Вы находитесь на странице: 1из 11

Finite Elements in Analysis and Design 45 (2009) 675-- 685

Contents lists available at ScienceDirect


Finite Elements in Analysis and Design
journal homepage: www. el sevi er . com/ l ocat e/ f i nel
Predicting the influence of overload and loading mode on fatigue crack growth:
Anumerical approach using irreversible cohesive elements
Haodan Jiang, Xiaosheng Gao

, T.S. Srivatsan
Department of Mechanical Engineering, The University of Akron, Akron, OH 44325, USA
A R T I C L E I N F O A B S T R A C T
Article history:
Received 12 November 2008
Received in revised form 20 May 2009
Accepted 20 May 2009
Available online 21 June 2009
Keywords:
Irreversible cohesive element
Fatigue crack growth
Damage accumulation
Overload
Loading modes
Compact-tension-shear specimen
This paper describes an irreversible cohesive zone model to simulate fatigue crack growth. The model
includes a three-dimensional (3-D) cohesive law that follows a distinct loading/unloading path and a
damage evolution mechanism that reflects a gradual degradation of cohesive properties of the material
under the influence of cyclic loading. To overcome convergence difficulties arising from nonlinearity of
the cohesive zone model, the stabilization technique and the viscous regularization of the constitutive law
are employed. For high-cycle fatigue applications, a damage extrapolation scheme is adopted to reduce
computational cost. The irreversible cohesive zone model is implemented in the finite element software
ABAQUS through a user defined subroutine and is used to predict fatigue crack growth in a compact-
tension-shear (CTS) specimen with an emphasis on the extrinsic influence of overload for different loading
modes. The numerical results show good agreement with experimental records documented in the open
literature and capture the essential features of fatigue crack growth for various loading conditions. This
indicates that the irreversible cohesive zone model can serve both as an accurate and efficient tool for
the prediction of fatigue crack growth.
2009 Elsevier B.V. All rights reserved.
1. Introduction
It is fairly well known that fatigue can be considered as a ma-
jor cause for inducing failure in engineering structures. Up until
now prediction of fatigue life remained very much an empirical art.
The most universally used method to predict fatigue crack growth
was put forth by Paris and co-workers [1], which relates the fatigue
crack growth rate (FCGR) to the applied stress intensity factor range
(K) through a power law relationship. However, Paris' law only
successfully describes the experimental data under restricted condi-
tions, such as, small scale yielding, constant amplitude loading and
long cracks and loses its predictive ability when these requirements
are not adhered to [2,3]. Since the 1960s, several modifications of
the Paris law have been proposed and put forth to enable improve-
ment in the prediction for cases that are different from the ideal
conditions. These cases include the following: (i) R-ratio effects [4],
(ii) threshold limits [5], (iii) crack closure [6], (iv) variable ampli-
tude loads [7], (v) small cracks [8,9], and others. On the other hand,
the J-integral range (J) and the crack tip opening displacement
range (CTOD) were also suggested for use in the Paris law under

Corresponding author. Tel.: +13309722415; fax: +13309726027.


E-mail address: xgao@uakron.edu (X. Gao).
0168-874X/$ - see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.finel.2009.05.006
conditions of gross plasticity [10,11]. Despite the widespread use of
the Paris law(including the modified forms), it is worth noting that it
provides a data correction scheme rather than a predictive capability
primarily because it does not capture the physics of fatigue crack
growth [3,12].
Mechanism-based concepts provide a key insight to support de-
velopment of computational models for the prediction of local frac-
ture initiation and continued propagation. The concept of cohesive
zone, initially put forth by Dugdale [13], Barrenblatt [14], Rice [15]
and others, considers fracture to be a gradual phenomenon in which
separation occurs between two adjacent virtual surfaces across an
extended crack tip (cohesive zone) and is resisted by the cohesive
forces. This theory of fracture leads to a new numerical approach
to simulate crack propagation. In this approach, the continuum is
characterized by two constitutive relations: (i) a volumetric consti-
tutive model describing the bulk behavior of the material and (ii) a
cohesive surface constitutive relation between the traction and dis-
placement characterizing the behavior of bond surfaces between the
elements (Fig. 1).
The loss of cohesion and thus crack formation or extension may
be viewed as a progressive decay of otherwise intact tension and
shear stresses across the adjacent surfaces. The introduction of non-
linear, interface constitutive laws specified between the tractions
and displacements across the surfaces provides a phenomenological
description for the progressive damage and eventual fracture of the
676 H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685
volumetric element
cohesive element
Fig. 1. Schematic representation of the concept of cohesive/volumetric finite ele-
ments. The triangular elements are separated for illustrative purpose: the cohesive
elements have no thickness and the adjacent nodes are superposed prior to the
onset of cohesive failure.
material. In essence, each finite element is tied to its neighbors
though a series of nonlinear springs which governs its interaction
with the surrounding. The creation of an internal surface is associated
with failure of a number of these springs. Such cohesive models
define an inherent toughness for the material, thereby, introducing
an intrinsic length-scale in the local fracture process, which enables
the fracture process zone(s) on the specimen/component scale to
gradually evolve as a natural outcome of the computation. Also, the
intrinsic length-scale enables converged solutions to be obtained
with increased mesh refinement, i.e., the solution does not depend
on size of the element to provide the length-scale over which the
separation processes occur.
Many variations of the cohesive zone models have been
proposed and successfully applied for the prediction of crack prop-
agation under monotonic loading [1619]. However, under fatigue
loading conditions, the material deteriorates with time and the
prevailing cohesive strength must be related to the loading history.
Nguyen et al. [20] revealed that a cracked specimen subjected to
constant-amplitude cyclic loading tends to shakedown when using
the non-dissipative unloadingreloading cohesive model to simu-
late fatigue crack propagation, i.e., after a small number of cycles,
the material undergoes an elastic cycle of deformation followed by
crack arrest. To overcome this problem, they introduced an irre-
versible cohesive law with unloadingreloading hysteresis where
the loading and unloading incremental stiffness are different. Roe
and Siegmund [21] introduced a damage parameter into the co-
hesive law using an effective stress concept of damage mechanics
[22]. This irreversible cohesive zone model has been successfully
applied for the prediction of various interface fatigue crack growth
problems [21,23,24].
The 3-D cohesive zone model presented in this paper takes on an
exponential form of tractionseparation relations. The irreversibility
is incorporated into the constitutive law with considerations of: (a)
specific loading and unloading paths; (b) accumulation of damage
under subcritical cyclic loading; and (c) compression or normal sur-
face contact behavior. The current cohesive tractions are determined
by the current amount of damage as well as by the current sepa-
ration. The constitutive laws are therefore history dependent and a
gradual degradation of the cohesive properties under cyclic loading
is reflected in the process zone ahead of the crack tip. This irre-
versible cohesive zone model is implemented in the finite element
software ABAQUS using a user defined subroutine and its features
are demonstrated through simple examples.
The service loading of engineering structures and components
is usually far from the simple case of constant amplitude loading.
Overloads come up constantly in the loading history and affect the
fatigue life to a variable extent. While cracks evolved under fa-
tigue loading displays complicated characteristics, the fatigue crack
propagation under variable amplitude loading interspersed with
overload and for different loading modes is of special interest.
Sander and Richard [2527] conducted an extensive experimental
and numerical study to investigate the different modes of overload
on fatigue crack growth of aluminum alloy 7075-T651. To validate
our irreversible cohesive zone model, the compact-tension-shear
(CTS) specimens tested by Sander and Richard was used. The numer-
ical model captures the essential features of fatigue crack growth
and the effects of overload for the different loading modes.
2. An irreversible cohesive zone model
The fracture process in engineering structures and components
can be viewed as a progressive decay of material strength across
two adjacent virtual surfaces. The cohesive tractions, induced by
the inter-atomic forces, between the virtual incipient crack surfaces,
work as the resistance to crack propagation. Under external load-
ing conditions, the distance between the atoms changes and the
inter-atomic traction decreases once a characteristic length scale is
reached. When the interfacial tractions diminish, the incipient vir-
tual surfaces separate from each other, defining the onset or forma-
tion of a macroscopic crack. The cohesive zone model is developed
to reflect the material separation process phenomenologically.
In the form of the principle of virtual work, the mechanical equi-
librium equation considering the contribution of the cohesive trac-
tions is written as
_
V
s : o

F dV
_
Sint

T o

dS =
_
Sext

T
ext
o udS (1)
where V, S
int
and S
ext
represent the specimen volume, the internal
(cohesive) surface and the external surface, respectively, s,

T and

T
ext
denote the nominal stress tensor, the cohesive traction vector
and the external traction vector, respectively,

F is the deformation
gradient and u is the displacement vector. The relative displacement
vector

A= u
+
u

represents the difference between two adjacent


cohesive surfaces. The contribution of the cohesive zone is described
by the integration term over the internal surface.
The traction vector

T acting on the cohesive surfaces can be de-
rived from an interfacial potential, or the free energy density poten-
tial [(

A) [28]

T =
*[(

A)
*

(2)
Under 3-D settings,

A and

T each has three components (normal,
tangential and transverse): (u
n
, u
t1
, u
t2
) and (T
n
, T
t1
, T
t2
). The
key features of a cohesive zone model include the following:
(1) the shape of the tractionseparation curve;
(2) the cohesive strength, defined by the peak value of the
tractionseparation curve;
(3) the cohesive energy density, defined by the area under the
tractionseparation curve; or the characteristic separation cor-
responding to the cohesive strength.
Different potential forms, mainly divided between polynomial
and exponential forms, have been adopted in the past by some
researchers [18,20,28,29]. The present work chooses the computa-
tionally convenient exponential form of the free energy density po-
tential.
[=o
max,0
eo
0
_
1.0 +
u
n
o
0
_
exp
_

u
n
o
0
_

_
(1.0 q) + q exp
_

u
2
t1
+u
2
t2
o
0
__
(3)
H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685 677
where o
max,0
is the initial normal cohesive strength under monotonic
loading. o
0
is the characteristic cohesive length, i.e. the material
separation required to reach the cohesive strength in normal loading.
q is the ratio between the normal cohesive energy density and shear
cohesive energy density under monotonic loading conditions.
With isotropic assumptions, the present work defines the initial
shear cohesive strengths (both tangential and transverse) under pure
shear conditions to be:
t
1 max,0
=t
2 max,0
=

2eqo
max,0
(4)
In this study, we assume t
1 max,0
= t
2 max,0
= o
max,0
, which corre-
sponds to q = 1/

2e 0.4289. The cohesive energy density, or the


work of separation per unit area of cohesive surface, is given by
I
c
=
_

0
T d (5)
With the cohesive law given by (3), I
c
= eo
max,0
o
0
.
2.1. Monotonic loading
We start by considering the monotonic loading situation. Accord-
ing to (2), the cohesive traction components can be computed from
derivatives of the potential function as
T
n
=o
max,0
u exp(1 u){(1 q) + q exp(v
2
1
v
2
2
)} (6a)
T
t1
=2qo
max,0
(1 +u)v
1
exp(1 u) exp(v
2
1
v
2
2
) (6b)
T
t2
=2qo
max,0
(1 +u)v
2
exp(1 u) exp(v
2
1
v
2
2
) (6c)
with u =u
n
/o
0
, v
1
=u
t1
/o
0
and v
2
=u
t2
/o
0
, respectively.
The features of the tractionseparation constitutive equations de-
fined above are shown in Fig. 2. Under pure normal loading, the
cohesive traction increases with increasing separation before the
characteristic cohesive length o
0
is reached, and decreases subse-
quently until it approaches the zero traction line. Similarly, under
pure shear (tangential/transverse) loading, the shear cohesive trac-
tions increase with increasing shear separations before the value
of

2o
0
/2 is reached, and decreases subsequently to approach the
zero shear traction line. The adopted form of the tractionseparation
model reduces the traction value exponentially to zero with increas-
ing separation. In order to provide a consistent definition for the
amount of crack extension, the advancing crack tip is defined at
u
n
=5o
0
or |u
ti
| =3o
0
.
2.2. Unloading/reloading
To account for fatigue crack propagation under cyclic loading,
the paths of unloading and reloading need to be considered in the
1
1
u
n
/
0
1
-1
1
-1
u
ti
/
0
T
ti
/
i max,0
T
n
/
max, 0
Fig. 2. Monotonic tractionseparation relation under: (a) pure normal loading and
(b) pure shear loading (tangential: i = 1 and transverse: i = 2).
cohesive zone model. When the separation of the corresponding co-
hesive surfaces becomes smaller than the previous value, unloading
occurs. In this research endeavors, unloading is prescribed to follow
the path back to the origin of the tractionseparation space (Fig. 3).
T
n
=u
n
k
n,0
, T
t1
=u
t1
k
t1,0
, T
t2
=u
t2
k
t2,0
k
n,0
= T
n,max
/u
n,max
, k
t,0
= T
t1,max
/u
t1,max
k
t2,0
= T
t2,max
/u
t2,max
(7)
where u
n,max
, u
t1,max
and u
t2,max
are values of the separation
components at the time when unloading starts, and T
n,max
, T
t1,max
and T
t2,max
are the corresponding traction values. When the cur-
rent separation overpasses the previous value but is smaller than
the maximum separation, the traction separation relation follows a
reloading path. Under monotonic loading, the reloading progresses
along the adverse direction of the unloading path (Fig. 3).
2.3. Compression/normal contact of broken cohesive surfaces
During unloading in the normal direction, the unloading path will
reach u
n
= 0 and the crack surfaces came into contact with each
other. In order to avoid overlapping and inter-penetration of the
material surfaces, a penalty is taken for the cohesive traction corre-
sponding with u
n
<0. The penalized equation for contact compu-
tation is given by the expression
T
n
= Ao
max,0
u exp(1 u) (8)
The stiffness multiplier, A, strongly penalizes any negative value of
u and thus prevents he solid elements surrounding the cohesive
element from interpenetrating into each other. In the present study,
a value of A = 30 is used.
2.4. Damage evolution
For the condition of cyclic loading, the constitutive equations
must be modified with a damage law that will result in failure of
the cohesive zone after cycling the element at subcritical loads. The
damage law takes several factors into consideration and these in-
clude the following:
(1) Damage should not begin to accumulate until the accumulated
separation is greater than a critical distance, o
0
.
(2) The increment of damage should be based on the increment of
separation.
(3) There should be an endurance limit for damage accumulation
based on the traction level, under which level the cyclic loading
can precede without causing failure.
Adamage variable Dis evaluated within the proposed constitutive
relations, defined as the ratio between the damaged cross-sectional
area and the initial cross-sectional area. Considering the essential
factors of the damage law, a damage evolution equation suggested by
Roe and Siegmund [21] is adopted. The derivative of damage under
cyclic loading,

D
c
, is defined as

D
c
=

u
4o
0
_

T
o
max
C
f
_
H( u
tot
o
0
) (9)
and the current amount of damage due to cyclic loading
D
c
=
_
t

D
c
dt (10)
where

T =
_
T
2
n
+(T
2
t1
+ T
2
t2
)/(2eq
2
) is the effective traction, u =
_
u
2
n
+(u
2
t1
+u
2
t2
)/(2eq
2
) is the effective separation for 3-D situ-
ation, u
tot
=

t
| u| is the accumulated separation at that instant.
678 H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685
1
-1
1
-1
u
t1
/
0
T
t1
/
1max, 0
compression
1
1
T
n
/
max, 0
u
n
/
0
Fig. 3. (a) Normal and (b) shear unloading/reloading paths.
With the Heaviside step function H( u
tot
o
0
) involved, the dam-
age variable will not start to evolve until the accumulated separation
at that instant is greater than the characteristic cohesive length o
0
.
The fatigue limit coefficient, C
f
, is defined as the ratio of the
uniaxial fatigue limit for a zero mean stress fatigue test over the
initial cohesive strength
C
f
=
o
f
o
max,0
(11)
This coefficient ensures the existence of a fatigue endurance limit.
A value of C
f
=2.5 is used in the present study.
In order to encompass failure of the element under the mono-
tonic loading condition, a damage mechanism for monotonic load-
ing is proposed. The amount of damage due to monotonic loading is
calculated as
D
m
=
u
max
o
0
4o
0
H( u
max
o
0
) (12)
The greater value of the two damage mechanisms is stored as the
current amount of damage
D =max(D
c
, D
m
), 0D1 (13)
With the incorporated damage evolution mechanism, modifica-
tions are made in the constitutive equations of the tractionseparation
relation. The cohesive strengths (normal, tangential and transverse)
degrade with increased damage accumulation
o
max
=o
max,0
(1 D), t
1 max
=t
1 max,0
(1 D)
t
2 max
=t
2 max,0
(1 D) (14)
Assuming the material to be isotropic, the replacement of the
strength term o
max,0
in Eq. (6) with the current strength o
max
achieves the purpose of reflecting the degradation of cohesive
strengths in the 3-D situation.
3. Finite element implementation and convergence concerns
The irreversible cohesive zone model discussed above is
implemented in the finite element software ABAQUS through a user
defined subroutine. For 3-D finite element models, the cohesive ele-
ment type COH3D8 is employed. The 3-D cohesive element has eight
nodes and four integration points as shown in Fig. 4. For the purpose
of predicting fatigue crack growth, cohesive elements with initially
zero thickness are defined along the crack path. With cyclic loading,
the upper and lower surface of the cohesive elements separate from
each other and damage is accumulated within the element. When
1
8 7
6 5
4
3
2
1 2
3
4
Initially zero
thickness
1
2
3
Integration points
Nodes
8-node element
Fig. 4. The COH3D8 cohesive element.
damage reaches the limit value of 1, the element is considered to be
broken at the current integration point.
The cyclic plasticity is taken into account with the nonlinear kine-
matic hardening model of Lemaitre and Chaboche [30], which is
implemented in ABAQUS. The evolution law of this model consists
of two components: a nonlinear kinematic hardening component,
which describes the translation of the yield surface in stress space
through the backstress, :, and an isotropic hardening component,
which describes a change of the equivalent stress defining the size
of the yield surface, o
0
, as a function of plastic deformation [31]. The
kinematic hardening component is defined to be an additive com-
bination of a purely kinematic term and a relaxation term, which
introduces the nonlinearity
a = C
1
o
0
(r a)

c
p
a

c
p
(15)
where C is the initial kinematic hardening modulus and deter-
mines the rate at which the kinematic hardening modulus decreases
with increasing plastic deformation. For aluminum alloy 7075-T651
tested by Sander and Richard [2527], these material parameters are
C = 9393 and = 34.96.
3.1. Viscous regularization and stabilization
Material models exhibiting softening behavior and stiffness
degradation often lead to severe convergence difficulties in implicit
finite element analysis. The cohesive zone model implemented is
distinctive of material softening after the characteristic separation is
reached and practical applications in ABAQUS/STANDARD often turn
out with convergence problems. A common technique to overcome
some of these convergence issues is the use of viscous regularization
of the constitutive equations, which causes the tangent stiffness
H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685 679
matrix of the softening material to be positive for sufficiently small
time increments [31].
The tractionseparation laws can be regularized using vis-
cosity by permitting stresses to be outside the limits set by the
tractionseparation law. The regularization process involves the use
of a viscous stiffness degradation variable, D
v
, which is defined by
the evolution equation

D
v
=
1
j
(D D
v
) =
dD
v
dt
(16)
where j is the viscosity parameter representing the relaxation time
of the viscous system and D is the damage variable evaluated in the
current cohesive zone model.
For incremental finite element calculations, the current viscous
stiffness degradation variable is calculated as
D
v
(t + dt) =
_
1 +
dt
j
_

_
D
v
(t) +
dt
j
D(t + dt)
_
(17)
When this viscous regularization technique is used, the damage term
in Eq. (14) is replaced with the current viscous stiffness degradation
variable D
v
. With a small value of the viscosity parameter, the vis-
cous regularization helps improve the rate of convergence without
compromising results. A value of j =10
5
is used in this study and
provides reliable results.
Top surface
Bottom surface
volumetric element
cohesive element
volumetric element
-1
-0.5
0
0.5
1
0
T
n

/

m
a
x
,
0
u
n
/
0
compression
unloading/reloading
loading
1 2 3 4 5 6 7
Fig. 5. (a) A three-element model containing two soild elements and one cohesive element and (b) normal tractionseparation relation with contact behavior.
T
t
i

/

m
a
x
,
0
u
ti
/
0
-1
-0.5
0
0.5
1
-0.015
u
n
/
0
0
0.2
0.4
0.6
0.8
1
0 0.02 0.04 0.06 0.08 0.1 -0.005 0.005 0.015
T
n
/

m
a
x
,
0
Fig. 6. The cyclic behavior of the cohesive element for: (a) pure normal (with R = 0) and (b) pure shear/transverse (with R = 1) under load-controlled cyclic loading.
To further improve the stability of the nonlinear finite element
analysis, the automatic stabilization technique provided in ABQUS
/STANDARD using a constant damping factor for static problems is
adopted [31]. The value of the damping factor is adjusted for different
geometry and loading cases to ensure minimum negative effects
related with the induced dissipated energy.
4. Numerical results
4.1. Features of the irreversible cohesive zone model and damage
extrapolation scheme for high cycle fatigue applications
A three-element finite element model is generated to examine
the loading/unloading and contact behaviors of the cohesive zone
model, where the cohesive element is sandwiched by two volumet-
ric solid elements as shown in Fig. 5(a). The boundary conditons of
the model are defined by restraining the 3 DOFs (degree of freedom)
of the bottom surface. A normal cyclic loading with a load ratio
R = 1 is applied to the top surface nodes in a way that the cohesive
element is loaded, unloaded and compressed before being com-
pletely damaged. Fig. 5(b) shows the normal tractionseparation
curve of the cohesive element under monotonic loading and con-
tact/compression. An exponential form is shown for the monotonic
loading, unloading/reloading is directed toward/from the origin and
negative separation is penalized to avoid material penetration under
680 H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685
0
0.2
0.4
0.6
0.8
0
-1
-0.5
0
0.5
1
-0.5
u
n
/
0
T
t
i

/

m
a
x
,
0
u
ti
/
0
T
n

/

m
a
x
,
0
0.1 0.2 0.3 0.4
-0.25 0 0.25 0.5
Fig. 7. The cyclic behavior of the cohesive element for: (a) pure normal (with R = 0) and (b) pure shear/transverse (with R = 1) under displacement-controlled cyclic loading.
contact or compression. With a monotonic shear/transverse loading,
the cohesive element presents similar exponential relations as pro-
posed (Fig. 2).
The general features of the irreversible cohesive model are tested
using a simple, one-element model for verification. Consider a co-
hesive element having square top- and bottom-surfaces. The nodes
on the bottom-surface are fixed. A cyclic load/displacement is im-
posed in the normal or shear (tangential/transverse) directions, for
pure normal or pure shear loading cases, respectively, on the nodes
at the top-surface, with the two other DOFs constrained. The cyclic
tractionseparation response of the cohesive element under constant
amplitude cyclic loading is shown in Fig. 6 for the load-controlled
loading and Fig. 7 for the displacement-controlled loading. For each
loading condition, damage accumulation due to cyclic loading de-
grades the cohesive strength, reflected in a continuous reduction in
the stiffness of the unloading/reloading path. As the traction value
as well as the unloading/reloading stiffness decreases, the cohesive
zone finally loses its load carrying capacity and failure of the mate-
rial occurs.
For high-cycle fatigue applications, the iterative damage calcu-
lations for crack propagation can be time-consuming or even com-
putationally prohibitive. In order to reduce the computation cost, a
damage extrapolation scheme is employed in the proposed cohe-
sive zone model. The extrapolation scheme extrapolates the damage
variable for the amount of damage at higher cycles after calcula-
tions of relatively small number of loading cycles. With reasonably
small extrapolation spans, a linear extrapolation scheme based on
first order Taylor expansion [32] of the damage variable can accom-
plish the goal within acceptable accuracy levels. The equation used
for the extrapolation of the damage variable D is as follows:
D
n+1
D
n
+
*D
*N

n
(N
n+1
N
n
) (18)
We compute the rate of change of D per cycle at N
n
, (*D/*N)|
n
,
required for extrapolation by a detailed step-by-step computation.
A linear approximation is used based on the damage change in the
range of a few previous loading cycles. Also to be noted, the cycle
increment (N
n+1
N
n
) used in the extrapolation formula is selected so
that the damage increment [D
n+1
D
n
] is sufficiently small. Following
the extrapolation of D, an equilibrium loop is entered which has
the effect of updating the remaining state variables in a manner
consistent with the constitutive relations.
0
0.02
0.04
0.06
0.08
0.1
0
t
D
without extrapolation
with extrapolation
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Fig. 8. Damage extrapolation scheme applied with the cohesive zone model. (100
analyzed loading cycles within the time period [0,1]).
The effect of the damage extrapolation scheme with the result
of the first cohesive element at the crack tip of a CTS specimen is
shown in Fig. 8. As demonstrated by the thinner line in the figure
the damage variable D is extrapolated 10 cycles after 15 normal
loading cycles (at the time t = 0.15) with the results abstracted from
cycles 10 to 15 (t [0.1,0.15]); and again after calculations of 25 total
loading cycles (at the time t = 0.25) with the results from cycles 20
to 25 (t [0.2,0.25]). The result of damage accumulation without the
extrapolation scheme (the thicker line) is also presented in Fig. 8
for comparison. As we can see in this case, the extrapolated damage
value on the thinner line coincide with the damage value after the
corresponding number of cycles plus the extrapolated cycles on the
thinner line, e.g. at time t = 0.15 (15 cycles plus extrapolated 10
cycles), the damage on the thinner line coincide with that on the
thicker line at time t = 0.25 (25 cycles). Although the accumulated
error may become manifest with the process of calculations going
on, considering the computational efficiency required for high cycle
H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685 681
5
4
t=10
1
8
0.9
M4
w=90
42.5
27 27 18
a
17
5
4
1
4
5

0
15
30 60
75
90
F
F

Fig. 9. CTS-specimen: (a) specimen dimensions (length unit: mm); and (b) loading device.
fatigue, the extrapolation scheme is useful with careful choice of
extrapolation spans.
4.2. Finite element modeling of fatigue crack growth in CTS specimens
Sander and Richard [2527] conducted extensive experimental
and numerical studies to investigate the influence of overload for the
different loading modes on fatigue crack growth of high strength alu-
minum alloy 7075-T651. The compact-tension-shear specimen (CTS)
developed by Richard [33] was used in these studies. Fig. 9 shows
the dimensions of the specimen and the designed loading device.
In this study one layer of 3-D elements with the plane strain con-
straints are used to model the specimen. Depending on the loading
angle :, the loading mode varies from pure Mode I (: = 0

) to pure
Mode II (: = 90

). The stress intensity factors K


I
and K
II
for the CTS-
specimen can be calculated from the following equations developed
by Richard [33]:
K
I
=
F
wt

a
cos :
1 a/w
_
0.26 +2.65a/(w a)
1 +0.55a/(w a) 0.08(a/(w a))
2
(19)
K
II
=
F
wt

a
sin :
1 a/w
_
0.23 +1.40a/(w a)
1 +0.67a/(w a) +2.08(a/(w a))
2
(20)
In the finite element model, six concentrated forces are applied to
the six holes as shown in Fig. 10(a). Depending on the loading angle
:, the corresponding force components from F
1
to F
6
are calculated
using the following equations developed by Richard [33], according
to the force F applied on the loading device
F
1
= F
6
= F
_
1
2
cos : +
c
b
sin :
_
, F
2
= F
5
= F sin :
F
3
= F
4
= F
_
1
2
cos :
c
b
sin :
_
(21)
The numerical analyses reported in this section were performed
under plane strain condition. For this purpose, one layer of eight-
node, 3-D linear hybrid brick elements with constant pressure
(C3D8H in the element library of ABAQUS) is used to model the CTS
specimen. Along the predefined crack path at the centerline, a row of
F1
F2
F3
F4
F5
F6
cohesive elements
Fig. 10. (a) Simplified loading for the finite element model; (b) in-plane finite
element mesh; and (c) close-up of the finite element mesh at the crack tip region.
cohesive elements (COH3D8) with zero initial thickness was laid in
between the surrounding solid elements (C3D8H). The size of each
cohesive element along the crack path is 0.025mm. Fig. 10(b) shows
682 H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685
Fig. 11. Definition of the characteristic parameters of an overload between constant
amplitude loading.
Table 1
Mechanical and fracture properties of the aluminum alloy 7075-T651.
Young's modulus, E 70,656MPa
0.2%-Yield strength, Rp0,2 517MPa
Tensile strength, Rm 579MPa
Poison's ratio, v 0.34
Fracture toughness, KIC 32.95MPa m
1/2
Threshold value, Kth (R = 0.1) 3.15MPa m
1/2
the in-plane finite element mesh and Fig. 10(c) shows the close-up
of the mesh at the crack tip region. All the nodal displacements are
constrained in the z-direction to satisfy the plane strain condition.
The simulations were carried out initially for Mode I baseline
loading with a constant amplitude of K
Bl
=7MPa m
1/2
and a stress
ratio R
Bl
=0.1, so that a steady crack growth of the preliminary state
is obtained. After this, a single Mode I, or Mode II overload with a
constant overload ratio of R
ol
is followed. Finally, the same baseline
loading is added following the overload. The loading sequence is
shown in Fig. 11.
The overload ratio R
ol
is defined as
R
ol
=
K
eq,ol
K
Bl,max
(22)
The equivalent stress intensity factor K
eq,ol
= K
I,ol
for Mode I, and
K
eq,ol
= :
1
K
II,ol
for Mode II. The parameter :
1
= 1.155 suggested by
Sander and Richard [2527] is used for this study.
The CTS specimens were made from aluminum alloy 7075 in
the T651 temper in the TL direction. The mechanical and fracture
properties of the alloy are given in Table 1 [25].
The cohesive energy density, I
c
= eo
max,0
o
0
, establishes the
material toughness when the background material follows a linear-
elastic response. The fracture toughness K
IC
is usually measured
from high-constraint specimens where plastic dissipation is neg-
ligible during crack initiation. Therefore, I
c
can be approximately
calculated from I
c
K
2
IC
(1 v
2
)/E under plane strain condition.
Using the material properties given in Table 1, the corresponding
cohesive energy density is calculated to be I
c
=13.59kJ/m
2
.
With the value of I
c
obtained, the cohesive law described by
Eq. (3) will be established once the value of o
max,0
is determined.
This is usually done by comparing the computed load versus load-
line displacement response or fracture resistance curve (JR curve)
with experimental records for a given fracture specimen. The fitting
process entails several finite element crack growth analysis of the
reference specimen using different o
max,0
-values. The o
max,0
-value
that provides the best agreement between the numerical results and
the experimental values will be used [18]. Due to lack of adequate
fracture data, the calibration process of o
max,0
is not performed here.
-300
0
300
600
900
1200
1500
0
R
ol
=2
R
ol
=1.5
R
ol
=3
R
ol
=2.5
Baseline loading

y

(
M
P
a
)

distance from the crack tip (mm)
0.5 1 1.5
Fig. 12. Change of the crack opening stress due to an overload of different load ratio.
Instead, the value of o
max,0
is chosen based on the results of several
previous studies [18,19,28] and a series of trial and error analyses.
According to the previous parameter studies, the cohesive strength
may vary in the range of (23.5) times of the yield strength. A higher
cohesive strength leads to lower crack propagation rate and vise
versa. To best match the experimental data, a value o
max,0
equal to
three times the yield strength of the material is used here.
4.3. Effects of the overload and loading modes
The change in stress along the crack path for a single Mode I
overload is shown in Fig. 12, in direct comparison with the baseline
level loading. The stress increase due to an overload is greater for
the higher overload ratio.
As a direct result of the overload-induced plasticity, stress redis-
tribution takes place in the solid elements around the crack tip. The
stress redistribution in the CTS model at the valley and peak cyclic
loading points after a Mode I overload, in comparison to the stress
state of the base line loading is shown in Fig. 13. At the valley load-
ing point following an overload (Fig. 13(a)), a compressive residual
stress develops, which causes either partial or complete crack clo-
sure. The compressive stress is manifest for a larger overload ra-
tio. At the peak loading point following the overload (Fig. 13(b)),
the peak stress that could be reached is reduced compared to the
baseline case, indicating a retardation effect on crack growth by the
overload. A higher overload ratio leads to a greater retardation ef-
fect. With continued cyclic loading following the overload, and the
crack tip advancing forward due to fatigue crack growth, the stress
distribution ahead of the crack tip gradually recovers to the state of
the baseline level loading.
With the same overload ratio of R
ol
=2.5, the level of compressive
residual stress caused by a single Mode II (: = 90

) overload is lower
than that followed after a Mode I overload (Fig. 14). Consequently,
the retardation on crack advance following a Mode II overload is
much smaller than the Mode I overload.
The fatigue crack growth rate (FCGR) can be obtained from nu-
merical analysis. A stable crack growth is observed for constant am-
plitude base line loading (Fig. 15). The nearly perfect straight line of
the variation of a with N is consistent with the Paris law. The slope
of the straight line indicates a constant crack growth rate of about
910
5
mm/cycle.
With the application of a single cycle of overload, the damage
accumulation in the cohesive elements is greatly accelerated during
H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685 683
-300
-200
-100
0
100
0
-100
0
100
200
300
400
500
600
R
ol
=3
R
ol
=2.5
R
ol
=2
R
ol
=1.5
Baseline
R
ol
=3
R
ol
=2.5
R
ol
=2
R
ol
=1.5
Baseline
distance from the crack tip (mm) distance from the crack tip (mm)

y

(
M
P
a
)

y

(
M
P
a
)
0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Fig. 13. Stress distribution at the: (a) valley loading point and (b) peak loading point after four cycles of baseline loading subsequent to the overload.
-200
-150
-100
-50
0
50
100
0
-100
0
100
200
300
400
500
600
0
baseline
Mode I
Mode II
baseline
Mode I
Mode II

y

(
M
P
a
)

y

(
M
P
a
)

distance from the crack tip (mm) distance from the crack tip (mm)
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
Fig. 14. Comparison of the stress distribution at the: (a) valley loading point and (b) peak loading point after a cycle of the Modes I and II overload.
0
0.05
0.1
0.15
0.2
0.25
0
N (cycle)

a

(
m
m
)
1000 2000 3000 4000 5000
Fig. 15. Stable crack growth under the baseline loading.
an overload. Due to the relatively large element size with respect
to the short application time of the overload, the complete failure
of one element at the crack tip is not observed during an overload.
However, the crack growth rate during an overload can be estimated
based on damage accumulation in the element at the existing crack
tip. Our numerical result suggests that crack growth rate during
an overload is approximately 710
3
mm/cycle, which is noticeably
higher than the crack growth rate under the baseline loading.
Due to plasticity-induced retardation, the crack growth rate is
decreased immediately following an overload (Fig. 16). The crack
growth rate will decrease to a minimum value and then will grad-
ually increase to the value under baseline loading as the crack tip
moves out of the region affected by the compressive residual stress
caused by the overload. The overload-induced crack retardation in-
creases the computation time exponentially. Based on available com-
putational facility, we were only able to grow the crack for a small
number of elements following an overload (in Fig. 16, the minimum
crack growth rate has not been reached).
5. Concluding remarks
An irreversible cohesive zone model incorporated with a damage
evolution mechanism is described in this paper. The monotonic and
irreversible cyclic tractionseparation (normal/shear) relations of the
model are verified via a simple three-element model. The gradual
degradation of cohesive properties under cyclic loading is effectively
captured with the damage mechanism of the proposed model. For
high-cycle fatigue applications, the damage extrapolation scheme
adopted greatly reduces the computation cost while maintaining
reasonable accuracy.
The CTS specimens tested by Sander and Richard [2527] are
analyzed with a row of cohesive elements putting along the crack
684 H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685
1.0E-07
1.0E-06
1.0E-05
1.0E-04
1.0E-03
1.0E-02
0
distance from the crack tip (mm)
d
a
/
d
N

(
m
m
/
c
y
c
l
e
)
0.05 0.1 0.15 0.2 0.25 0.3
Fig. 16. Prediction of the crack growth rate for the baseline loading interspersed with a single Mode I overload.
growth path. The numerical results show the s tress redistribution at
the crack tip after a single overload interspersed in a constant ampli-
tude loading. The compressive residual stress caused by the overload
lead to the decrease of fatigue crack propagation. The retardation
effect is less obvious for lower overload ratio. The retardation effect
is also dependent of the mode of the overload: it is less for Mode II
overload than that for Mode I overload.
With the irreversible cohesive zone model, the fatigue crack
growth rate becomes a natural result of the numerical analysis.
The predicted crack growth rates agree with the values reported
by Sander and Richard [2527]. Although the cohesive element de-
veloped in this study is fully three-dimensional, only plane strain
analyses are reported in this paper. Fully three-dimensional analy-
sis is currently underway and the results will be reported in future
publications. In this study, only one row of cohesive elements are
placed directly ahead of the original crack front, which allows crack
growth in the original direction only. For mixed mode loading, the
actual crack growth may deviate from the original crack plane. This
can be simulated by putting cohesive elements along every solid
element boundaries of an arbitrary or quasi-arbitrary mesh, and the
deviating crack path will come out as the natural results of different
loading situations. This phenomenon will also be considered in the
future work.
Acknowledgment
The authors would like to thank Dr. M. Sander of University of
Paderborn, Germany for providing material properties used in the
finite element analysis.
References
[1] P. Paris, F.A. Erdogan, Critical analysis of crack propagation laws, Journal of
Basic Engineering 85 (1963) 528534.
[2] M. Klesnil, P. Lukas, Influence of strength and stress history on growth and
stabilization of fatigue cracks, Engineering Fracture Mechanics 4 (1972) 7792.
[3] S. Suresh, Fatigue of Materials, second ed., Cambridge University Press,
Cambridge, UK, 1998.
[4] O. Wheeler, Spectrum loading and crack growth, Journal of Basic Engineering
94 (1972) 181186.
[5] C. Laird, Mechanisms and Theories of Fatigue, Fatigue and Microstructure,
American Society for Metals, Metals Park, OH, 1979 pp. 149203.
[6] R. Foreman, V. Keary, R. Engle, Numeral analysis of crack propagation in cyclic-
loaded structures, Journal of Basic Engineering 89 (1967) 459464.
[7] J. Willenborg, R. Engle, R. Wood, Technical Report AFFDL-TM-71-1-FBR, Air
Force Flight Dynamics Laboratory, 1971.
[8] W. Elber, Fatigue crack closure under cyclic tension, Engineering Fracture
Mechanics 2 (1970) 3745.
[9] M. El Haddad, N. Dowling, T. Topper, K. Smith, J-integral applications for short
fatigue cracks at notches, International Journal of Fracture 16 (1980) 1530.
[10] N. Dowling, J. Begley, in: Mechanics of Crack Growth, ASTM STP 590, American
Society for Testing and Materials, Philadelphia, PA, 1976, pp. 83104.
[11] P. Neumann, The geometry of slip processes at a propagating fatigue crack-II,
Acta Metallurgica 22 (1974) 11671178.
[12] J.P. Bailon, S.D. Antolovich, Effect of microstructure on fatigue crack propagation:
a review of existing models and suggestions for further research, in: J. Lankford,
D.L. Davidson, W.L. Morris, R.P. Wie (Eds.), Fatigue Mechanisms: Advances
in Quantitative Measurements of Physical Damage ASTM STP 811, American
Society for Testing and Materials, Philadelphia, PA, 1983, pp. 313349.
[13] D.S. Dugdale, Yielding of steel sheets containing slits, Journal of the Mechanics
and Physics of Solids 8 (1960) 100104.
[14] G.I. Barrenblatt, The mathematical theory of equilibrium of cracks in brittle
fracture, Advances in Applied Mechanics 7 (1962) 55129.
[15] J.R. Rice, Mathematical analysis in the mechanics of fracture, in: H. Liebowitz
(Ed.), FractureAn Advanced Treaties, vol. 2, Academic Press, New York, London,
1968, pp. 191311.
[16] X. Xu, A. Needleman, Numerical simulations of fast crack growth in brittle
solids, Journal of the Mechanics and Physics of Solids 42 (1994) 13971434.
[17] G. Camacho, M. Ortiz, Computational modeling of impact damage in brittle
materials, International Journal of Solids and Structures 33 (1996) 28992938.
[18] S. Roychowdhury, Y.A. Roy, R.H. Dodds, Ductile tearing in thin aluminum panels:
experiments and analyses using large-displacement, 3-D surface cohesive
elements, Engineering Fracture Mechanics 69 (2002) 9831002.
[19] W. Li, T. Siegmund, An analysis of crack growth in thin sheet metal via a
cohesive zone model, Engineering Fracture Mechanics 69 (2002) 20732093.
[20] O. Nguyen, E.A. Repetto, M. Ortiz, R.A. Radovitzky, A cohesive model of fatigue
crack growth, International Journal of Fracture 110 (2001) 351369.
[21] K.L. Roe, T. Siegmund, An irreversible cohesive zone model for interface fatigue
crack growth simulation, Engineering Fracture Mechanics 70 (2003) 209232.
[22] J. Lemaitre, A Course on Damage Mechanics, Springer, Berlin, 1996.
[23] B. Wang, T. Siegmund, A numerical analysis of constraint effects in fatigue crack
growth by use of an irreversible cohesive zone model, International Journal of
Fracture 132 (2005) 175196.
[24] B. Wang, T. Siegmund, Simulation of fatigue crack growth at plastically
mismatched bi-material interfaces, International Journal of Plasticity 22 (2006)
15861609.
[25] M. Sander, H.A. Richard, Finite element analysis of fatigue crack growth with
interspersed mode I and mixed mode overloads, International Journal of Fatigue
27 (2005) 905913.
[26] M. Sander, H.A. Richard, Experimental and numerical investigations on the
influence of the loading direction on the fatigue crack growth, International
Journal of Fatigue 28 (2006) 583591.
[27] M. Sander, H.A. Richard, Fatigue crack growth under variable amplitude
loading part II: analytical and numerical investigations, Fatigue and Fracture
of Engineering Materials and Structures 29 (2006) 303319.
H. Jiang et al. / Finite Elements in Analysis and Design 45 (2009) 675-- 685 685
[28] A. Needleman, Micromechanical modeling of interfacial decohesion,
Ultramicroscopy 40 (1992) 203214.
[29] M. Ortiz, A. Pandolfi, Finite-deformation irreversible cohesive elements
for three-dimensional crack propagation analysis, International Journal for
Numerical Methods in Engineering 44 (1999) 12671282.
[30] J. Lemaitre, J. Chaboche, Mechanics of Solid Materials, Cambridge University
Press, Cambridge, 1994.
[31] ABAQUS/Standard User's Manual, Version 6.3, Hibbit, Karlsson and Sorensen
Inc., 2002.
[32] A. De-Andres, J.L. Perez, M. Ortiz, Elastoplastic finite element analysis of three-
dimensional fatigue crack growth in aluminum shafts subjected to axial loading,
International Journal of solids and structures 36 (1999) 22312258.
[33] H.A. Richard, Fracture Mechanical Predictions for Cracks with Superimposed
Normal and Shear Loading, VDI-Verlag, Dusseldorf, 1985 (in German).

Вам также может понравиться