Fiber metal laminates (FMLs) are composed
of laminated thin aluminum layers bonded with intermediate glass
fiber-reinforced epoxy layers. FMLs are of
great interest in the aerospace industry due to their superior properties, such
as high fracture toughness and low-density when compared to solid aluminum
sheets.
Cohesive elements are used to model the interlaminar delamination, and three different
material models are used to predict the behavior of the fiber-reinforced layer:
Hashin damage model for unidirectional fiber-reinforced materials (Hashin Criterion).
Damage model proposed by Linde et al. (2004), which is implemented in user subroutine
UMAT.
Both Abaqus/Standard and Abaqus/Explicit are used for simulation when the Hashin damage model is used for the fiber-reinforced epoxy
layers.
This type of problem is important in the aerospace industry since blunt notches (for example,
fastener holes) commonly occur in airplane structures; the strength of the structure
containing a blunt notch is a crucial design parameter. The models presented in this example
demonstrate how to predict the blunt notch strength, the failure patterns of the fiber and
matrix within the fiber-reinforced epoxy layer, and the delamination between different layers
of FMLs.
Problem description and material characteristics
Figure 1
shows the geometry of the laminate containing the blunt notch for this example.
The laminate is subjected to uniaxial tension in the longitudinal direction.
The laminate is made of three layers of aluminum and two layers of 0°/90° glass
fiber-reinforced epoxy. Only 1/8 of the laminate needs to be modeled, with
appropriate symmetric boundary conditions applied as shown in
Figure 2.
Figure 2
also shows the through-thickness lay-up of the 1/8 model.
The material behavior of aluminum is assumed to be isotropic elastic-plastic with isotropic
hardening. The Young’s modulus is 73,800 MPa, and the Poisson’s ratio is 0.33; the isotropic
hardening data are listed in Table 1.
The material behavior of the glass fiber-reinforced epoxy layers is assumed to be orthotropic,
with stiffer response along the fiber direction and softer behavior in the matrix. The
elastic properties—longitudinal modulus, ; transverse modulus, ; shear moduli, and ; and Poisson’s ratios, and —are listed in Table 2. The subscript “L” refers to the longitudinal direction (or fiber direction), and the
subscript “T” refers to the two transverse directions orthogonal to the fiber direction.
The damage initiation and evolution behavior is also assumed to be orthotropic. Table 3 lists the ultimate values of the longitudinal failure stresses, and ; transverse failure stresses, and ; and in-plane shear failure stress, . The superscripts “t” and “c” refer to tension and compression,
respectively. The fracture energies of the fiber and matrix are assumed to be =12.5 N/mm and =1.0 N/mm, respectively.
Three material models that use the parameters listed above are considered, as follows:
The material is modeled using the damage model proposed by Linde et al. (2004). This damage
model is implemented in user subroutine UMAT and is referred to in this
discussion as the UMAT model. Details of the UMAT model are provided below.
The material is modeled using the multiscale material model in Abaqus (see Mean-Field Homogenization), and
the damage model is implemented in user subroutine
USDFLD at the constituent level. Details
of the multiscale material and the damage model are provided below.
The adhesive used to bond neighboring layers is modeled using interface
layers with a thickness of t=0.001 mm. To simulate the
interlaminar delamination, these interface layers are modeled with cohesive
elements. The initial elastic properties of each interface are assumed to be
isotropic with Young’s modulus E=2000 MPa and Poisson’s
ratio =0.33.
The failure stresses of the interface layers are assumed to be
===50
MPa; the fracture energies are ===4.0
N/mm. The subscripts “n,” “s,” and “t” refer to the normal direction and the
first and second shear directions (for further discussion of the constitutive
modeling methods used for the adhesive layers, see
Defining the Constitutive Response of Cohesive Elements Using a Traction-Separation Description).
The plate is loaded with displacement boundary conditions applied at the
right edge. To simplify the postprocessing, the displacement loading is applied
at a reference point and an equation constraint is used to constrain the
displacement along the loading direction between the right edge and the
reference point. Except for those files designed exclusively to study the
effect of the loading direction on the strength, the loading direction (along
the global X-direction) aligns with the fiber direction of
the 0° fiber-reinforced epoxy layer.
UMAT model for fiber-reinforced epoxy layers
For fiber-reinforced epoxy layers, the primary model considered is based on the Hashin damage
model for unidirectional fiber-reinforced composites available in both Abaqus/Standard and Abaqus/Explicit. Alternatively, in Abaqus/Standard, the damage in the fiber-reinforced epoxy is also simulated using the model proposed by
Linde et al. (2004), which is implemented in user subroutine UMAT and is discussed below.
In the UMAT model, the damage initiation
criteria are expressed in terms of strains. Unlike the Hashin damage model in Abaqus, which uses four internal (damage) variables, the UMAT model uses two damage variables
to describe damage in the fiber and matrix without distinguishing between tension and
compression. Although the performance of the two models is expected to be similar for
monotonic loads, such as in this example problem, the results obtained might differ
considerably for more complex loads in which, for example, tension is followed by
compression. For the UMAT model, if the material is
subjected to tensile stresses that are large enough to cause partial or full damage (the
damage variable corresponding to this damage mode is greater than zero), both tensile and
compressive responses of the material are affected. However, in the case of the Hashin
damage model, only the tensile response is degraded while the material compressive response
is not affected. In many cases the latter behavior is more suitable for modeling
fiber-reinforced composites. In this section the governing equations for damage initiation
and evolution as proposed by Linde et al. (2004) are discussed, followed by a description of
the user subroutine UMAT implementation.
Damage in the fiber is initiated when the following criterion is reached:
where ,
,
and
are the components of the elasticity matrix in the undamaged state. Once the
above criterion is satisfied, the fiber damage variable,
,
evolves according to the equation
where
is the characteristic length associated with the material point. Similarly,
damage initiation in the matrix is governed by the criterion
where ,
,
and .
The evolution law of the matrix damage variable, ,
is
During progressive damage the effective elasticity matrix is reduced by the
two damage variables
and ,
as follows:
The use of the fracture energy-based damage evolution law and the
introduction of the characteristic length
in the damage evolution law help to minimize the mesh sensitivity of the
numerical results, which is a common problem of constitutive models with strain
softening response. However, since the characteristic length calculation is
based only on the element geometry without taking into account the real
cracking direction, some level of mesh sensitivity remains. Therefore, elements
with an aspect ratio close to one are recommended (for a discussion of mesh
sensitivity, see
Concrete Damaged Plasticity).
In user subroutine
UMAT the stresses are updated according to the following
equation:
The Jacobian matrix can be obtained by differentiating the above equation:
The above Jacobian matrix is not symmetric; therefore, the unsymmetric
equation solution technique is recommended if the convergence rate is slow.
To improve convergence, a technique based on viscous regularization (a generalization of the
Duvaut-Lions regularization) of the damage variables is implemented in the user subroutine.
In this technique we do not use the damage variables calculated from the aforementioned
damage evolution equations directly; instead, the damage variables are “regularized” using
the following equations:
where
and
are the matrix and fiber damage variables calculated according to the damage
evolution laws presented above,
and
are the “regularized” damage variables used in the real calculations of the
damaged elasticity matrix and the Jacobian matrix, and
is the viscosity parameter controlling the rate at which the regularized damage
variables
and
approach the true damage variables
and .
To update the “regularized” damage variables at time
,
the above equations are discretized in time as follows:
From the above expressions, it can be seen that
Therefore, the Jacobian matrix can be further formulated as follows:
Care must be exercised to choose an appropriate value for
since a large value of viscosity might cause a noticeable delay in the
degradation of the stiffness. To estimate the effect of viscous regularization,
the approximate amount of energy associated with viscous regularization is
integrated incrementally in user subroutine
UMAT by updating the variable
SCD as follows:
where
is the damaged elasticity matrix calculated using the damage variables,
and ;
and
is the damaged elasticity matrix calculated using the regularized damage
variables,
and .
To avoid unrealistic results due to viscous regularization, the above
calculated energy (available as output variable ALLCD) should be small compared to the other real energies in the
system, such as the strain energy ALLSE.
This user subroutine can be used with either three-dimensional solid
elements or elements with plane stress formulations. In the user subroutine the
fiber direction is assumed to be along the local 1 material direction.
Therefore, when solid elements are used or when shell elements are used and the
fiber direction does not align with the global
X-direction, a local material orientation should be
specified. The damage variables—,
,
,
and —are
stored as solution-dependent variables.
Multiscale model for fiber-reinforced epoxy layers
The material of the fiber-reinforced epoxy layers is also modeled with the multiscale
material model available in Abaqus/Standard. The Mori-Tanaka homogenization method is used, and the volume fraction of the fiber is
set to 50%, which is typical for fiber-reinforced epoxy composites. Both the matrix material
and fiber material are modeled with a linear elastic material model. The elastic moduli are
calibrated to match the composite properties listed in Table 2. The matrix material is assumed to be isotropic. The Poisson's ratio is 0.41, and the
calibrated Young's modulus is 6161.4 MPa. To better match the composite properties, the
fiber material is modeled with an orthotropic linear elastic material with the calibrated
properties listed in Table 4. The shape of the fiber is assumed to be cylindrical with infinite
length.
In the multiscale material model, the damage initiation criteria are expressed at the
constituent level rather than at the composite level. The stress limit values of the matrix
material and fiber material are obtained from microlevel solutions when the stress limits
listed in Table 3 are reached at the composite level.
Damage in the fiber material is initiated when one of the following criteria is reached:
Fiber tension :
Fiber compression :
Damage in the matrix material is initiated when one of the following criteria is reached:
Matrix tension :
Matrix compression :
The parameters in the above equations are listed in Table 5. The maximum values of the failure indices , , , and are saved as solution-dependent state variables. Once the maximum value of
the failure index exceeds 1.0, the damage variable is set to 1.0 and continues to have the value of 1.0 even though the
stresses might reduce significantly, which ensures that the material does not "heal" after
it is damaged. To improve convergence, the damage variables are "regularized" using the
following equations:
The value of the "regularized" damage variable is assigned to the field variable used to
define the elastic properties of the matrix and fiber materials. The Poisson's ratios remain
unchanged while the other elastic moduli reduce by 1×10−6 when the field
variables reach 1.0.
Both damage initiation and the viscous regularization of the damage variables are
implemented in user subroutine USDFLD.
Finite element model
The finite element model uses a separate mesh for each of the respective layers shown in Figure 2: two aluminum layers, two fiber-reinforced epoxy layers, and three adhesive layers. While
not required, a similar finite element discretization in the plane of the laminate, such as
that shown in Figure 3, can be used for all layers.
Modeling considerations for aluminum layers
Due to the interactions with the fiber-reinforced epoxy layers, the stress
state within the aluminum layers (especially surrounding the notch tip) cannot
be approximated using the plane stress assumption. To model this
three-dimensional plasticity stress state accurately, solid elements must be
used for the aluminum layers. In
Abaqus/Standard
incompatible mode elements (C3D8I) are used since local bending might exist in the post-failure
region surrounding the notch. For the
Abaqus/Explicit
analysis, reduced-integration elements (C3D8R) are used for modeling the aluminum layers.
Modeling considerations for glass fiber-reinforced epoxy layers
The plane stress assumption can be used safely within the fiber-reinforced epoxy layers;
therefore, either solid elements or shell elements can be adopted for these layers.
However, it is important to have an accurate representation of the through-thickness
geometry to model the interface between the adhesive and the fiber-reinforced epoxy
realistically. This is achieved most conveniently with solid elements or continuum shell
elements instead of conventional shell elements. The Hashin damage model for
unidirectional fiber-reinforced materials is available only for elements with a plane
stress formulation. Therefore, continuum shell elements are used with this model. Models
are also included in which continuum elements
(C3D8R or
C3D8) are used along with user subroutine
UMAT to model the fiber-reinforced
epoxy layers. The multiscale material model using continuum shell elements appears to give
the best results for the damage model described above. The model is also available for
plane stress elements and three-dimensional continuum solid elements.
Modeling considerations for adhesive layers
Cohesive elements (COH3D8) are used for the interface layers. The elastic response is
defined in terms of a traction-separation law with uncoupled behavior between
the normal and shear components. For convenience, a constitutive thickness of
1.0 mm is used so that we do not need to distinguish between the separation
displacement and the nominal strain (NE). However, since the actual thickness is 0.001 mm, the diagonal
terms in the elasticity matrix need to be scaled by the inverse of the actual
thickness as follows:
The quadratic nominal strain criterion is used for the damage initiation:
Results for each analysis are discussed in the following sections.
Abaqus/Standard
results
Damage to the fiber-reinforced epoxy plays a key role in the response for the loading considered.
Figure 4 shows the load-displacement curve for the 0° loading direction for the Linde and Hashin
(built-in) the damage models considered for the fiber-reinforced epoxy. The response shows
a “bilinear” shape before the sudden loss of loading capacity; that is, an initial linear
curve representing the initial elastic region, a smoothly deflecting nonlinear curve
representing the local plasticity, and a second linear curve representing the net section
yielding. The effect of the element type was studied using the UMAT model and
C3D8R,
C3D8, and
SC8R elements; and the results are
summarized in Figure 5 and Table 7. The numerical results obtained using different element types and different damage
models are similar and show a good agreement with the experimental results of De Vries
(2001).
The fiber and matrix damage patterns in the 0° fiber-reinforced epoxy layer at the failure load
are shown in
Figure 6 and Figure 7, for the Hashin damage model for fiber-reinforced materials;
It can be seen that the fiber damage in the 0° fiber-reinforced epoxy layer propagates
along the ligament above the blunt notch tip (that is, orthogonal to the loading
direction).
Figure 12 shows the matrix damage in the 90° layer for the damage model of Linde et al. (2004).
There is no fiber damage in the 90° fiber-reinforced epoxy layer prior to the sudden
fracture. Interlaminar damage is most severe between the 0° fiber-reinforced epoxy layer
and the aluminum layer. These observations are in agreement with the experimental results
of De Vries (2001).
The load-displacement results for different values of the viscosity parameter, are given in
Figure 13 and Table 6, for the Hashin damage model for fiber-reinforced materials;
The smaller the viscosity, the more abrupt the failure and the smaller the failure
strength. Although a viscosity of 0.001 seems to overestimate the failure strength by a
few percent (Table 6 and Table 8), the convergence is noticeably improved. Therefore, a viscosity of 0.001 is used for
all the other studies in this example.
For the Hashin damage model for fiber-reinforced materials, only the viscosity in the
fiber direction was varied while the viscosity in the matrix direction was kept constant
at 0.005. This improved convergence and did not markedly affect the results. For the
multiscale material model, the viscosity in the matrix material was also kept at 0.005,
while the viscosity in the fiber was varied between 0.0025 and 0.002. Since the viscosity
parameter in this model also dictates the damage evolution, it has a larger impact on the
failure strength, as shown in Table 9.
The effect of the loading direction on the blunt notch strength is studied
using the three-dimensional element, C3D8R, with the
UMAT model. Three tests are performed in which the local
material orientations in the 0°/90° fiber-reinforced epoxy are rotated by an
angle of 15°, 30°, and 45°, respectively. For example, for a loading angle of
15° the fiber orientation in the 0° fiber-reinforced epoxy layer would be at a
15° angle with respect to the X-direction, while the fiber
orientation in the 90° fiber-reinforced epoxy layer would be at an angle of
−75° with respect to the X-direction (Figure 16).
As can be seen in
Figure 17,
strain hardening is smaller for the larger loading angles. As can be seen in
Figure 18,
the failure strength decreases with the increasing loading angle and reaches
the minimum at the 45° loading angle (the response for even larger loading
angles is expected to be approximately symmetric with respect to the 45° angle
due to the symmetric nature of the 0°/90° fiber-reinforced epoxy layer). As
stated by De Vries (2001), this is expected and reflects the poor shear
properties of the fiber-reinforced epoxy layer.
In the above discussions the net blunt notch strength is defined as
,
where
is the length of the ligament above the notch and t is the
total thickness of the laminate. This example demonstrates that the approach
employed in the study can be used to predict the blunt notch strength of the
fiber metal laminates.
Abaqus/Explicit
results
In the Abaqus/Explicit simulation, we only consider loading along the 0° ply. The simulation is conducted
without damage stabilization, and no mass scaling is used. However, in order to reduce the
computational time, the total loading is applied in a short interval of time (0.001 s).
The overall load-displacement curve obtained from the explicit dynamic simulation is
compared with the Abaqus/Standard result (with viscosity of 0.001) in Figure 19. The results from the explicit dynamic simulation are presented using an antialiasing
filter to remove high frequency noise (see Filtering Output and Operating on Output in Abaqus/Explicit). The overall response compares well with the Abaqus/Standard results with some differences in the peak value of the load and in the post-peak
response. Damage stabilization is used in the Abaqus/Standard simulation to achieve convergence and is likely to change the overall response
(especially in the post-peak portion of the load-displacement curve). On the other hand,
the Abaqus/Explicit simulation does not use damage stabilization and is better able to capture the dynamic
behavior inherent in the damage and failure processes. The contour plots of various damage
variables in the 0° and 90° plies agree qualitatively with the corresponding plots
obtained from the Abaqus/Standard simulation using the Hashin damage model.
SC8R used in the fiber-reinforced
epoxy layer, a loading angle of 0°, and a viscosity of 0.002 in the fiber material
(using multiscale material model).
SC8R used in the fiber-reinforced
epoxy layer, a loading angle of 0°, and a viscosity of 0.001 in the fiber material
(using multiscale material model).
SC8R used in the fiber-reinforced
epoxy layer, a loading angle of 0°, and a viscosity of 0.0005 in the fiber material
(using multiscale material model).
SC8R used in the fiber-reinforced
epoxy layer, a loading angle of 0°, and a viscosity of 0.00025 in the fiber material
(using multiscale material model).
SC8R elements used in the fiber-reinforced epoxy
layer and a loading angle of 0° (using Hashin damage model).
References
De
Vries, T.J., “Blunt
and Sharp Notch Behavior of Glare
Laminates,” Ph.D dissertation, Delft
University
Press, 2001.
Hagenbeek, M., C. Van
Hengel, O. J. Bosker, and C. A. J. R. Vermeeren, “Static
Properties of Fibre Metal Laminates,” Applied
Composite
Materials, vol. 10207–222, 2003.
Linde, P., J. Pleitner, H. De
Boer, and C. Carmone, “Modelling
and Simulation of Fiber Metal Laminates,” Abaqus
Users’
Conference, 2004.
Tables
Table 1. Isotropic hardening data for aluminum.
Yield stress (MPa)
Plastic strain (%)
300
0.000
320
0.016
340
0.047
355
0.119
375
0.449
390
1.036
410
2.130
430
3.439
450
5.133
470
8.000
484
14.710
Table 2. Orthotropic elastic properties of fiber-reinforced epoxy.
(MPa)
(MPa)
(MPa)
(MPa)
55000
9500
5500
3000
0.45
0.33
Table 3. Orthotropic damage initiation properties of fiber-reinforced
epoxy.
(MPa)
(MPa)
(MPa)
(MPa)
(MPa)
2500
2000
50
150
50
Table 4. Orthotropic elastic properties of fiber material.
(MPa)
(MPa)
(MPa)
(MPa)
103739
12170
29655.4
4261.56
0.242
0.1594
Table 5. Damage initiation properties of matrix and fiber in the multiscale material.
(MPa)
(MPa)
(MPa)
(MPa)
(MPa)
(MPa)
4550
3700
48
142
30
66
Table 6. Net blunt notch strength (MPa) for different values of the viscosity parameter in fiber
direction (using Hashin damage model, viscosity in the matrix direction =0.005).
Numerical results
(SC8R, 0o loading angle)
Experimental results(De Vries,
2001)
=0.001
=0.0005
=0.00025
462.1
456.4
453.2
446
Table 7. Net blunt notch strength (MPa) for different element types used in the
fiber-reinforced epoxy layers (using UMAT model).
Numerical
results (=0.001,
0o loading angle)
Experimental results
(De Vries, 2001)
C3D8R
C3D8
SC8R
463.7
467.1
458.7
446
Table 8. Net blunt notch strength (MPa) for different values of the viscosity
parameter (using UMAT model).
Numerical results
(C3D8R, 0o loading angle)
Experimental results(De Vries,
2001)
=0.001
=0.0004
=0.00016
=0.000064
463.7
453.8
449.2
448.2
446
Table 9. Net blunt notch strength (MPa) for different values of the viscosity parameter in the
fiber material (using multiscale model, viscosity in the matrix material =0.005).