lagrange multiplier and
penalty techniques, it has the advantage of no additional equations to the discretized system, and
no additional ill-conditioning induced.

MATERIAL AND SPECIMENS

The analyses were conducted on the two specimen types shown in Fig. 2, compact tension with
side grooved (CTsg) and surface crack tension (SCT) specimens. All specimens were made of
2024FC aluminium alloy. Figure 3 plots the uniaxial tensile stress–strain curve obtained from the
experiment. This was used as the input of the stress–strain response in the finite element analyses.
The mechanical constants of this aluminium alloy are summarized as follows:
Young’s modulus: E=72 000 MPa
Poisson’s ratio: n=0.3

Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173 © 1998 Blackwell Science Ltd
Crack extension in aluminium alloy 2024FC 1163

Fig. 2. Schematic of crack geometries. (a) Side-grooved compact tension specimen. (b) Surface-crack
tension specimen.

Fig. 3. Experimental true stress–strain curve for aluminium alloy 2024FC in uniaxial tension.

Mean hardening: n=3.3


Yield stress: s =80.7 MPa
0.2
Ultimate tensile stress: s =224 MPa (s =259 MPa)
u u
Strain at s : e =15.6%
u u
It is noted that, in the finite element calculations, the limit of elasticity, s =60.3 MPa, was taken
e
as the yield stress instead of s .
0.2

DETERMINATION OF THE COHESIVE ZONE PARAMETERS

Based on the studies of the crack growth resistance under conditions of Mode I SSY and LSY
in two previous studies [1,2], it is concluded that the crack initiation always takes place at C #C ,
i 0
which is independent of the crack tip constraint arising from geometry and loading mode. As
noted above, C is identical to J at least for a certain amount of crack extension. Therefore, it is
R R
suggested that C can be taken as the value of the J-integral at crack initiation as determined from
0
any kind of fracture toughness testing specimen. The J value should give a very good approximation
i
© 1998 Blackwell Science Ltd Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173
1164 G. L et al.

Fig. 4. Predictions of a round notched tensile bar. (a) Geometry. (b) Comparison of predicted and
measured load versus displacement curves.

of C . It is noted that the specimen should be fatigue precracked to correspond to the initially
0
sharp crack tip and the J value should be carefully evaluated to obtain a real initiation J value.
i
Theoretically, the definition of the J here should be the J value at Da=0. Hancock, Reuter and
i
Parks have shown that at a small amount of crack extension, e.g. Da=200 mm, the J value for
centre-cracked panels can be approximately four times greater than that of the highly constrained
deeply cracked bend bars and CT specimens. However, they have also shown that when Da
approaches zero, the J value is essentially independent of the constraint. A suitable procedure for
determining J can be found in Ref. [30]. Once we have determined C , the second CZM parameter,
i 0
T , can be obtained by the common way of fitting the crack growth resistance curve. It should be
0
noted that now only one parameter needs to be fitted, thus the determination of this parameter
through a fitting procedure is unique. Alternatively, the parameter T can also be determined
0
through fitting the load–deformation curve of a notched tensile bar (Fig. 4). This will be addressed
in the following.
Firstly, from experiments on CTsg specimens, J was determined as 10 N/mm and set equal to
i
C . The parameter T was then determined through fitting the tensile curve of a round notched
0 0
tensile bar. The finite element calculation was carried out under the condition of axisymmetry.
The finite element mesh is shown in Fig. 4(a). Four-node linear axisymmetric elements with a
selected integration scheme were used in the calculations. Figure 4( b) shows the results predicted
using three values for T ==400, 420 and 430 N/mm2. The thick solid line is the experimental
0
curve. The solid, dotted and dashed lines correspond to T =400, 420 and 430 N/mm2, respectively.
0
It is seen that T =420 N/mm2 exactly reproduces the fracture of the tensile bar. Thus, T was
0 0
taken as 420 N/mm2.

THREE-DIMENSIONAL FINITE ELEMENT ANALYSES

Three-dimensional finite element analyses were conducted on the two specimens shown in Fig. 2
by using the finite element code FECGS. Two-dimensional interface elements have been developed

Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173 © 1998 Blackwell Science Ltd
Crack extension in aluminium alloy 2024FC 1165

Fig. 5. Finite element mesh for the compact tension specimen. (a) One-quarter of specimen. ( b) Crack tip
mesh.

for simulating the 3D crack extension using the CZM, see Ref. [28] for details of the elastic–
plastic finite element analysis with crack extension. Comparisons were made between measured
and calculated global mechanical behaviour, e.g. load against displacement and crack growth
resistance, for both specimens. The results will be reported in the following.

Side-grooved compact tension specimen


Geometry and finite element model
Figure 2(a) shows the analysed specimen. The dimensions of the specimens are also shown in
the figure. Due to symmetry, only one-quarter of the geometry needs to be analysed, the finite
element mesh of which is shown in Fig. 5(a). The element arrangement in the vicinity of the crack
front is shown in Fig. 5( b), which corresponds to the dashed rectangle in Fig. 5(a). Six layers of
elements were used through the half-net section thickness (z-direction) in the region near the crack
tip. Each layer has equal thickness. One extra layer was used for the side-groove part of the
specimen, as shown in Fig. 5(b). The side-groove is modelled by a sharp notch. Three-dimensional
20-node quadratic elements with 14 reduced integration points were used in the calculations. Two-
dimensional nine-node quadratic interface elements were arranged on the ligament plane to
simulate crack extension. The transition elements with 22 nodes, elements with variable numbers
of nodes [31], were used to make the mesh compatible with nine-node quadratic interface elements.
The size of the regular small elements near the crack tip was 0.2 mm. The crack tip was modelled
as initially sharp. The FE mesh for the specimen under study consisted of 1258 elements and 7081
nodes. This mesh was generated by FEMESH [32].

Results
Figure 6(a) displays load–CMOD curves for the CTsg specimen. Here, CMOD is measured at
the centre plane of the specimen for the 3D analysis. However, no difference is observed between
the centre and surface planes. The thick solid line indicates the experimental test curve. Two-
dimensional calculation was also conducted in the context of the plane strain assumption to study
the difference between 2D and 3D predictions. The dotted line corresponds to the 2D plane strain
prediction with net section force. The dashed line corresponds to the 2D plane strain prediction
with effective section force, and the effective section was calculated by B 앀B Ω B . The open
eff= net
circle indicates the crack initiation as determined in the experiment and the open square is the
crack initiation numerically predicted. The 3D prediction by the CZM agrees excellently with the
experimental records. The 2D plane strain simulation with net section force underestimates
the experimental curve, whereas the result of the 2D plane strain calculation with effective section
force overestimates the load. Moreover, the crack initiation predicted by the 3D finite element
analysis agrees very well with the experimentally measured one.

© 1998 Blackwell Science Ltd Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173
1166 G. L et al.

Fig. 6. Predictions of a CTsg specimen. (a) Comparison of the predicted and experimental load–
deformation (CMOD) curves. ( b) Variation of crack extension along the thickness direction for various
loading stages. (c) Comparison of experimental and predicted crack fronts. The top figure is from
experimental measurement, and the bottom figure is from numerical prediction. The dashed line in the
top figure corresponds to the curve indicated by an arrow in the bottom figure. (d) Comparison of the
predicted and experimental CMOD–resistance curves.

Figure 6( b) shows the predicted crack extension along the thickness direction for several loading
stages. Crack initiation started at the free surface. It is observed that the crack extended much
faster at the free surface than in the centre plane of the specimen. This is due to the side groove
at the free surface of the specimen, whereby the crack tip constraint near the free surface is much
higher than that in the centre plane. A comparison of experimentally observed and numerically
predicted crack front is shown in Fig. 6(c). The top sketch of Fig. 6(c) shows the experimental
measurement, and the bottom sketch of Fig. 6(c) shows the predicted crack fronts corresponding
to the loading stages in Fig. 6( b). The dashed line in the top sketch is the predicted one indicated
by an arrow in the bottom sketch, which is taken from the last record in the FE calculation. It is

Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173 © 1998 Blackwell Science Ltd
Crack extension in aluminium alloy 2024FC 1167

Fig. 7. Curves of d versus CMOD and values of d along the crack front for the CTsg specimen.
5 5
(a) Comparison of the predicted and experimental d –CMOD curves. (b) Distribution of d along the
5 5
thickness direction for several loading stages.

seen that the prediction agrees very well with the experimental data, except the region near the
side-grooved free surface. It is noted that the FE calculation was terminated much earlier than the
experiment. However, as has been shown in Fig. 6(b), after a certain amount of crack extension,
the shape of the crack front does not change, and thus the comparison is reasonable.
The CMOD as a function of crack extension Da is plotted in Fig. 6(d). Again, the values of
CMOD were measured at the centre plane. The experimental curve is displayed as a solid thick
line. The thin solid line indicates CMOD versus crack extension, Da, at the centre plane from the
3D finite element analysis. The dashed line is the 2D plane strain prediction. It is seen that the
3D prediction using the CZM agrees very well with the experimental results, whereas the 2D
prediction overestimates the experimental CMOD.
Figure 7(a) compares the numerically predicted and experimentally measured d versus CMOD
5
curves. Again, the experimental curve is shown as a solid thick line. Note that d is a type of
5
CTOD measured over a gauge length of 5 mm at the crack tip [33]. The dotted line is for the 2D
plane strain prediction. Here, a discrepancy between the predicted and measured curves is observed.
The numerical prediction underestimates the d . Presumably, this is attributed to the side groove
5
simulated as a sharp notch, although actually its root was a blunt notch. The result is an
overestimation of constraint. Figure 7(b) shows the distribution of d along the thickness direction
5
for several loading stages. It is shown that the variation of d along the z-direction is minor. The
5
difference between the centre plane and free surface is within 5%.

Surface crack tension specimen


Geometry and finite element model
The specimen geometry considered is shown in Fig. 2( b). The initial crack front is semi-elliptical,
which is directly taken from the experiment. Here, mode I crack extension is only considered, thus
the crack grows only in the crack plane, which was also experimentally observed. Due to symmetry,
only one-quarter of the geometry needs to be analysed. The specimen was fatigue precracked, thus
an initially sharp crack front was assumed in the finite element analysis. Figure 8(a) displays the

© 1998 Blackwell Science Ltd Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173
1168 G. L et al.

Fig. 8. Finite element mesh for the surface crack tension specimen. (a) One-quarter of specimen. (b) Crack
tip mesh.

finite element mesh for one-quarter of the SCT specimen. The element arrangement in the vicinity
of the crack tip is shown in Fig. 8( b), which corresponds to the dashed rectangle in Fig. 8(a). The
initial crack was simulated by simply freeing the element nodes within the ellipse, as shown in
Fig. 8(b). Similar to the CTsg specimen, three-dimensional 20-node quadratic elements with 14
reduced integration scheme were used in the calculations. Transition elements with 22 nodes were
used to make the mesh compatible with nine-node interface elements. The smallest element size
was about 0.25 mm. The FE mesh for the specimen under study consisted of 1645 elements and
9044 nodes. Note that the element sizes in the CTsg and SCT specimens are not the same. However,
it has been demonstrated in a previous study [27], that the CZM has the advantage of mesh size
independence.

Results
Figure 9(a) compares predicted and experimental load–CMOD curves for the SCT specimen.
The solid and open circles indicate the data sets obtained on two identical specimens [34]. The
result from the 3D finite element analysis using the CZM is shown as a solid curve in the figure.
The finite element calculation was continued up to almost the maximum load. One can see that
the FE prediction by the CZM agrees very well with the experimental records.
Figure 9( b) shows the predicted crack extension along the crack front at several load levels.
Crack initiation is found to start at an angle h of about 30°. With increasing load, a canoe-shaped
crack front is formed. The crack propagates mostly along the crack front from h=25° to h=90°.
The maximum crack extension is found between h=30° and h=40°. Up to the highest applied
load, only a small amount of crack extension is observed at the free surface, h=0°. Since the FE
calculation was terminated at an early stage of loading, there is no experimental result at this load
level for comparison.
d versus CMOD is plotted in Fig. 10(a) for h=0°, 30° and 90°. Figure 10( b), a cross-plot of
5
Fig. 9(a), shows the d distribution along the crack front. Here, d is normalized by the correspond-
5 5
ing d at the centre plane of the specimen. It can be seen that d changes significantly before the
5 5
crack starts to grow. The maximum d is obtained at the free surface. The minimum d is observed
5 5
at the centre plane, h=90°. d decreases continuously along the crack front from h=0° to 90°,
5
which is much different from the behaviour observed for the through-crack specimen CTsg. With
increasing load, the d at the free surface increases much faster than that of the centre plane. This
5
Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173 © 1998 Blackwell Science Ltd
Crack extension in aluminium alloy 2024FC 1169

Fig. 9. Predictions of SCT specimen. (a) Comparison of predicted and experimental load–deformation
(CMOD) curves. ( b) Crack extension along the crack front under different loading stages.

Fig. 10. Curves of d versus CMOD and variation of calculated d values along the initial crack front
5 5
for the SCT specimen. (a) d versus CMOD curves. (b) d normalized by d at the centre plane (d )
5 5 5 5-centre
along the crack front under several loading stages.

also implies that the constraint along the crack front changes significantly for the surface crack.
However, after the crack has initiated, the variation of d along the crack front remains almost
5
constant.

DISCUSSION

The GKSS version of the CZM utilizes a constant traction-separation law, Fig. 1(a), with the
traction, T . Earlier studies [20,25] have shown that the shape of this law is of no importance.
0
Although the shape assumed by other authors, e.g. Refs. [16,21], in principle seems to have more
physical relevance to the crack tip conditions, the exact shape is not known. For these reasons,

© 1998 Blackwell Science Ltd Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173
1170 G. L et al.

the more convenient rectangular shape is readily justified. In the development of the model,
particular emphasis has been placed on generating an engineering tool for solving practical
problems. Therefore, an important aim of the present analysis was to validate the model by means
of experiments. As described in previous studies [25,27], the CZM parameters were determined by:
$ Equating the cohesive energy, C , to the J-integral at initiation of a pre-existing crack, J , as
0 i
determined in an R-curve test; and
$ Fitting T to the R-curve, or, alternatively, by determining it from a notched tensile specimen.
0
The CZM assumes that these quantities can be regarded as parameters representing the fracture
behaviour of a given material the observed size and geometry dependence of the crack growth
resistance curve is then due to the response of plastic deformation to a variation of these quantities.
This has been perfectly demonstrated by our results: crack extension is achieved by unzipping the
material ahead of the crack with C and T as the controlling parameters, the rest is a 3D elastic–
0 0
plastic analysis. The predicted load–CMOD curves of both specimen types, Figs 6(a) and 9(a),
and the CMOD–Da curve of the CTsg specimen, Fig. 6(d), are in perfect agreement with experimen-
tal data. It is also seen that a 2D analysis gives much less accurate answers [Fig. 6(a) and (d)].
A further point of interest concerns the development of the crack shape. The predicted crack
front agrees well with experimental observation, except in the region near the side groove [Fig. 6(c)].
The accelerated crack extension near the surface of the CTsg specimen seems to be somewhat
unrealistic. This is attributed to the side-groove simulated as a sharp notch [Fig. 5( b)], although
actually its root was a blunt notch as required by the test method [30]. Figure 7(a) confirms this
interpretation: d is underestimated as compared to the experiment, due to too high constraint.
5
For the surface crack, the predicted crack shape development [Fig. 9( b)] follows what has been
observed in experiments [34], in particular the maximum crack extension between h=30° and
40°. Since the FE calculation was terminated much earlier than the experiment, therefore, a direct
comparison with measured Da characteristics was not possible.
The excellent performance of the CZM is also confirmed by an earlier study [26] which showed
the ability of the CZM to handle the constraint problem caused by a soft layer between two elastic
substrates, where the layer is under very different constraint conditions when its thickness varies.
Other means of varying the constraint conditions can also be dealt with: small scale yielding
studies by Tvergaard and Hutchinson using their trapezoidal traction-law [22], and by Lin and
Cornec [1], predicted the well-known effect of the T -stress on the crack growth resistance.
Our results suggest that the CZM is able to solve the transferability problem. From the above
discussion, a workable strategy can be derived:
$ A standard test on a high constraint specimen serves for the material characterization and
for determining C and T .
0 0
$ Alternatively, T can be determined by means of a notched tensile test.
0
$ Transferability of the thus determined fracture properties to a structural component is achieved
by a 3D elastic–plastic FE analysis.
As compared to the Gurson–Needleman–Tvergaard (GTN) model, the CZM does not claim to
model the actual fracture mechanisms in the process zone. The CZM is tuned via the parameters
C and T which are easily accessible by experiments. The physical meaning of C and T is covered
0 0 0 0
by the void initiation and extension modelling of the GTN model.
Recently, Shih and Xia [35] presented a fracture model which consists of a thin layer behaving
as a Gurson material in a continuum field. They introduced a length scale, i.e. the thickness of the
layer, which can be identified as the mean spacing between the voids. Starting with an initial

Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173 © 1998 Blackwell Science Ltd
Crack extension in aluminium alloy 2024FC 1171

porosity, f , the crack growth resistance curve is modelled for a given material constitutive law
0
and a given geometry. Their results bear strong similarities to our present and previous [1] results,
if we connect the initial porosity f of their model with the cohesive stress T in our traction-
0 0
separation law. A higher T corresponds to a lower f .
0 0
From the present work, an interesting side effect emerges: the usefulness of the CTOD definition
of d , which is used at GKSS for fracture mechanics testing and flaw assessment procedures [33]
5
is supported by the findings illustrated by Figs 7(b) and 10( b). In the CTsg specimen, d not vary
5
by more than 3% across the thickness, which means that d measured at the specimen’s surface
5
as required by the test technique [33] represents the overall behaviour of the specimen very well.
In the case of the surface crack, the variation of d along the crack front is roughly 25%. Since
5
the GKSS flaw assessment method ETM is aimed at estimating the d occurring at the surface
5
[36], a conservative assessment of a surface-flawed component can be achieved.

CONCLUSIONS

The version of the Cohesive Zone Model (CZM) developed at GKSS has been used to model
3D cases of crack extension and load–deformation behaviour. The 3D cases were represented by
a side-grooved CT specimen and a tensile plate containing a semi-elliptical surface crack. Previous
experiments on these specimen types were used to validate the CZM analyses. The main conclusions
are as follows:
$ Concrete values can be assigned to the CZM parameters, e.g. C and T , by means of
0 0
experimental calibration. The term C can be set equal to the experimental J-integral at
0
initiation of crack extension, while T can be obtained from a fit to the experimental crack
0
growth resistance, or, alternatively, to the load–deformation curve of a notched tensile bar,
thus providing a unique set of parameters for a given material. This enables the CZM to go
beyond parameter studies and to treat specific engineering problems.
$ Predicted load–CMOD curves agree excellently with the experimental results.
$ The predicted development of the crack front geometry in the surface-cracked tensile specimen
resulted in the same canoe-shape as observed in the experiments.
$ In the side-grooved CT specimen, the crack advanced more rapidly near the surface than
near the centre plane of the specimen. This is attributed to constraint induced by the
side-grooves.
$ In the side-grooved CT specimen, the d CTOD was almost constant along the crack front,
5
whereas d of a surface crack was about 25% higher at the surface than near the deepest point.
5
$ It has been demonstrated that the CZM is capable of handling 3D crack problems.
Acknowledgement—The authors thank the Deutsche Forschungsgemeinschaft for the support of this work as part of the
Sonderforschungsbereich 371.

REFERENCES
1. G. Lin and A. Cornec (1995) Simulation of crack growth resistance curves using a Cohesive Zone Model:
Small Scale Yielding. In: Berichtsband der 27 T agung des DVM-Arbeitskreises Bruchvorgänge, Köln,
pp. 265–278 (in German).
2. A. Cornec and G. Lin (1996) Simulation of crack growth resistance curves using a Cohesive Zone Model:
Large Scale Yielding. In Berichtsband der 28 T agung des DV M-Arbeitskreises Bruchvorgänge, Bremen,
pp. 199–214 (in German).
3. J. R. Rice (1968) A path independent integral and the approximate analysis of strain concentration by
notches and cracks. J. Appl. Mech. 35, 379–386.
4. J. W. Hancock, W. G. Reuter and D. M. Parks (1993) Constraint and toughness parameterized by T. In:

© 1998 Blackwell Science Ltd Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173
1172 G. L et al.

Constraint EVects in Fracture, ASTM STP 1171, American Society for Testing and Materials, Philadelphia,
pp. 21–40.
5. J. A. Joyce and R. E. Link (1994) Effect of constraint on upper shelf fracture toughness. In: AST M ST P
Proceedings of 26th National Symposium on Fracture Mechanics (Edited by W. G. Reuter, J. H.
Underwood and J. C. Newman), American Society for Testing and Materials, Philadelphia.
6. W. Brocks and J. Olschewski (1986) On J-dominance of crack-tip fields in largely yielded 3D structure.
Int. J. Solids Structures 22, 693–708.
7. W. Brocks, G. Künecke, H.-D. Noack and H. Veith (1989) On the transferability of fracture mechanics
parameters to structures using FEM. Nuclear Engng Design 112, 1–14.
8. Z.-Z. Du and J. W. Hancock (1991) The effect of non-singular stresses on crack-tip constraint. J. Mech.
Phys. Solids 39, 555–567.
9. N. P. O’Dowd and C. F. Shih (1991) Family of crack-tip fields characterized by a triaxiality parameters—
I. Structure of fields. J. Mech. Phys. Solids 39, 989–1015.
10. N. P. O’Dowd and C. F. Shih (1992) Family of crack-tip fields characterized by a triaxiality parameters—
II. Fracture applications. J. Mech. Phys. Solids 40, 939–963.
11. C. F. Shih, N. P. O’Dowd and M. T. Kirk (1993) A framework for quantifying crack tip constraint. In:
Constraint EVects in Fracture, ASTM STP 1171, American Society for Testing and Materials,
Philadelphia, pp. 2–20.
12. G. Lin, A. Cornec and K.-II. Schwalbe (1994) Two parameters characterization of crack front fields in
thin ductile center-cracked geometries. In Structure Integrity: Experiments–Models–Applications,
Proceedings of 10th European Conference on Fracture, Berlin, Germany, Vol. 1, pp. 341–348.
13. D. S. Dugdale (1960) Yielding of steel sheets containing slits. J. Mech. Phys. Solids 8, 100–104.
14. G. I. Barenblatt (1962) The mathematical theory of equilibrium cracks in brittle fracture. Adv. Appl.
Mech. 7, 55–129.
15. A. H. Cottrell (1963) Mechanisms of fracture. In: T ewksbury Symposium on Fracture, University of
Melbourne, pp. 1–27.
16. A. Needleman (1987) A continuum model for void nucleation by inclusion debonding. J. Appl. Mech.
54, 523–531.
17. A. Needleman (1990) An analysis of tensile decohesion along an interface. J. Mech. Phys. Solids
38(3), 289–324.
18. A. Needleman (1990) An analysis of decohesion along an imperfect interface. Int. J. Fracture 42, 21–40.
19. J. Planas and M. Elices (1991) Nonlinear fracture of cohesive material. Int. J. Fracture 51, 139–157.
20. H. Yuan, A. Cornec and K.-H. Schwalbe (1991) Numerical simulation of ductile crack growth on thin
aluminium CT specimens using the cohesive zone model (in German). In: Proceedings of 23.
Vortragsveranstaltung des DV M-Arbeitskreises Bruchvorgänge, Berlin, pp. 221–237.
21. V. Tvergaard and J. W. Hutchinson (1992) The relation between crack growth resistance and fracture
process parameters in elastic–plastic solids. J. Mech. Phys. Solids 40, 1377–1397.
22. V. Tvergaard and J. W. Hutchinson (1994) Effect of T-stress on Mode I crack growth resistance in a
ductile solid. Int. J. Solids Structures 31, 823–833.
23. A. Cornec, H. Yuan and G. Lin (1994) Cohesive zone model for ductile fracture. In: Numerical Predictions
of Deformation Processes and the Behavior of Real Materials, Proceedings of the 15th RISO International
Symposium on Material Science, pp. 269–275.
24. V. Tvergaard and J. W. Hutchinson (1996) On the toughness of ductile adhesive joints. J. Mech. Phys.
Solids 44, 789–800.
25. H. Yuan, G. Lin and A. Cornec (1996) Verification of a cohesive zone model for ductile fracture. J. Engng
Mater. T echnol., ASME T rans. 118, 109–116.
26. G. Lin, Y.-J. Kim, A. Cornec and K.-H. Schwalbe (1997) Fracture simulation of undermatched interleaf
in bending using a cohesive zone model. In: Mis-matching of Interfaces and Welds (Edited by K.-H.
Schwalbe and M. Koçak), GKSS Research Center Publications, Geethacht, FRG, pp. 343–356.
27. G. Lin (1997) Numerical investigation of crack growth behaviour using the cohesive zone model. Ph.D.
thesis, Technical University Hamburg, Hamburg.
28. G. Lin (1997) A finite-element program system (FECGS). GKSS internal technical report.
29. J. A. Landers and R. L. Taylor (1986) An augmented lagrangian formulation for the finite element
solution of contact problem. NCEL Contract Report, Naval Civil Engineering Laboratory, Port
Hueneme, CA.
30. ESIS procedure for determining the fracture behaviour of materials, ESIS P2-92, January 1992.

Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173 © 1998 Blackwell Science Ltd
Crack extension in aluminium alloy 2024FC 1173

31. T. J. R. Hughes (1987) T he Finite Element Method—L inear Static and Dynamic Finite Element Analysis,
Prentice-Hall, Englewood Cliffs, NJ.
32. G. Lin (1997) A finite element program system (FECGS)—pre-processor FEMESH. GKSS internal
technical report.
33. K.-H. Schwalbe (1995) Introduction of d as an operational definition of the CTOD and its practical
5
use, fracture mechanics. In: AST M ST P Proceedings of 26th National Symposium on Fracture Mechanics
(Edited by W. G. Reuter, J. H. Underwood and J. C. Newman), American Society for Testing and
Materials, Philadelphia, pp. 763–778.
34. H.-M. Bauschke, D. T. Read and K.-H. Schwalbe (1989) Crack growth resistance at surface cracks in
three aluminium alloys. European Symposium on Elastic–Plastic Fracture Mechanics: Elements of Defeat
Assessment, Freiburg, F.R. Germany, 9–12 October.
35. C. F. Shih and L. Xia (1994) Modeling of crack growth resistance using computational cells with
microstructurally-based length scales. In: Constraint EVects in Fracture: T heory and Applications, AST M
ST P 1244 (Edited by M. Kirk and A. Bakker), American Society for Testing and Materials, Philadelphia.
36. K.-H. Schwalbe (1994) The crack tip opening displacement and J integral under plane strain control
and fully plastic conditions estimated by the Engineering Treatment Model for plane stress tension. In:
Fracture Mechanics: 24th Volume, AST M ST P 1207 (Edited by W. G. Reuter, J. H. Underwood and J. C.
Newman), American Society for Testing and Materials, Philadelphia, pp. 636–651.
37. G. Lin (1997) A finite element program system (FECGS)—post-processor FEPOST. GKSS internal
technical report.
38. K.-H. Schwalbe, B. K. Neale and J. Herrens (1994) The GKSS test procedure for determining the fracture
behaviour of materials: EFAM GTP 94, GKSS 94/E/60, GKSS Forschungszentrum Geesthacht,
Geesthacht, Germany.
39. J. R. Rice and D. M. Tracey (1969) On the ductile enlargement of voids in triaxial stress fields. J. Mech.
Phys. Solids 17, 201–217.
40. A. L. Gurson (1977) Continuum theory of ductile rupture by void nucleation and growth: Part I—yield
criteria and flow rules for porous ductile media. J. Engng Mater. T echnol. 99, 2–15.
41. G. Rousselier (1987) Ductile fracture models and their potential in local approach of fracture. Nuclear
Engng Design 105, 97–111.
42. K.-H. Schwalbe and J. Herrens (1997) R-curves testing and relevance to structural assessment. Fatigue
Fract. Engng Mater. Struct., this issue.
43. W. M. Garrison, Jr. and N. R. Moody (1987) Ductile fracture. J. Mech. Phys. Solids 48, 1035–1074.
44. D.-Z. Sun and W. Schmitt (1990) Application of micromechanical models to the analysis of ductile
fracture resistance behavior. Numerical Methods in Fracture Mechanics, Proceedings of the 5th International
Conference (Edited by A. R. Luxmoore and D. R. J. Owen), Freiburg, FRG, 23–27 April.

© 1998 Blackwell Science Ltd Fatigue & Fracture of Engineering Materials & Structures, 21, 1159–1173

Vous aimerez peut-être aussi