Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Comput. Methods Appl. Mech. Engrg.

192 (2003) 3887–3908


www.elsevier.com/locate/cma

Numerical modelling of crack propagation: automatic


remeshing and comparison of different criteria
P.O. Bouchard *, F. Bay, Y. Chastel
Centre de Mise en Forme des Mat
eriaux, Ecole des Mines de Paris, UMR 7635, B.P. 207, 06904 Sophia-Antipolis Cedex, France
Received 13 December 2002; received in revised form 18 April 2003; accepted 23 May 2003

Abstract

Modelling of a crack propagating through a finite element mesh under mixed mode conditions is of prime importance
in fracture mechanics. Three different crack growth criteria and the respective crack paths prediction for several test
cases are compared. The maximal circumferential stress criterion, the strain energy density fracture criterion and the
criterion of the maximal strain energy release rate are implemented using advanced finite element techniques. A fully
automatic remesher enables to deal with multiple boundaries and multiple materials. The propagation of the crack is
calculated with both remeshing and nodal relaxation. Several examples are presented to check for the robustness of the
numerical techniques, and to study specific features of each criterion.
 2003 Elsevier B.V. All rights reserved.

Keywords: Finite element method; Crack propagation criteria; Automatic remeshing; Strain energy release rate; Gh method; Strain
energy density

1. Introduction

Since the first studies in the early 1920s by Inglis [1], Griffith [2] and Irwin [3], the research in linear elastic
fracture mechanics has led to the development of a vast number of theories and applications. The prop-
agation of a crack in a part leads to an important displacement discontinuity. The more accurate way of
modelling such a discontinuity in a finite element mesh is to modify the part topology (due to the crack
propagation) and to perform an automatic remeshing. However, several methods have been proposed in
the literature to model this discontinuity without any remeshing. Belytschko has suggested a meshless
method (Element free Galerkin method [4]) where the discretisation is achieved by a model which consists
of nodes and a description of the surfaces of the model. More recently, two interesting finite element
techniques have been presented to deal with such discontinuities without any remeshing stage. The Strong
Discontinuity Approach (SDA) [5–7] in which displacement jumps due to the presence of the crack are

*
Corresponding author. Tel.: +33-4-93-957432; fax: +33-4-92-389752.
E-mail addresses: pierre-olivier.bouchard@ensmp.fr, bouchard@cemef.cma.fr (P.O. Bouchard).

0045-7825/$ - see front matter  2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0045-7825(03)00391-8
3888 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

embedded locally in each cracked finite element without affecting neighbouring elements. The amplitude of
displacement jumps across the crack are defined using additional degrees of freedom related to the plastic
multiplier, leading to a formulation similar to standard non-associative plasticity models. The Ex-
tended Finite Element Method (XFEM) [8–10] in which the displacement-based approximation is enriched
near a crack by incorporating both discontinuous fields and the near tip asymptotic fields through a
Partition of Unity method [11]. However, these recent techniques still have to be improved in order to deal
with complex configurations such as multiple cracks, large deformation crack propagation, etc. When
remeshing is possible, a real mesh discontinuity representing the crack seems to be more accurate. This
technique is used, in particular, in the 2D and 3D fracture numerical code FRANC [12,13]. However,
whatever the technique used, the accuracy of the crack path directly depends on the accuracy of kinking
criteria.
In a recent paper, Bouchard et al. [14] have introduced an interesting remeshing technique to model crack
propagation accurately using the discrete crack approach. This technique is used here for three different
criteria. First the numerical tools used to propagate a crack through a mesh are described. The theoretical
aspects and the numerical difficulties when dealing with the three criteria are then studied. In the third part
several examples are presented to compare each criterion and to show the robustness of the numerical
schemes. Finally a discussion about advantages and disadvantages of the three criteria is presented and
examples of propagation in multimaterial structures are investigated.

2. Numerical tools for crack propagation

The technique presented in [14] has been implemented in a finite element code based on a 2D model for
viscoplastic or elastic–viscoplastic behaviour, FORGE2 [15]. It has been developed for plane strain or
axisymmetric cases in large strain. The mesher and automatic remesher of FORGE2 is based on P2/P0
triangles––quadratic velocity and constant pressure for each element––and allows to deal with multiple
edges and multiple materials.
When a crack propagates through a mesh, the accuracy at the crack tip is of prime importance. In
FORGE2, a nodal relaxation is combined with a remeshing technique. This enables to avoid the problem of
distortion at the crack tip and to continue with a new, undistorted and well suited mesh. The transfer of
data from the old to the new mesh at each remeshing stage is performed so as to preserve the fields of the
various state variables. This transfer is a closeness-based interpolation technique which respects the plastic
criterion.
Fracture may be decomposed into two steps: the crack initiation and the crack propagation.
The stage of crack initiation is crucial but quite problematic. General fracture mechanics codes avoid this
problem because their aim is to study the evolution of a pre-existing crack. Damage-based numerical
models are more adapted to this problem because they study the evolution of damage continuously and a
crack is initiated for a critical damage value. However these codes are often unable to model the crack
propagation without a local collapse criterion.
In fact, it is very difficult to determine the location of a new crack. Micro-failure and inclusions always
induce local stress concentrations which are at the origin of failure and cracks. As all these defects cannot
be taken into account numerically (unless if we use statistical approaches), either the material is assumed to
be perfect or homogeneised, or the initiation location is imposed by positioning a pre-crack. Moreover, the
initiation of a crack in a mesh induces a severe topological change which is rarely supported by numerical
codes.
Many initiation criteria have been proposed in the literature. They often depend on the materials studied.
Some of them are based on critical values for mechanical state variables, stress or strain. An other pos-
sibility is to use a damage law and a critical damage parameter to localize the crack initiation.
P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3889

In this paper, the selected crack initiation criterion is based on a critical stress value, which is a char-
acteristic of the material. So, at each time step, the maximum principal stress value is computed and its
position determined on the boundary. Once this value reaches the critical one, a crack is initiated in two
steps:

• an internal outline is added to the geometrical definition of the domain in order to model the crack ini-
tiation;
• the automatic remeshing procedure is carried out on the new domain, and the nodes of the new outline
are split to make the opening of the crack possible (see Fig. 1).

Moreover, many numerical tools have been developed to improve the accuracy at the crack tip. A
concentric mesh around the crack tip may be coupled with singular elements [16] to model the stress field
singularity (Fig. 2a). These elements are studied in detail in [14]. A ring of elements (Fig. 2b) can also be
introduced in order to compute the strain energy release rate [17,18]. Finally mesh refinement around the
crack tip enables to keep a good precision in the vicinity of the crack. As the crack tip moves along, the
areas which need to be refined will change; a new mesh is created and refined only in the areas where it is
needed in order to optimise calculation time (Fig. 2c). More details about the mesh structure at the crack tip
may be found in [14].

3. Crack growth criteria

When a crack is initiated, one needs to check, at each time step, if the crack is going to propagate (crack
propagation criteria), and in which direction (crack kinking criteria).

Fig. 1. Crack initiation.

Fig. 2. (a) Concentric mesh with singular elements, (b) ring of elements and (c) evolutionary mesh refinement at the tip of the crack,
and unrefinement elsewhere.
3890 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

Stress intensity factors––strength singularity at the crack tip––are often used for crack propagation. In
mode I loading, they are compared with a critical value KIc of the material. For example,
pffiffiffiffiffiffiffiin the maximum
circumferential stress criterion proposed by Erdogan [19], the crack propagates when 2prrhh ¼ KIc , where
rhh is the circumferential stress, and r represents the distance to the crack tip.
Other criteria, based on energetic parameters––strain energy release rate [2], strain energy density [20]––
are also available.
For elastic–plastic materials, the crack tip is blunted by plasticity. Then the crack tip opening dis-
placement (CTOD) introduced by A.A. Wells in 1961, may be used as a material crack propagation para-
meter.
Then, determination of the crack direction may be obtained by kinking criteria:

• Some criteria are based on the local fields at the crack tip following a local approach, e.g. the maximum
circumferential stress criterion introduced by Erdogan and Sih [19], or the maximum strain criterion [21].
• Other criteria are based on the energy distribution throughout the cracked part, following a global ap-
proach, e.g. the maximal strain energy release rate criterion [22].
• Some authors have also proposed to determine the crack extension using a micro-void continuum dam-
age model [23,24]. These theories are based on the assumption that the void initialisation and the void
growth control the crack growth direction [25].

When selecting one of these criteria, one has to consider two main points:

• Does the criterion comply with the physical characteristics of the considered material?
• Can the criterion be implemented so as to be calculated accurately?

More often, the accuracy of the direction computation is directly linked to the accuracy of the numerical
computation of local or energetic parameters.
In the following, we briefly present the maximum circumferential stress criterion, the minimum strain
energy density fracture criterion and the maximum strain energy release rate criterion. For each of them,
the influence of numerical parameters on the results is pointed out.

3.1. Maximum circumferential stress criterion (MCSC)

This criterion, introduced by Erdogan [19] for elastic materials, states that the crack propagates in the
direction for which the circumferential stress is maximum. It is a local approach since the direction of crack
growth is directly determined by the local stress field along a small circle of radius r centered at the crack
tip.
The kinking angle of the propagating crack is computed by solving the following system:
8
< KII sinðh=2Þ < 0;
KI sinðhÞ þ KII ð3 cosðhÞ  1Þ ¼ 0 with h 2   p; p½; ð1Þ
:
KI > 0;
where KI and KII are respectively the stress intensity factors corresponding to mode I and mode II loading,
and h is the kinking angle. This criterion also proves the existence of a limit angle corresponding to pure
shear: h0 ¼
70:54.
This criterion is widely used in the literature, but numerical details concerning its implementation are
rarely mentioned. When stress intensity factors are not computed by the finite element software, the
computation of the kinking angle has to be based on the circumferential stress rhh at each integration point
at the crack tip (Fig. 3). The crack propagation is then performed toward the integration point that
P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3891

Fig. 3. Direction of propagation with the maximum circumferential stress criterion.

Fig. 4. Two different implementations of the MCSC leading to two different crack paths.

maximises rhh . However a direct application of this method would be mesh-dependent since the crack
direction would then directly depend on the number of elements at the crack tip (Fig. 4a)––the crack di-
rection being provided by the integration point location with respect to the crack tip.
We propose here an alternative method based on the maximal principal stress evaluated at each inte-
gration point nearest to the crack tip (Fig. 3):

• The integration points nearest to the crack tip are identified.


• For each of them, eigenvalues and eigenvectors of the stress tensor are computed.
• These eigenvalues provide the principal stresses and enable to find the direction of propagation for each
integration point.
• The final direction of the crack propagation is obtained by a weighted average of each direction with
respect to the distance between the integration point and the crack tip. This leads to a mesh indepen-
dence at the crack tip since the kinking direction is not directly associated to a specific integration point
at the crack tip.

This way of computing the crack propagation direction may be questionable since the direction is merely
determined using the maximal principal stress. The direction of the maximal principal stress at the crack tip
corresponds to the direction of the maximal tensile stress. The approximation performed here is based on
the fact that a crack propagates perpendicularly to the maximum tensile stress. Moreover, the advantage of
such a technique is to improve the mesh-independence since the direction of crack propagation is not di-
rectly associated to the position of an integration point (Fig. 4b).
These two different implementations of the same criterion can thus lead to different crack paths, even for
simple applications. Fig. 4 shows the crack path in a part containing two symmetric holes during a tensile
test.
3892 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

Details about different ways of implementing the maximum circumferential stress criterion can be found
in [26].
The maximum circumferential stress criterion is one of the most used criteria, because it is easy to im-
plement. However it may be questionable because the stress field in the immediate vicinity of the crack tip is
only approximated. Moreover the existence of a non-elastic zone at the crack tip undoubtedly modifies the
stress distribution. That is why energy-based criteria, based on global values of the structure, may be more
adequate. Their implementation is generally more complex, but they lead to more accurate results since the
crack propagation direction is computed further away from the crack tip.

3.2. Minimum strain energy density criterion (MSEDC)

Sih [20] considers that high values of strain energy, We , tend to prevent crack growth. Then  the crack
grows in the direction that minimises this energy. Let w be the strain energy density: w ¼ dW dV
e
, where We is
the sum of the volumetric part of the strain energy Wv and the pdistortion
ffiffi energy W d . This quantity is
proportional to the square of stress, and since the stress has a 1= r singularity at the crack tip (where r
represents the distance to the crack tip), the strain energy density w has a 1=r singularity. Therefore the so-
called strain energy density factor, S ¼ rw, remains bounded.
Parameter S can be computed using two different techniques:

• An analytical formulation. The strain energy density is inversely proportional to the distance r to the
crack tip [17]. Then S represents the intensity of the local energy field:
 

1þm m 2
S ¼ rw ¼ r r211 þ r222 þ r233  ðr11 þ r22 þ r33 Þ þ 2r212 :
2E 1þm
• A numerical formulation. In the simulation, the strain power Pe is computed at each time step and for
each integrationR point. A time integration of Pe gives the strain energy for each integration point of
t
the mesh: We ¼ 0 Pe dt.

In the numerical formulation, parameter S is computed by introducing a ring of elements around the crack
tip (Fig. 2b). The curve SðhÞ is plotted for each element of this ring as a function of the angle between the
centre of the element and the crack axis (Fig. 5).
The kinking angle h0 is the angle corresponding to the local minimum of the curve SðhÞ:
8 
>
> oS
>
< oh ¼ 0;
h¼h0
 2 
>
> oS
>
: P 0:
oh2 h¼h0

θ
-70˚ θ0 +70˚
Fig. 5. Strain energy density as a function of the angle h.
P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3893

One should stress that it is important to compute the local minimum of the curve, and not the global
minimum which can lead to results corresponding to angles which do not belong to ½70:54; 70:54, and
therefore are not relevant.
Moreover, the computation accuracy is quantitatively linked to the number of elements in the ring
around the crack tip.
At last, this criterion may not exactly be considered as a global criterion since the computation of the
strain energy density is based on local mechanical parameters.

3.3. Maximum strain energy release rate criterion (MSERRC)

The strain energy release rate G represents the energy required to increase the crack length by one unity.
The criterion states that among all virtual and kinematically admissible crack length displacements, the
real increase is the one which maximises the strain energy release rate. The kinking angle is then deter-
mined by
8 
>
> dG
>
< dh ¼ 0;
 2 h¼h0
>
> dG
>
: 6 0:
dh2 h¼h0

Numerous numerical techniques can be used to compute G. The most commonly used methods are briefly
described in the following.

3.3.1. Real crack extension


Based on the definition of G, this method consists in computing the total potential energy Wp for the
initial crack length a, and for the increased crack length (a þ da). As the strain energy release rate is the
decrease of the total potential energy Wp , G can be approximated by
oWp Wp ða þ daÞ  Wp ðaÞ
G¼  ;
oA b da
where oA is the surface increment corresponding to the crack increase da, and b represents the thickness in
the out-of-plane direction. This technique, based on the physical definition of G, is easy to implement.
However it is expensive in terms of computational time since it requires a significant refinement at the crack
tip as well as two mechanical computations for each crack length increase.

3.3.2. Path independent J integral


Among all the path independent integral methods [27,28], the J integral method introduced by Rice [29]
is the most commonly used:
Z  
oui
J¼ wðeÞn1  rij nj ds;
C ox

where w is the elastic strain density, u is the displacement field at point M of path C of outward normal n. r
and e are respectively the stress and strain fields. Rice showed that J is path-independent in quasi-static
isothermal conditions, if no load is applied on the crack edges and for self-similar crack growth. Moreover,
J is equal to the strain energy release rate. In plasticity, this statement is only verified in experiments
without unloading. In practice, it is shown that the accuracy of the computation is path-dependent. The
most accurate results are obtained with paths far away from the crack tip [30].
3894 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

3.3.3. Virtual crack extension


Hellen [18] and Parks [31] suggest a computation of G by using its definition and a single crack length
computation. The crack is propagated by moving nodal points rather than by removing nodal tensions at
the crack tip and by performing a second analysis. The strain energy release rate can then be computed
using the rigidity matrix K, the displacement vector u and the load vector f of the numerical system
½Kfug ¼ ff g:

 
dWp 1 dK df
G¼ ¼  fugt fug þ fugt :
da 2 da da

3.3.4. Surface integral


De Lorenzi [32] introduces a method based on a continuum mechanics formulation of the virtual crack
extension formulation. The strain energy release rate can be computed as a surface integral––in 2D:
ZZ  
1 ouj oDx1
G¼J ¼ rij  wdi1 dA;
da A ox1 oxi
where A is the surface between paths C0 and C1 , and Dx1 represents the virtual crack extension. Numerical
applications of this method [28,33] show its great precision, and its independence to surface integration.

3.3.5. Gh method
In 1983, Destuynder [17] has introduced a new method based on a virtual displacement field h. The
strain energy release rate is the decrease in the total potential energy during a growth of area dA of the
crack. All derivatives with respect to the crack growth can be computed by using the Lagrangian method
[17,34].
Let X be a cracked solid (Fig. 6) and F d an infinitesimal geometrical perturbation d in the vicinity of the
crack tip:
F d : R3 ! R3
8M 2 X; F d ðMÞ ¼ M d ¼ M þ edðMÞ;
where the field h gives the location of each point of the perturbated solid using its initial position before
perturbation (Fig. 7).
If perturbation d is sufficiently small, Destuynder showed that the stress field r and the displacement field
u on the perturbed configuration may be expressed as

Γ1

Γ0

Fig. 6. Virtual extension method.


P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3895

Fig. 7. Cracked solid.

rd ¼ r þ dr1 ;
ud ¼ u þ du1 ;

where r1 and u1 are the first variations of the stress and displacement fields during the infinitesimal per-
turbation d on X. The total potential energy variation during a crack extension is then obtained when e aims
dW W d W
towards 0: dap ¼ limd!0 p d p .
The virtual displacement field h represents the virtual kinematics of the crack. This field has the following
properties:

• h is parallel to the crack plan (obvious in 2D);


• h is normal to the crack front;
• the support of h is only needed in the vicinity of the crack;
• khk is constant in a defined area around the crack tip.

In practice, we define two paths C1 and C2 around the crack tip. These paths divide the part into three
domains (Fig. 8).

Fig. 8. Contours and domains used to compute G with the Gh method.


3896 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

For example, the virtual displacement field hðh1 ; h2 Þ may be expressed as


8  
>
> IM
>
< h 1 ¼ 1  cosðhÞ;
IJ
 
>
> IM
> h2 ¼ 1 
: sinðhÞ;
IJ
where O is the crack tip, M is an integration point, I and J are the intersections between OM and C1 and C2
respectively.
Thus the values of h in the three domains are:

• in Cint , the norm of h is constant and equal to 1;


• in Cext , h is 0;
• in the ring Cring , the norm of h varies continuously from 1 to 0.

Then, if there is neither thermal strain nor load applied directly to the crack, the total potential energy may
be expressed as
Z Z
1
Wp ¼ TrðrrU Þ dX  fU dX;
2 X X
where r and U are respectively the stress and displacement field, and f the external loads. If an infinitesimal
perturbation d is performed, derivatives and integrals on the perturbed part may be expressed using a first
order ‘‘limited development’’ of operations related to the non-perturbed part.
This technique enables to express the strain energy release rate as
Z Z
G¼ TrðrrU rhÞ dX  w divðhÞ dX;
X X

where w is the strain energy density.


For an elastic material, we obtain
Z Z
1
G¼ TrðrrU rhÞ dX  TrðrrU Þ divðhÞ dX:
X 2 X
Besides, as h varies only inside Cring , the integration may be performed only over Cring . The strain energy
release rate computation is then performed by integration over the seven integration points of each element
of the ring:
Z

1
G¼ TrðrrU rhÞ  TrðrrU Þ divðhÞ dC;
ring 2

X X
G¼ ðr11 u1;1 þ r12 u1;2 Þh1;1 þ ðr11 u2;1 þ r12 u2;2 Þh1;2 þ ðr12 u1;1 þ r22 u1;2 Þh2;1
elements int p

1
þ ðr12 u2;1 þ r22 u2;2 Þh2;2  ½ðr11 u1;1 þ r12 u1;2 þ r12 u2;1 þ r22 u2;2 Þðh1;1 þ h2;2 Þ wint dAreaint ;
2
where wint is the Gauss weight of the integration point int and dAreaint the associated area.
All these fields are computed in a finite element analysis, so that it is possible to evaluate G accurately,
using a single mesh and a single mechanical computation.
Comparisons to other techniques [26] show that the Gh method is very accurate and completely mesh
independent.
The Gh method has been implemented in FORGE2. At each crack increment, GðhÞ is computed for h
varying from )70 to +70 with degrees steps of 1, 5 and 10. The GðhÞ curve is increasing and then
P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3897

G
(N.m-1)

-70˚ θ0 +70˚ θ
Fig. 9. GðhÞ curve for the maximum strain energy release rate criterion.

decreasing (Fig. 9), so that the determination of the angle h0 corresponding to the maximum strain energy
release rate is straightforward.
The Gh method can also be used for elastic–plastic materials. However, it is then restricted to stationary
cracks. Indeed, the computation of G for elastic–plastic materials is based on an analogy between an
elastic–plastic behaviour and a non-linear elastic behaviour. This analogy is only possible without any
global unloading. A crack propagation itself represents a local unloading of the material, therefore G is
only available for a unique crack extension with an elastic–plastic behaviour.

4. Applications

The maximum circumferential stress criterion (MCSC), the minimum strain energy density criterion
(MSEDC), and the maximum strain energy release rate criterion (MSERRC)––using the Gh method––have
been implemented in FORGE2 and tests have been performed successfully.
In this section, we compare the different crack trajectories for various applications. In each example, the
material is purely elastic with a Young modulus E ¼ 98,000 MPa, and a Poisson ratio m ¼ 0:3.

4.1. Rectangular part with an oblique pre-crack

A rectangular part with an oblique crack is submitted to a vertical tensile test (Fig. 10).

4.1.1. Maximum circumferential stress criterion


The maximum circumferential stress criterion is used to compute the direction of the crack propagation
at each time step. As expected, the crack propagates in the cleavage mode (mode I), perpendicularly to the
maximal principal stress which is vertical due to the applied load (Fig. 11).
Fig. 12 shows the shape of the equivalent stress (MPa) field at the crack tip. It is similar to the theoretical
shape in plane strain.

4.1.2. Minimum strain energy density criterion


In this criterion, the accuracy is directly dependent on the number of elements in the ring around the
crack tip. The strain energy density is computed for each element of this ring and the local minimum is then
evaluated using the SðhÞ curve.
3898 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

Fig. 10. Rectangular part with an oblique pre-crack.

Fig. 11. Crack trajectory with the MCSC.

Fig. 12. Comparison between the theoretical maximum stress zone (a) and the numerical simulation (b).
P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3899

However, there is a slight difference between values from external elements of the ring (ext in Fig. 13a)
and values from internal elements (int in Fig. 13a). This difference makes the computation of the local
minimum difficult. In this case, it is recommended to separate values for external and internal elements (Fig.
13b).
The computation of the local minimum may be improved by taking the local minimum and the values in
the two neighbour elements. The local minimum is then computed as the minimum of the parabola fitting
these three points.
As previously observed, the crack propagates horizontally (Fig. 14b). Besides it is possible to visualise the
strain energy during propagation (Fig. 14a). It is concentrated around the crack tip, and in the direction for
which this energy is minimum.

Fig. 13. (a) Ring elements and (b) calculated SðhÞ curve for the MSEDC with the numerical formulation.

Fig. 14. (a) Strain energy field and (b) crack propagation predicted by the MSEDC.
3900 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

Fig. 15. Crack propagation with the MSEDC without numerical improvements.

Fig. 16. Crack trajectories calculated with the MSERRC for an angle scan of (a) 10, (b) 1 and (c) superimposed trajectories.

If the technique of the parabola is not used, and if all elements of the ring are taken into account, the
same criterion appears to be rather inaccurate (Fig. 15). The global direction of propagation is the ap-
propriate one, but the crack trajectory is perturbed. This is not important in this example, but gives raise to
wrong trajectories for more complex parts or loading conditions.

4.1.3. Maximum strain energy release rate criterion


In this criterion, the direction of propagation is computed from the maximum of the GðhÞ curve. The
accuracy depends on the angle step chosen. In Fig. 16, the crack trajectories are compared for a step of 10
(Fig. 16a) and a step of 1 (Fig. 16b).
The global trajectories are similar. But the crack direction obtained with 10 steps is not really horizontal
as expected: it comes from the fact that the only way for the crack to propagate horizontally is to bifurcate
at 10, which is big in this test case. This shows that a slight difference in the first load step may involve
important differences in the final trajectory. This remark is also valid for other criteria that could be mesh-
dependent. This is why it is crucial that the implementation of the MCSC be mesh-independent.
However, the GðhÞ curve (Fig. 17) is far more regular than the SðhÞ curve, and it is then easier to compute
accurately the maximum. The parabola technique may also be used, and gives good results even with a 10
scanning.

4.1.4. Discussion
The three criteria give good results on the previous simple example. However, we have noticed that
energy criteria need a ring of elements around the crack tip. The number of elements and the radius of
P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3901

Fig. 17. GðhÞ curve.

Fig. 18. Pre-cracked part with two holes.

this ring, as well as the accuracy of scanning––for the Gh––are critical when calculating the crack trajec-
tory.
In the following example, we compare the criteria on a more complex problem with two cracks and two
holes, and particular attention is paid to the mesh refinement and the CPU time required to calculate a
complete crack propagation.

4.2. Pre-cracked part with two holes

In [14], an example of a crack propagation in a planar domain with an off-centre hole has been stud-
ied, and it is shown that the crack is moving towards the hole since it creates a stress drop. Once the crack
tip has moved beyond the hole, the crack reorients horizontally in the mode I loading. We consider here a
pre-cracked part with two holes (Fig. 18). When dealing with multiple cracks, one has to compute inde-
pendently the velocity of each crack. Some authors [35] suggest to propagate cracks one after the other,
according to the value of their strain energy release rate or stress intensity factors. In the following example,
the two cracks can be propagated with the same length since they are symmetric.
3902 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

Fig. 19. Crack trajectory in a pre-cracked part with two holes, calculated with the MCSC.

4.2.1. Maximum circumferential stress criterion


At the beginning, each crack is growing towards the nearest hole (Fig. 19a). Then, the cracks reorient
horizontally (Fig. 19b) since they modify the stress distribution at each otherÕs tip (Fig. 19c). At the end,
they are attracted again by the opposite hole (Fig. 19d). In Fig. 19e, the influence of each crack on the other
is presented in terms of the Von Mises equivalent stress field.

4.2.2. Minimum strain energy density criterion


In this example, three simulations are performed with a small ring around the crack tip (Fig. 20a), a
medium one (Fig. 20b) and a large one (Fig. 20c). For these three cases, the number of elements in the mesh
is the same. Fig. 20 shows that results are different in each configuration.
When a small radius is used, the crack moves increasingly towards the hole and it does not reorient
horizontally. For a medium radius, the global trajectory is the proper one but it is locally perturbed. Finally
a large radius leads to a good computation of the trajectory. Indeed, mechanical fields are perturbed in the
close vicinity of the crack tip. The SðhÞ curve (Fig. 20d) is so much perturbed that the minimum compu-
tation is approximate. When the computation is performed further from the crack tip, the results are more
accurate.

4.2.3. Maximum strain energy release rate criterion


This criterion has also been applied with three different radii for the ring of elements. However, as shown
in Fig. 17, the GðhÞ curve is extremely regular. This is still true for a little radius, and the maximum of G is
P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3903

Fig. 20. Crack trajectory calculated with the MSEDC with different ring radii.

easily evaluated, even for a 10 scanning. In Fig. 21, the superposition of the trajectories for the three cases
show the mesh-independence of this method.

4.2.4. Discussion
This more complex example shows that the MSEDC is mesh-dependent whereas the MCSC and the
MSERRC give good results without any significant mesh dependence. The crack trajectories obtained with
the MSEDC and MSERRC (Fig. 22a), and with the MSERRC and the MCSC (Fig. 22b) are compared.
Energy criteria give the same global trajectory. The crack path is more perturbed trajectory for the
MSEDC. The superposition in Fig. 22b shows that the local criterion (MCSC) appears to be more influ-
enced by holes, but no significant difference in the global trajectory is observed.
The CPU time computation is also an important aspect in the choice of numerical methods. Conse-
quently the CPU times spent for the crack propagation with the three criteria, and with the same automatic

Fig. 21. Superposition of crack trajectories calculated with the MSERRC, for three different mesh refinements.
3904 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

Fig. 22. Crack trajectories comparison.

Table 1
CPU time comparison
Criterion MCSC MSEDC MSERRC
Number of nodes 3053 3075 3186
CPU time t 1.3t 1.08t

remeshing parameters are compared. The results in Table 1 show that the MSEDC is slower than the two
others methods that are time equivalent.

5. Multimaterial applications

Previous results have shown the robustness of our technique to propagate cracks in one material.
However it is interesting to use the multimaterial capabilities of FORGE2 to study the crack propagation in
a part containing several materials. In this section, two examples are presented in order to demonstrate the
influence of bimaterial rigidity on crack paths.

5.1. Crack growth from a fillet

This example, performed experimentally by Sumi, shows the growth of a crack from a fillet in a structural
member. The crack propagation depends on the welding residual stresses and the bending stiffness of the
structure. For sake of simplicity, residual stresses are not taken into account. The bending stiffness of the
structure is modified by varying the size of the bottom I-beam presented in Fig. 23 from 15 to 315 mm with
an intermediate value of 115 mm.
A linear elastic plane strain simulation is performed with a Young modulus and a Poisson ratio of
E ¼ 200 GPa, and m ¼ 0:3 respectively. The initial crack length is a0 ¼ 5 mm, and the MCSC is used to
compute the crack path.
Numerical simulations in Fig. 24 show the important influence of the bottom I-beam rigidity on the crack
path.
The same example has been studied by Fleming el al. [36] using the element-free Galerkin method in-
troduced by Belytschko [37]. Crack paths obtained with this method and the results of our discrete crack
propagation technique are compared on a part containing a flexible I-beam. Fig. 25 shows the excellent
correlation of the two techniques.
P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3905

Fig. 23. Crack growth from a fillet.

Fig. 24. Influence of the bottom I-beam rigidity on the crack path.

Fig. 25. Comparison of crack paths calculated with the element free Galerkin method [36] and with our discrete crack propagation
method.

5.2. Pre-cracked part with an inclusion

In [14] we have studied the propagation of a crack in a planar part with an off-center hole. It was shown
that the crack was attracted by the hole since it creates a stress fall in this region.
3906 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

In the present example, the hole is replaced by an inclusion, and the influence of this inclusion on the
crack path is studied. A rectangular part is pre-cracked and submitted to a tensile test. This part contains an
inclusion which may be more rigid or less rigid than the matrix. R is defined as the ratio of the matrix and
inclusion Young modulus: R ¼ Ematrix =Eincl .
The numerical simulation is performed in plane strain, and the MSERRC is used.
Fig. 26 shows that for an inclusion less rigid than the matrix, the crack is attracted to the inclusion. The
crack reorientation is however less pronounced than the one obtained with a hole.
Conversely, if the inclusion is more rigid than the matrix (Fig. 27), the crack is moving away from the
inclusion.

Fig. 26. Propagation of a crack in a part containing a soft inclusion R ¼ 10.

Fig. 27. Propagation of a crack in a part containing a hard inclusion R ¼ 0:1.


P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908 3907

6. Conclusion

An advanced remeshing technique is used to propagate cracks through finite element meshes. Every step
in the multiple applications presented above is completely automatic. It shows the robustness and the
accuracy of the coupled numerical techniques associated with the discrete crack propagation.
Three crack kinking criteria have been implemented and compared. All these comparisons show that the
MSEDC is less accurate than the two other criteria. The MCSC and MSERRC are quite equivalent in
terms of accuracy and time computation. The MCSC appears to be the easiest to implement in any finite
element code. However it requires a refined mesh at the crack tip. The MSERRC is the most complex but
gives good results. It requires a ring of elements around the crack tip, but its accuracy is mesh-independent.
It should be interesting to study the validity and accuracy of each of these two criteria on elastic–plastic
examples.
The multimaterial characteristics of the finite element formulation allows to study various applications,
such as the propagation into composites. In the future, crack propagation should be modelled at interfaces
between different materials.

References

[1] C.E. Inglis, Stresses in a plate due to the presence of cracks and sharp corners, Proc. Inst. Naval Architects 60 (1913) 219–241.
[2] A.A. Griffith, The phenomena of rupture and flow in solid, Phil. Trans. Roy. Soc. London A 221 (1920) 163–197.
[3] G.R. Irwin, D.C. Washington, Analysis of stresses and strains near the end of a crack traversing a plate, J. Appl. Mech. (1957)
361–364.
[4] T. Belytschko, Y.Y. Lu, L. Gu, Element free Galerkin methods, Int. J. Numer. Methods Engrg. 37 (1994) 229–256.
[5] J.C. Simo, J. Oliver, F. Armero, An analysis of strong discontinuities induced by strain softening in rate dependent inelastic solids,
Int. J. Numer. Methods Engrg. 12 (1993) 277–296.
[6] K. Garikipati, On strong discontinuities in inelastic solids and their numerical simulations, Ph.D. Thesis, Stanford University,
1996.
[7] J. Oliver, Continuum material failure in strong discontinuity settings, in: E. O~ nate, D.R. Owen (Eds.), COMPLAS 2003, VII
International Conference on Computational Plasticity, CIMNE, Barcelona, 2003.
[8] T. Belytschko, T. Black, Elastic crack growth in finite elements with minimal remeshing, Int. J. Numer. Methods Engrg. 45 (1999)
601–620.
[9] N. Mo€es, J. Dolbow, T. Belytschko, A finite element method for crack growth without remeshing, Int. J. Numer. Methods Engrg.
46 (1999) 131–150.
[10] N. Mo€es, E. Bechet, Modeling stationary and evolving discontinuities with finite elements, in: E. O~ nate, D.R. Owen (Eds.),
COMPLAS 2003, VII International Conference on Computational Plasticity, CIMNE, Barcelona, 2003.
[11] I. Babuska, J.M. Melenk, The partition of unity method, Int. J. Numer. Methods Engrg. 40 (1997) 727–758.
[12] T. Bittencourt, P. Wawrzynek, J. Sousa, A. Ingraffea, Quasi-automatic simulation of crack propagation for 2D LEFM problems,
Engrg. Fract. Mech. 55 (2) (1996) 321–334.
[13] B.J. Carter, P.A. Wawrzynek, A.R. Ingraffea, Automated 3D crack growth simulation, Gallagher Special Issue of Int. J. Numer.
Methods Engrg. 47 (2000) 229–253.
[14] P.O. Bouchard, F. Bay, Y. Chastel, I. Tovena, Crack propagation modelling using an advanced remeshing technique, Comput.
Methods Appl. Mech. Engrg. 189 (2000) 723–742.
[15] Y. Chastel, C. Magny, F. Bay, An elastic–viscoplastic finite element model for multimaterials, Engrg. Comput. 15 (1) (1998) 139–
149.
[16] R.S. Barsoum, On the use of isoparametric finite elements in linear fracture mechanics, Int. J. Numer. Methods Engrg. 10 (1976)
25–37.
[17] P. Destuynder, M. Djaoua, S. Lescure, Quelques remarques sur la mecanique de la rupture elastique, J. Meca. Theo. Appl. 2 (1)
(1983) 113–135.
[18] T.K. Hellen, On the method of virtual crack extentions, Int. J. Numer. Methods. Engrg. 9 (1975) 187–207.
[19] F. Erdogan, G.C. Sih, On the crack extension in plane loading and transverse shear, J. Basic Engrg. 85 (1963) 519–527.
[20] G.C. Sih, B. Macdonald, Fracture mechanics applied to engineering problems––strain energy density fracture criterion, Engrg.
Fract. Mech. 6 (1974) 361–386.
3908 P.O. Bouchard et al. / Comput. Methods Appl. Mech. Engrg. 192 (2003) 3887–3908

[21] 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. Fract. 24 (1984) 5–22.
[22] M.A. Hussain, S.L. Pu, J.H. Underwood, Strain energy release rate for a crack under combined Mode I and Mode II, Fract.
Analysis, ASTM STP 560, Philadelphia, 1974, pp. 2–28.
[23] K. K€ onke, G. Schmid, in: Meskouris 1 Wittek (Ed.), Prediction of Crack Propagation Directions under Plane Normal and Shear
Loading, Aspects in Modern Comp. Struct. Analysis, Balkema Press, Rotterdam, 1997, pp. 431–441.
[24] J.C.W. van Vroonhoven, R. de Borst, Combination of fracture and damage mechanics for numerical failure analysis, Int. J. Solids
and Structures 36 (1999) 1169–1191.
[25] C. K€ onke, Damage evolution in ductile materials: from micro- to macro-damage, Comp. Mech. 15 (1995) 497–510.
[26] P.O. Bouchard, Contribution a la modelisation numerique en mecanique de la rupture et structures multimateriaux, These de
doctorat de lÕEcole Nationale Superieure des Mines de Paris, Septembre 2000.
[27] H.B. Bui, Dualite entre les integrales de contour, Compte Rendu Acad. Sciences, vol. 276, Paris, Mai 1973.
[28] X.B. Zhang, Etude numerique de la propagation de fissures par la mecanique de la rupture, These de doctorat de lÕuniversitede
Clermont-Ferrand II, Juin 1992.
[29] J.R. Rice, A path independent integral and the approximate analysis of strain concentrations by notches and cracks, J. Appl.
Mech. 35 (1968) 379–386.
[30] F.Z. Li, C.F. Shih, A. Needleman, A comparison of methods for calculating energy release rate, Engrg. Fract. Mech. 21 (2) (1985)
405–421.
[31] D.M. Parks, A stiffness derivative finite element technique for determination of crack tip stress intensity factors, Int. J. Fract. 10
(4) (1974) 487–502.
[32] H.G. de Lorenzi, Energy release rate calculations by the finite element method, Engng Fract. Mech. 21 (1) (1985) 129–143.
[33] S.C. Lin, J.F. Abel, Variational approach for a new direct-integration form of the virtual crack extension method, Int. J. Fract. 38
(1988) 217–235.
[34] Ph. Gilles, Ph. Mourgue, M. Rochette, Precision de calcul de la force dÕextension de fissure G: effets du maillage et avantage de la
methode G-h, in: Hermes et INRIA (Ed.), Acte du Colloque National en Calcul de Structures, vol. 2, 1993, pp. 639–670.
[35] J. Wang, Development and application of a micromechanics-based numerical approach for the study of crack propagation in
concrete, Ph.D. Thesis from Ecole Polytechnique Federale de Lausanne, 1994.
[36] M. Flemming, Y.A. Chu, B. Moran, T. Belytschko, Enriched element-free Galerkin methods for crack tip fields, Int. J. Numer.
Methods Engrg. 40 (1997) 1483–1504.
[37] T. Belytschko, Y.Y. Lu, L. Gu, Element Free Galerkin methods, Int. J. Numer. Methods Engrg. 37 (1994) 229–256.

You might also like