Bosch2008 Article OnTheDevelopmentOfA3DCohesiveZ

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Comput Mech (2008) 42:171–180

DOI 10.1007/s00466-007-0184-8

ORIGINAL PAPER

On the development of a 3D cohesive zone element in the presence


of large deformations
M. J. van den Bosch · P. J. G. Schreurs ·
M. G. D. Geers

Received: 13 October 2006 / Accepted: 11 April 2007 / Published online: 15 May 2007
© Springer-Verlag 2007

Abstract Delamination is typically modelled using cohe- possible because there are no emissions of volatile organic
sive zone models. In this paper, it is shown that the validity of compounds and lacquering is superfluous [5,10].
most models is limited to small displacements and/or defor- The polymer coating can delaminate because it is sub-
mations at the interface and the surrounding bulk materials. jected to the same deformation as the metal substrate. This
A large displacement formulation (LDF) is proposed that causes the loss of protective and attractive properties of the
overcomes issues of the classical formulation with large dis- product [3] and is therefore unacceptable. Adequate material
placements and deformations at the interface. Subsequently, description and process simulations are required to predict
a 3D cohesive zone element with this LDF is introduced and delamination, which allows the adjustment of process param-
its numerical implementation is elaborated. Then, a 3D FEM eters to prevent it.
model is proposed and the determination of several model In-situ observation of the delamination process of poly-
parameters is substantiated. Finally, the implementation is mer coated steel reveals the presence of large deformations
validated by comparing numerical results with experimental and displacements in both the interface and the bulk mate-
observations. rials. The delamination occurs by the initiation, growth and
fracture of fibrils at the interface, a process called fibrillation.
Keywords Cohesive zone · Large displacements · In the numerical tool, the delamination process is
Delamination · Laminates · Coatings described by cohesive zone models, which represent a rela-
tion between the traction and the relative displacement of
associated points of crack surfaces. Cohesive zones are typi-
1 Introduction cally implemented in finite element codes as interface
elements.
Polymer coated metals are used nowadays for an extend- Most (2D) cohesive zone models provide a constitutive
ing variety of products, e.g., beverage cans and aerosols. A relation between the normal opening displacement and the
polymer coating is applied to the metallic substrate prior normal traction and a separate relation between the tangen-
to forming a product. Polymer coated metals have several tial opening displacement and the tangential traction (see
advantages over uncoated metals. The polymer coating pro- a.o. [4,7,18] for an overview). The tractions are sometimes
tects the substrate from corrosion, without influencing the coupled by making them a function of both the normal and
taste of the beverage or food. Moreover, costs savings are tangential opening displacement [17,21]. The normal and
tangential direction are defined with respect to a local coor-
M. J. van den Bosch
dinate basis that is placed inside the cohesive zone element.
Netherlands Institute for Metals Research (NIMR),
PO Box 5008, 2600 GA Delft, The Netherlands This approach is accurate when there are small displacements
present at the interface, e.g., when the cohesive zones are used
M. J. van den Bosch · P. J. G. Schreurs · M. G. D. Geers (B) to simulate the indentation-induced delamination of a brit-
Department of Mechanical Engineering,
tle film [1]. Therefore these models are next referred to as
Eindhoven University of Technology, PO Box 513,
5600 MB Eindhoven, The Netherlands cohesive zone models with a small displacement formulation
e-mail: m.g.d.geers@tue.nl (SDF). However, in the presence of large displacements at

123
172 Comput Mech (2008) 42:171–180

the interface, these models, as shown later, may give unex-


pected results. To overcome the problems of the SDF, a large
displacement formulation (LDF) for fibrillation is proposed.
Large displacement cohesive zone models have already
been developed [14,15]. However, these cohesive zone
models were used to predict crack growth with large dis-
placements in the bulk material but the displacements at the
interface prior to failure were relatively small and should
therefore be considered as SDF models.
In this paper first some experimental results on the delam-
ination of the polymer coating are shown, illustrating the
formation of fibrils. Secondly, the SDF and its problems to
describe large displacements unambiguously are discussed. Fig. 2 SEM micrograph of the delamination front and the presence of
fibrils
Subsequently, the LDF is introduced and it is shown how it
handles large displacements correctly. Then, the formulation
of a 3D cohesive zone element with a LDF is elaborated. constitutive relations (2D/3D), that describe the relations
Finally, the constitutive model for the coating is introduced, between the tractions and the opening displacements. Here a
on the basis of which a 3D peel model is built and compared reversible exponential cohesive zone law is used [18]:
with experiments.       
φt ∆t ∆n ∆n ∆2t
Tt = 2 1+ exp − exp − 2 (1)
2 Experimental observation of fibrillation δt δt δn δn δt
and
The substrate is a batch annealed steel. It is extrusion coated      
φn ∆n ∆n ∆2
on one side with poly-ethylene terephthalate (PET). The sub- Tn = exp − exp − 2t (2)
strate is 210 µm thick and the PET layer has a thickness of δn δn δn δt
30 µm. where T represents a traction, φ a work-of-separation, δ a
Peel test were performed in-situ inside a scanning electron characteristic cohesive length and ∆ an opening displace-
microscope (SEM) as shown in Fig. 1 [19]. At the delami- ment. The subscripts ( )t and ( )n refer to the tangential and
nation front, the formation, growth and debonding of fibrils normal directions with respect to a local basis.
was observed (see Fig. 2). This process is called fibrillation Figure 3 shows a quadrilateral cohesive zone element with
and is considered to be the main mechanism by which the four nodal points and two integration points. The opening dis-
coating delaminates from the substrate under the imposed placements are determined by decomposing the total open-
loading conditions. ing displacement with respect to a local orthonormal basis
{et , en }. The local basis is commonly defined with respect
3 Small displacement formulation to a reference line. In most cases this is the cohesive zone
mid-line, which is line A–B in Fig. 3.
Most cohesive zone models in literature use a SDF to describe If there are small displacements at the interface, all ref-
quasi-brittle fracture. These models may have two or three erence lines within a cohesive zone element are parallel to
each other, i.e., the choice for a specific reference line will

Fig. 3 Schematical representation of the initial and deformed geom-


etry of a planar cohesive zone element with large displacements and
Fig. 1 In-situ delamination setup three possible orientations of a local basis

123
Comput Mech (2008) 42:171–180 173

1.5 Line 1−2


Line A−B
Line 4−3

1
∆n / l [−]
0

0.5 Fig. 5 An interface between two materials with an opening displace-


ment ∆ between two points that coincided initially

0
0.2 0.4 0.6 0.8 1 only transfer a load along their axes. A LDF is therefore pro-
t / τ [−] posed to resolve the ambiguity induced by the choice of a
local basis, where no distinction will be made between nor-
Fig. 4 The calculated normal opening displacement (normalized by
the initial cohesive zone length 0 ) of integration point 1 (ip 1 in Fig. 3)
mal and tangential loadings. Only a single constitutive rela-
for three different local bases as a function of the time t normalized by tion between the traction T and the opening displacement 
time τ at which the maximum opening is achieved is proposed:
T = f(). (3)
not significantly influence the magnitude of the normal and
The opening displacement is calculated between two points,
tangential opening displacement. However, when large dis-
on both sides of the interface, which coincided initially as
placements are present at the interface, the choice of a refer-
sketched in Fig. 5. The vectors are no longer decomposed
ence line may influence the orientation of the local basis and
with respect to a local basis but are resolved globally. There-
hence the decomposition of the total opening displacement
fore, the deficiencies of the SDF are overcome, since there
in a normal and tangential contribution [6]. To illustrate this,
is no need for a local basis. Adopting the geometrically
a single cohesive zone element is monotonically deformed
nonlinear solution procedure proposed in [19], tractions are
in a typical deformation mode shown in Fig. 3. The normal
resolved as Piola-Kirchhoff tractions. In here, the required
opening displacements, calculated with respect to the three
cohesive zone length is the initial length, which can be deter-
local bases, are shown in Fig. 4. From Fig. 4 it is clear that the
mined unambiguously.
normal opening displacement cannot be determined unam-
The constitutive relation (3) is next explicated on the basis
biguously due to the differences in orientation of the local
of the normal traction of the SDF cohesive zone law (Eq. 2).
bases.
It relates the traction T = T e to the opening displacement
So, the use of a local basis has several consequences when
 = ∆e , where e is the unit vector along the line between
large displacements occur at the interface. First, the orienta-
the associated points of the interface:
tion of the local basis will influence the magnitude of ∆t    
and ∆n and consequently the contribution of φt and φn to φ ∆ ∆
T = exp − (4)
the total dissipated energy. Secondly, in most SDF imple- δ δ δ
mentations the integration of the tractions is performed over where φ is the work-of-separation and δ characteristic open-
the length of the reference line. Upon large displacements in ing length, see Fig. 6. The maximum traction Tmax is related
one of the joined materials, the lengths of the reference lines to φ and δ by:
will change during deformation and as a result, the integrated
φ
tractions will have different magnitudes. Tmax = . (5)
Since delamination in polymer coated steel is character- δ exp(1)
ized by fibrillation and large displacements at the interface, As shown in Fig. 6, upon unloading, the cohesive zone law
a solution is required to remedy the inconsistencies that may can show a reversible response (i) or two types of irrevers-
result from the choice of the local basis. ible behaviour: linear elastic unloading to the origin, i.e.,
elasticity-based damage (ii) and unloading with the initial
stiffness of the cohesive zone, i.e., plastically damaged (iii).
4 Large displacement formulation All three cases are shown in Fig. 6.
It has been shown previously that the shape of the cohesive
If there are large displacements present at the interface, it zone law, even when a large deformation formulation is used,
is no longer physical to uniquely distinguish between a nor- has no significant influence on the macroscopic response of
mal and tangential opening. In the particular case of polymer the model [19]. In this paper, an exponential cohesive zone
coated steels, the interfacial gap is bridged by fibrils that can law is used because its traction and derivative are continuous

123
174 Comput Mech (2008) 42:171–180

0.8

0.6 (i)
[−]
max
T/T

0.4 (ii)

0.2
(iii) Fig. 7 An eight-noded 3D cohesive zone element with four integration
0 points
0 1 2 4 6 8
∆/δ [−]
A 3D cohesive zone element is shown in Fig. 7. Both the local
Fig. 6 Cohesive zone law of equation (4) with the maximum traction coordinates, ξ and η, span the range [–1 1]. In the deformed
∆ = δ and three types of unloading behaviour state a traction T connects associated (= with the same (ξ ,η)-
coordinate) points P and Q located on the faces S1 and S2 ,
respectively. Focussing attention to the cohesive zone ele-
in forward loading, which is attractive from a computational
ment, traction equilibrium across the interface involves the
point of view.
(Cauchy) traction vector σ rather than the full stress tensor
There are a number of advantages of a LDF cohesive
S. Integration over the cohesive zone element volume with
zone model compared to a SDF one. Firstly, there is no
area A yields:
local basis, rendering the decomposition of the total displace- 
ment superfluous. Therefore, the total work-of-separation is f i = ∆w · σ d A (8)
always equal to φ, independent of the mode-mixity. This is a
A
valid simplification of a mode-dependent cohesive zone law
if there are only fibrils at the interface. Secondly, the cohe- where the weighting function can be considered as a kine-
sive zone law of equation (4) can be derived from a poten- matically admissible virtual displacement. A pull-back to
tial function, which makes the dissipated energy intrinsically the initial undeformed state is performed, where the integra-
path-independent. Finally, the first Piola-Kirchhoff tractions tion takes place over the initial area A0 . Accordingly, Eq. (8)
are integrated in a consistent unambiguous manner by using becomes:
the initial cohesive zone element length.

fi = ∆w · T d A0 (9)
5 Three-dimensional cohesive zone element A0

where T is the first Piola-Kirchhoff traction vector. The


The cohesive zone is implemented in the commercial finite weighting function w and the traction T are a function of
element code MSC.MARC as a user element in a large dis- the local coordinate η. The integration is expressed in this
placement framework. The weak form of the weighted resid- coordinate, leading to:
ual integral of the equilibrium equation reads:
   1 1
A0
(∇w)c : S d V = w · q d V + w · t d A ∀ w(x) (6) fi = P T N T (ξ, η) T (ξ, η) dξ dη (10)
˜ 4 ˜
V V A −1 −1

where ∇ is the gradient operator and w arbitrary continuous where T (ξ, η) is the column with the tractions and P is an
weighting function. The boundary tractions at the external operator˜ matrix that transforms the nodal quantities to the
boundary A are given by t = S · n, whereas the body forces points A,B,C and D:
q apply to the volume V . The sum of the two integrals with the  
P = −I 12 I 12 (11)
external load vectors are denoted as the external force inte-
gral f e . The remaining integral on the left hand side is the in which I 12 is a 12 × 12 unity matrix. N (ξ, η) is the matrix
internal force integral f i , which expresses the virtual work with the interpolation functions:
of the internal stresses and the cohesive zone tractions. The ⎡ ⎤
ψ1 0 0 ψ2 0 0 ψ3 0 0 ψ4 0 0
weighted residual integral equation is then compactly written N (ξ, η) = ⎣ 0 ψ1 0 0 ψ2 0 0 ψ3 0 0 ψ4 0 ⎦
as: 0 0 ψ1 0 0 ψ2 0 0 ψ3 0 0 ψ4
f i (w, S) = f e (w, t, q) ∀ w. (7) (12)

123
Comput Mech (2008) 42:171–180 175

in which ψi are the four shape functions associated with the


points A, B, C and D:

1 1
ψ1 = (1 − ξ ) (1 − η) ψ2 = (1 + ξ ) (1 − η)
4 4
(13)
1 1
ψ3 = (1 + ξ ) (1 + η) ψ4 = (1 − ξ ) (1 + η).
4 4
Following a standard iterative procedure, similar to the classi-
cal 2D implementations, yields the approximate incremental Fig. 8 The normal vector n is determined in point M by the cross
value of the internal load integral and the element stiffness product of two tangent vectors
matrix:
The normal vector is given by the normalized cross product
A0
T T ip
4
f i∗
= P N (ξ, η) T ∗ (ξ, η) (14) of the two tangent vectors:
˜ 4 ˜
ip=1 tξ × tη
n= (18)
and ||tξ × tη ||
The magnitude of penetration parallel to the surface normal
A0
T T ip
4
n is ∆c =  · n, which leads to a contact penetration vector:
K∗ = P N (ξ, η) M(ξ, η) N (ξ, η) P (15)
4
ip=1 c = ∆c n = ( · n) n. (19)

where ( )∗ denotes the incremental value and M is the cohe-


sive tangent operator. Finally, the iterative element equation 5.2 Tractions
reads:
The adopted constitutive relation that relates the opening vec-

K δu = f e − f i∗ (16) tor  to its corresponding traction is given by the exponential
˜ ˜ ˜ cohesive zone law [19]:
where δu is a column with the iterative nodal displacements.  
˜ φ∆ ∆
T = exp − (20)
δ δ δ
5.1 Contact
where ∆ = |||| and φ and δ are the cohesive zone parame-
ters. The traction vector becomes:
A contact algorithm is required to detect the penetration of
surface S1 into S2 , or vice versa. The opening displacement T=Tm (21)
in the LDF is always positive (∆ = ||||) and can therefore
where m is a unit vector directed along . In the case con-
not be used to detect penetration. The approach presented
tact is present, i.e., ∆c < 0, an extra penetration term Tc is
here firstly determines a normal vector n to the middle sur-
present, directed along the normal of the plane:
face Sm of the cohesive zone. Then, it decomposes  to find
 2
the component parallel to n. If this component is negative, ∆c
there is penetration and a contact penalization is invoked. Tc = −kc and Tc = Tc n (22)
δ
This approach apparently introduces a local basis into the
where kc is the contact stiffness. The contact penetration term
formulation. However, this only affects the solution in the
is added to the constitutive traction resulting from the phys-
case of contact without significant influences on the solution
ical opening displacement . This yields the total traction
in relation to other local bases.
vector:
Since the surface Sm has four nodes, it can be curved and
there is no consistent normal vector. The normal vector is T = T + Tc . (23)
determined by taking the cross product of two tangent vec-
tors. The two tangent vectors, tξ and tη , are determined with 5.3 Tangent operator
respect to the isoparametric coordinate directions in the iso-
parametric mid-point M (ξ = 0, η = 0), as shown in Fig. 8. The consistent tangent stiffness tensor M relates infinitesimal
The tangent vectors are given by: variations of the traction vector to infinitesimal variations of
the opening displacement vector:
∂x ∂x
tξ = and tη = (17)
∂ξ M ∂η M δT = M · δ. (24)

123
176 Comput Mech (2008) 42:171–180

 
With respect to a Cartesian basis, e x , e y , ez , Eq. 24
becomes:
δT = M δ∆ (25)
˜ ˜
where components are stored in columns δT and δ∆ and
˜ ˜ the (a) (b)
matrix M. The consistent tangent stiffness matrix M for
LDF follows from a standard analysis of a 3D truss element Fig. 9 a Compressible Leonov model. b Schematic true stress–strain
under large displacements and rotations: curve of with decomposition of the intrinsic deformation behaviour
 
∂T ∗ T∗ T∗
M= m m T + ∗ m 2 m 2T + ∗ m 3 m 3T (26) where G r is the strain hardening modulus and B̃d is the devi-
∂∆ ˜ ˜ ∆ ˜ ˜ ∆ ˜ ˜
atoric part of the isochoric left Cauchy–Green deformation
where m , m 2 and m 3 are columns containing the components tensor. The driving stress is decomposed into a deviatoric
˜ ˜ m, m
of the vectors ˜ and m that constitute an orthonormal
2 3 stress Sds and a hydrostatic stress Ssh :
vector basis. For the exponential cohesive zone law, with
characteristic length δ, it can be verified that: Sds = G B̃de and Ssh = κ (J − 1) I (33)
 
T∗ δ ∂T ∗ where G is the shear modulus, B̃de the deviatoric part of the
= (27)
∆∗ δ − ∆∗ ∂∆ isochoric elastic left Cauchy–Green strain tensor, κ the bulk
modulus, J the volume change ratio, and I the unity tensor.
which leads to:
  The evolution of J and B̃de is given by the following kine-
∂T ∗ 1 matical equations:
M=
∂∆ ∆∗ (δ − ∆∗ )
⎡ ∗ ⎤ J˙ = J tr(D) (34)
δ∆ − ∆2x −∆x ∆ y −∆x ∆z
⎣ −∆x ∆ y δ∆∗ − ∆2y −∆ y ∆z ⎦ (28) and
−∆x ∆z −∆ y ∆z δ∆∗ − ∆2z ◦
B̃e = (Dd − D p ) · B̃e + B̃e · (Dd − D p ). (35)
where ∆i = ∗ · ei , with i the global x, y and z directions. ◦
The tangent operator for the contact penalization equals: Here B̃e is the objective Jaumann rate of B̃e and Dd the de-
 
∂ Tc ∗  T  viatoric part of the rate of deformation tensor. The plastic
Mc = nn . (29) deformation rate tensor D p is related to the deviatoric part of
∂∆c ˜˜
the driving stress Sds :
So, in the case of contact the total tangent operator becomes:
Sds
M = M + M c. (30) Dp = . (36)

The viscosity η was originally described by an Eyring rela-
tionship [8]. It was extended to incorporate pressure depen-
6 Constitutive modelling
dency and intrinsic strain softening in [9]:
 
A 3D constitutive model is used for description of large strain, µp τ̄ /τ0
η(T, p, τ̄ , S) = η0,r (T ) exp exp(S)
time dependent mechanical behaviour of the polymer layer. τ0 sinh(τ̄ /τ0 )
The constitutive model is a compressible generalization of (37)
the Leonov model [13], proposed by [2]. In the model a
where η0,r denotes the zero-viscosity for the completely reju-
distinction is made between the contribution of secondary
venated state, T is the temperature, µ is the material param-
interactions between polymer chains, which determine the
eter describing the pressure dependence and S is a parameter
(visco-) elastic properties at small deformations and plastic
that captures the thermal and mechanical history of the mate-
flow, and the entangled polymer network, which governs the
rial. The hydrostatic pressure p, the characteristic stress τ0
strain hardening. Accordingly, based on the original work
and the equivalent stress τ̄ are defined as:
of Haward and Thackray [11], the total Cauchy stress S is
decomposed in a driving stress Ss and a hardening stress Sr : 1
p = − tr(S) (38)
3
S = Ss + Sr . (31) kT
τ0 = ∗ (39)
V
The hardening is modelled with a neo-Hookean relation [16]: 
1  d d
Sr = G r B̃d (32) τ̄ = tr Ss · Ss (40)
2

123
Comput Mech (2008) 42:171–180 177

Table 1 PET material parameters A mesh convergence study is performed to find the
E(GPa) ν (–) Sa (–) η0,r (MPas) τ0 (MPa) optimum element discretization, which is a compromise
between the accuracy and required computational time. The
1.25 0.4 28.4 0.486 0.95 mesh is divided into a coarse and a fine region, as depicted
in Fig. 10. Only the part of the coating that delaminates (and
µ (–) G r (MPa) r0 (–) r1 (–) r2 (–)
thus deforms) needs to have a fine mesh, whereas the bulk
0.047 28.2 0.990 200 −4 coating can be handled with a relative coarse mesh. In this
section the discretization of the fine coating part is further
with k Bolzmann’s constant and V ∗ the activation volume. scrutinized.
The state parameter S is decomposed into a factor Sa , which The number of elements in x, y and z direction are varied
captures the thermo-mechanical history of the material, and and responses of the models are compared by considering
Rγ (γ̄ p ), which describes the softening kinetics [12]: the quasi-static peel-force, where a fine mesh should provide
the best accuracy. Therefore, the peel-force F is normalized
S(γ̄ p ) = Sa Rγ (γ̄ p ) (41) by the peel-force Fr calculated by the model with the most
refined mesh. Evidently, the resulting conclusions are only
where γ̄ p is the effective plastic strain. The initial value valid for the range of geometries and discretization studied.
of S equals Sa since the value of Rγ is normalized at 1 In Fig. 11a the influence of the amount of elements in
and decreases towards 0 with increasing plastic strain. The y-direction (width) of the model is shown for two model
Carreau-Yassuda function describes the softening character- widths. With four elements in y-direction the relative error
istic: is 0.5%, which is considered as acceptable. The two mod-
(1 + (r0 exp(γ̄ p ))r1 )(r2 −1)/r1 els with different widths yield almost identical results. This
Rγ (γ̄ p ) = (42) might indicate that the width of the model has little influ-
(1 + r0 r1 )(r2 −1)/r1
ence on its response and that the geometry can be reduced to
where r0 , r1 and r2 are fitting parameters. The material param- plane-strain. This is investigated in Sect. 9.
eters for PET have been determined in a previous study [20] The amount of elements in z-direction have a more pro-
and are given in Table 1. nounced influence on the peel force, up to 2% when only
a single element over the thickness is used (see Fig. 11b).
An error smaller than 0.5% is achieved with at least nine
7 Numerical model elements over the coating thickness.
The influence of the element size in the x-direction is
A 3D FEM model of the peel experiment is made. The model shown in Fig. 11c, where it is also clear that the character-
geometry and boundary conditions are shown in istic length, δ, does not have a pronounced influence on the
Fig. 10. The model consists of a polymer coating layer, a convergence. An element length of le = 2 µm leads to a
cohesive zone layer and a substrate. The steel is modelled maximum error of 0.5%.
as an isotropic linear elastic solid with a Young’s modulus The amount of elements in a model approximately equals:
of 210 GPa and a Poisson’s ratio of 0.3. The interface has a n elem = (1 + n z ) n x n y . For the number of degree-of-free-
initial zero thickness and is modelled with 3D cohesive zones doms (DOFs) an empirical expression is used: n DOF =4n elem .
and the presently proposed LDF. The constitutive model of The computational time for a single load step is related to
the coating layer was outlined in Sect. 6. The coating and sub- n DOF as: tinc = (n DOF )n , where n = 1.34 (see Fig. 11d).
strate are meshed with eight-node hexahedral element with Simulations with different load step sizes have been car-
trilinear interpolation functions. Due to symmetry conditions ried out and it was found that the relative error remains well
the model only encompasses a quarter of the specimen, which below 1% for all step sizes. However, with increasing step
is properly incorporated in the boundary conditions applied. sizes, not surprisingly, convergence may be lost. The largest
The model is loaded by a displacement u x on one end of the step size that resulted in a stable simulation was 100 nm/inc.
polymer coating, as shown in Fig. 10. It was found that this step size is dictated by the stability of
the PET constitutive model.
It may be worth verifying whether the steps are not too
large compared to the characteristic length of the cohesive
zone (δ). Theoretically, a cohesive zone should dissipate the
interfacial energy (per unit area) φ, which is equal to the
area under the curve in Fig. 6. If, however, the incremental
steps are too large, the numerical integration of the traction–
Fig. 10 The 3D peel model with boundary conditions separation law may yield an inaccurate dissipated energy. On

123
178 Comput Mech (2008) 42:171–180

Fig. 11 The normalized (a) 1.012 (b)


width = 1.0 mm
quasi-static peel forces as a width = 0.5 mm 1
function of the number of 1.01

elements in y-direction (a), 1.008


0.995
z-direction (b) or element length
1.006
(c). The influence of the number

r
r

F/F
F/F
of DOF on the CPU time of a 1.004 0.99
single increment (d)
1.002
0.985
1 thickness = 30 µm
thickness = 15 µm
0.998
0.98
0 5 10 15 20 0 5 10 15 20 25 30
# elem in y # elem in z

(c) 1.005 (d) 10 Sim data


1 Lin. fit

0.995

0.99

∆t [s]
F / Fr

1
10
0.985

0.98

0.975 δ = 3.7 µm
δ = 1.2 µm
0
0.97 10
−1 0 1 2 3 4 5
10 10 10 10 10 10 10
le [µm] #DOF

this basis, a simple analysis is performed to estimate the error


as a function of the incremental opening of a cohesive zone.
From this analysis it follows that the incremental cohesive
zone opening should be smaller than 13 δ in order to dissipate
more than 99% of φ.
Since the displacement is controlled at the end of the poly-
mer coating and the coating stretches at least 50%, the max-
imum incremental step size increases to 21 δ. For a stable
simulation the step size may not be larger than 100 nm/inc.
So, as long as δ > 200 nm, this maximum load step size can
be retained.
Fig. 12 Optical microscope image of the tensile specimen during a
peel experiment after 3.7 mm of delamination

8 Application
cohesive zones (see Fig. 6) do not have a noticeable influence
A 3D peel-off model is made with ten elements in y-direction on the simulation results because of the continuous tensile
and with three elements in z-direction. The elements have loading conditions of the structure.
a length of 2 µm in the x-direction. The load step size is
0.1 µm/inc. Numerical results are here compared qualita-
tively with the experiments. These results will be used in 9 Plane-strain and 3D
future work to identify the interfacial parameters also quan-
titatively. Here the interfacial parameters are taken as φ = Three-dimensional simulations are computational expensive,
200 J m−2 and Tmax = 20 MPa, which are in a realistic range. so it is worthwhile to investigate the possibility to use a model
In Fig. 12 an image is shown of the peel specimen. The with a 2D plane-strain geometry instead. In Fig. 14 the 3D and
deformed geometry can be compared with the results of the 2D models are compared by plotting their force–displacement
simulations, shown in Fig. 13 and a good qualitative agree- curves. Clearly, the plane-strain model overestimates the
ment is found. The three types of unloading behaviour of the quasi-static peel force by 4%. This overestimation is

123
Comput Mech (2008) 42:171–180 179

tent way because these models use a local coordinate basis to


decompose the opening displacement. A LDF is proposed to
overcome the shortcomings of the classical SDF. In the LDF a
global basis is used and only one relation between the traction
and the opening displacement is defined. As a consequence
there is only a single mode-independent work-of-separation,
which is a valid assumption regarding the presence of fibrils
at the interface. A more general LDF cohesive zone model
with a mode-dependent work-of-separation may be more rel-
evant for other material systems.
A 3D cohesive zone element with the LDF is developed
and implemented in a finite element solution framework. The
detection of contact in a cohesive zone element with the LDF
Fig. 13 Model predictions of the deformed geometry and the equiv-
is no longer straightforward and has been solved by defining
alent Von Mises stress after a delamination of 1.0 mm. Element edges
and cohesive zones are made invisible a normal vector with respect to the cohesive zone element
mid-plane.
The simulation of a peel experiment with the use of the
1.65 presented 3D model illustrates the added value of the devel-
oped 3D cohesive zone element with the LDF. Good qualita-
1.6
Peel−off force [N]

tive agreement is found between the model predictions and


1.55 experimental observations of a peel-off test. Future work con-
1.5
sists in quantitative determination of the interface parameters
on the basis of detailed experimental measurements.
1.45

3D simulation Acknowledgments This research was carried out under project num-
1.4
2D simulation ber MC2.03149 in the framework of the Strategic Research Programme
1.35 of the Netherlands Institute for Metals Research in The Netherlands
0.5 1 1.5 2 2.5 3 3.5 (www.nimr.nl).
Displacement [mm]

Fig. 14 The peel force as a function of the displacement as calculated


by a 3D and a plane-strain model
References

caused by the plane-strain condition, which prevents defor- 1. Abdul-Baqi A, Van der Giessen E (2001) Indentation-induced
mations in y-direction. The advantage of the 2D model with interface delamination of a strong film on a ductile substrate. Thin
respect to the 3D model is the significant reduction in calcu- Film Solids 381:143–154
2. Baaijens FPT (1991) Calculation of residual stresses in injection-
lation time. Depending on the width of the model and thus the molded products. Rheol Acta 30:284–299
element discretization in the width direction, the calculation 3. Boelen B, Den Hartog H, Van der Weijde H (2004) Product perfor-
time is typically reduced by a factor 15–50. This makes a mance of polymer coated packaging steel, study of the mechanism
plane-strain model an attractive alternative to perform qual- of defect growth in cans. Prog Org Coat 50:40–46
4. Chandra N, Li H, Shet C, Ghonem H (2002) Some issues in the
itative parameter studies and useful in the first steps towards application of cohesive zone models for metal–ceramic interfaces.
an iterative fitting procedure on quantitative experimental Int J Solids Struct 39:2827–2855
results. 5. Chvedov D, Jones R (2004) Frictional behavior of rolled surfaces
coated with polymer films. Surface Coat Technol 188–189:544–
549
6. Cornec A, Scheider I, Schwalbe KH (2003) On the practical appli-
10 Conclusion cation of the cohesive model. Eng Frac Mech 70:1963–1987
7. De Borst R (2003) Numerical aspects of cohesive-zone models.
Eng Frac Mech 70:1743–1757
It is shown that delamination in polymer coated steels typi- 8. Eyring H (1936) Viscosity, plasticity, and diffusion as examples of
cally takes place through a process of fibrillation. This mech- absolute reaction rates. J Chem Phys 4:283–295
anism involves large displacements and deformations at the 9. Govaert LE, Timmermans PHM, Brekelmans WAM (2000) The
interface and in the surrounding bulk materials. Cohesive influence of instrinsic strain softening on strain localization in poly-
carbonate: modeling and experimental validation. J Eng Mat Tech
zone models are a convenient tool to describe this failure 122:177–185
mechanism. However, classical cohesive zone models cannot 10. Graziano F (2000) Coil and sheet coating. Metal Finish 98:175–
cope with large displacements and deformations in a consis- 176

123
180 Comput Mech (2008) 42:171–180

11. Haward R, Thackray G (1968) The use of a mathematical model to 17. Tvergaard V (1990) Effect of fibre debonding in a whisker-
describe isothermal stress–strain curves in glassy polymers. Proc reinforced metal. Mat Sci Eng A 125:203–213
R Soc Lond A 302:453–472 18. Van den Bosch MJ, Schreurs PJG, Geers MGD (2006) An
12. Klompen ETJ, Engels TAP, Govaert LE, Meijer HEH (2005) Mod- improved description of the exponential Xu and Needleman
eling of the postyield response of glassy polymers: inlfuence of cohesive zone law for mixed-mode decohesion. Eng Frac Mech
thermomechanical history. Macromolecules 38:6997–7008 73:1220–1234
13. Leonov AI (1976) Nonequilibrium thermodynamics and rheology 19. Van den Bosch MJ, Schreurs PJG, Geers MGD (2007) A cohesive
of viscoelastic polymer media. Rheol Acta 15:85–98 zone model with a large displacement formulation accounting for
14. Qiu Y, Crisfield MA, Alfano G (2001) An interface element for- interfacial fibrilation. Eur J Mech 26:1–19
mulation for the simulation of delamination with buckling. Eng 20. Van der Aa HCE, Van der Aa MAH, Schreurs PJG, Baaijens FTP,
Frac Mech 68:1755–1776 Van Veenen WJ (2000) An experimental and numerical study of
15. Roychowdhury S, Arun Roy Y, Dodds RH (2002) Ductile tear- the wall ironing process of polymer coated sheet metal. Mech Mat
ing in thin aluminium panels: experiments and analyses using 32:423–443
large-displacement, 3-D surface cohesive elements. Eng Frac Mech 21. Xu XP, Needleman A (1993) Void nucleation by inclusions deb-
69:983–1002 onding in a crystal matrix. Mod Sim Mat Sci Eng 1:111–132
16. Tervoort TA, Govaert LE (2000) Strain-hardening behavior of
polycarbonate in the glassy state. J Rheol 44:1263–1277

123

You might also like