Specifying Constituent Properties for a Composite Material
In multiscale material modeling, each composite material consists of one matrix material
and one or more inclusions including voids. In Abaqus/Explicit, the composite material must consist of one matrix material and only one inclusion; the
inclusion material must be linear elastic with constant properties, and the matrix material
can be only linear elastic or elastic-plastic with isotropic hardening.
Input File Usage
Use the following option to define the properties of the matrix in the
composite:
Mean-Field Homogenization for Linear Elastic Composites
The mean-field strain and stress in each phase I is defined as
where
The averaged macro fields can be written as
where is the volume fraction of the matrix phase and is the volume fraction of the inclusion phase. For the single-inclusion
case, if all the constituents are linear elastic, the strain in the inclusion is related to
the strain in the matrix through a concentration tensor .
The strain and stress in the inclusion/matrix can also be related to the macro
strain/stress through the strain/stress concentration tensor and :
The form of the concentration tensor defines the homogenization methods below.
Voigt and Reuss Models
The simplest mean-field homogenization formulations are the Voigt and Reuss models. These
models do not take into account the shape or the orientation of the inclusions; however,
they provide upper and lower bounds of the macro stiffness modulus and, therefore, can be
used for validation. The Voigt model is also used in a two-step approach to model
composites with multiple inclusions.
The Voigt model assumes uniform strain in the composite , which gives
The Reuss model assumes uniform stress in the composite , which gives
where is the stiffness of the matrix and is the stiffness of the inclusion.
Input File Usage
Use the following option to specify the Voigt model as the homogenization
method:
Micromechanical Models for Unidirectional Composites
The Chamis model (Chamis, 1989) and Composite Cylinder Assemblage (CCA) model (Christensen, 1979) are micromechanical homogenization models
developed and validated for unidirectional composites.
With the composite stiffness known from the micromechanical models, the concentration
tensor can be computed using
where is the composite stiffness, is the stiffness of the matrix, and is the stiffness of the inclusion.
Input File Usage
Use the following option to specify the Chamis model as the homogenization
method:
The more sophisticated mean-field homogenizations are based on Eshelby's solution. In his
1957 paper (Eshelby, 1957), Eshelby solved a single-inclusion
problem described in Figure 1. Figure 1. Eshelby's problem: an ellipsoid is cut out of an infinite matrix, undergoes an
eigenstrain, and is inserted back into the matrix.
Eshelby found that the strain inside the constrained ellipsoid is uniform and given by
where represents a point inside the inclusion and is the eigenstrain. The Eshelby tensor (E) is a
function of the stiffness of the matrix and also the shape and orientation of the
inclusion. To calculate Eshelby's tensor usually involves numerical integration over the
surface of the ellipsoidal inclusion; however, analytical solutions are available (Mura, 1987) for simple shapes such as a sphere,
cylinder, penny, oblate, and prolate. Based on Eshelby's solution, the following models
are proposed based on different assumptions about the properties of the outer material
surrounding the inclusions.
Mori-Tanaka Model
The Mori-Tanaka formulation assumes that each inclusion behaves like an isolated
inclusion and the strain in the matrix is considered as the far-field strain. Therefore,
the concentration tensor can be written as
The inverse Mori-Tanaka formulation assumes a high volume fraction of the inclusion and
permutes the properties of the matrix and inclusion in a single inclusion problem, which
gives
The Mori-Tanaka and inverse Mori-Tanaka formulations give the upper and lower bounds of
the composite stiffness for low and high volume fractions. The balanced formulation is
the interpolative formulation, which can be written as
where is the smooth interpolation function proposed by Lielens (1999)
You can define the concentration tensor directly by specifying all of its components. The user-defined
concentration tensor must be defined in the local coordinate system of the inclusion. You
can use homogenization models other than those listed above, calibration from experimental
results, or FE-RVE models to obtain the value of the
concentration tensor.
Input File Usage
Use the following option to define the fourth-order strain concentration tensor used
to calculate the strain in a constituent:
Mean-field homogenization can also be used to compute effective thermal properties and
micro-level thermal fields of the constituents.
Conductivity
The effective conductivity of composite materials can also be computed using Eshelby's
single-inclusion method described in Mean-Field Homogenization for Linear Elastic Composites. For example, Mori-Tanaka method gives
the following conductivity concentration tensor
where is the modified 3x3 second-order Eshelby's tensor and are second-order conductivity tensor of the constituents.
The relationship between the temperature gradient in the constituents is given by
and the macro-level temperature gradient is the volume-averaged temperature
gradient of the constituents. Chamis and CCA formulations are not supported for
conductivity.
Specific Heat and Latent Heat
The specific heat and latent heat of the composite can be compute using the mass-averaged
value of the constituents.
Incremental Mean-Field Homogenization for Nonlinear Composites
An incremental linearization approach is used to extend the homogenization models to
composites with nonlinear behavior. The tangent stiffness matrix is used in place of the
linear elastic modulus to compute the concentration tensor
Since the tangent matrix is not constant in the nonlinear case and is usually a function of
the strain, iterations are performed to guarantee a converged solution of the strain
increment in each constituent.
Isotropization Methods
Analytical expression of Eshelby's tensor is available only if the inclusions are
ellipsoidal and the matrix material is isotropic or transversely isotropic. For matrix
materials that are anisotropic, using the stiffness directly to compute the Eshelby's
tensor can result in stiffer prediction for the composite behavior (Doghri, 2003). The isotropic part of the tangent
stiffness matrix can generally be written as
where is the bulk modulus and is the shear modulus. and are the volumetric and deviatoric part of the fourth-order identity
tensor. The method used to extract the isotropic part of the stiffness matrix is not
unique, and the accuracy of the prediction varies depending on the properties of the
matrix material.
General Method
The general method is the most common isotropization method, and it extracts the
isotropic part using
The general method can be used for any anisotropic tangent matrix. This method is the
default isotropization method.
The spectral isotropization method can be used if isotropic hardening is defined in the
matrix material (Doghri, 2003). The isotropic part of the tangent
matrix is given by
where is the elastic bulk modulus, is the elastic shear modulus, is the yield stress, and is the equivalent plastic strain.
For matrix materials with significant softening after yielding, both the general method
and the spectral method might give predictions that are too stiff. The modified spectral
method proposed by Selmi et. al. (Selmi, 2011) improves the accuracy of the spectral
method by evaluating the derivative of the yield stress at a larger equivalent plastic
strain
where , , and are the isotropization parameters. You can calibrate these parameters
by comparing the mean-field solution to experimental data or to FE-RVE solutions. The
isotropization parameters are irrelevant if the constituent does not have plastic
behavior.
Input File Usage
Use the following option when all the parameters are constant and is the only parameter with a nondefault value:
Options are available to use the isotropic projection of the tangent matrix to compute
different parts of the concentration tensor. For better predictions, Doghri et al. (Doghri, 2010) recommended the following:
If the general method is used, only Eshelby's tensor is computed with the isotropic
projection.
If the spectral/modified spectral method is used, Hill's tensor given by is computed with the isotropic projection.
It is possible that some choices can lead to unphysical prediction of microlevel strain
resulting in unstable material behavior in the composite. In addition, in Abaqus/Standard analyses, the tangent stiffness matrix for the composite might lose its symmetry after
homogenization; therefore, you must use the unsymmetric solver to achieve convergence.
Input File Usage
Use the following option to use the isotropic projection of the matrix stiffness to
compute all parts of the concentration tensor by default:
For composites with multiple inclusions, a multistep homogenization approach is used, as
shown in Figure 2. The composite is decomposed into "grains," with each grain
containing one inclusion family and the matrix. The inclusions in each family have the same
material properties, aspect ratio, and orientation. In the first step homogenization is
performed in each grain using the user-specified formulation; in the second step, the Voigt
formulation is used to compute the properties of the overall composite.
An alternative approach is to use the Mori-Tanaka scheme in both the first and the second
step, assuming the average strain in the matrix is uniform across all grains. This approach
is equivalent to the direct Mori-Tanaka approach proposed by Benveniste (Benveniste et al., 1991). In an Abaqus/Standard analysis, the second approach has the drawback that it might result in an unsymmetric
effective modulus when the inclusions are misaligned and nonisotropic. In such cases, use of
the unsymmetric solver might be necessary.
Use the following option to use the direct Mori-Tanaka formulation for the multiphase
composite, assuming the average strain in the matrix is uniform across all grains:
You can specify the microstructure of the composite by giving the shape, volume fraction,
aspect ratio, and fiber orientation of each constituent. You can use a distribution to
define the volume fraction, aspect ratio, and the orientation tensor (see Distribution Definition).
The orientation tensor can be used only with three-dimensional solid elements.
Specifying the Shape, Volume Fraction, and Aspect Ratio
The volume fraction and aspect ratio can be specified as spatial distributions. The
aspect ratio is given by as shown in Figure 3.
Figure 3. Different shapes of the inclusion.
The aspect ratio of a spherical inclusion is 1.0, and the aspect ratio of a cylindrical
inclusion is infinity; therefore, you do not need to specify the aspect ratio in these
cases.
Input File Usage
Use the following option to specify a prolate-shaped inclusion ():
Here and are the radii of the elliptical cross-section along the axis in the
local 2- and 3-directions.
Specifying a Fixed Orientation for the Inclusion and Void Phases
By default, the orientation of the inclusion or void phase is fixed and is aligned with
the x-axis of the local coordinate system specified in the section
definition. You can specify a different orientation by giving the three components of the
orientation vector in the local coordinate system, as shown in Figure 4. The components of the orientation unit vector, , are
The local 3-direction of the inclusion (void included) is the projection of the local
z-axis defined in the section onto the surface normal to the
inclusion orientation (1-direction). If the local z-axis is within
0.1 degree of being parallel to the 1-direction, the local 3-direction is the projection
of the local y-axis onto the surface. The local 2-direction is then
at right angles to the local 3-direction, so that the local 2-direction, local
3-direction, and the inclusion orientation (1-direction) form a right-handed set. The
local 2 and 3 directions are relevant only if the shape of the inclusion (void included)
is an elliptical cylinder.
Figure 4. Orientation of the inclusion (void included).
Input File Usage
Use the following option to specify a fixed fiber orientation:
Specifying a Second-Order Orientation Tensor in Abaqus/Standard
For composites containing short fibers, a second-order orientation tensor, , is usually used to describe the dispersion of the fiber orientation
where is the integral over the surface of the unit sphere (all possible
directions of ) and is the orientation distribution function
(ODF). The spatial distribution of the orientation tensor
is usually given by injection molding software and can be used directly to compute the
effective moduli of the composite material that is linear elastic. If the material is
nonlinear, the average fields given by must be computed during the analysis, and the
ODF is required to calculate the integral. When only the
second-order orientation is available, Abaqus recovers the ODF using the approach proposed by Onat
and Leckie (1988)
for three-dimensional orientations, where and are the tensor basis functions
and and are the deviatoric versions of the orientation tensors
The fourth-order orientation tensor is required during the ODF reconstruction. It is
computed using a closure approximation, which is a formula that approximates the
fourth-order tensor in terms of the known second-order tensor. A popular closure
approximation is the hybrid closure approximation proposed by Advani and Tucker (1987), which is constructed using the interpolation of
linear and quadratic closure approximation
where the linear closure is given by
and the quadratic closure is given by
The weight function is defined as for three-dimensional orientations.
Other available closure approximations are the orthotropic smooth approximation (Advani and Tucker, 1987), the orthotropic fitted (ORF) approximation
(Cintra and Tucker, 1995), and the orthotropic equaivalent of the
natural (ORE) approximation (VerWeyst, 1998).
Only one constituent in the composite is allowed to have orientations specified with the
second-order orientation tensor.
Taking into account the symmetry of the ODF , we compute the integral numerically as follows:
where and . The numerical integral is computed by subdividing the macro point into subdomains, with each subdomain having the same fiber orientation . Then a multistep homogenization (as described in Figure 2) is used to compute the macrolevel solutions. The
ODF is more accurately recovered if a large value of
N is used; however, the computational cost and memory usage also
increases substantially.
When the fiber direction is random in the three-dimensional space, the second-order
orientation tensor is , and the ODF is constant across all
subdomains.
When the direct Mori-Tanaka homogenization approach is used, the fiber response can be
calculated either individually in each subdomain or just once using the averaged state
over all subdomains. In both cases only the averaged fiber state is available for output.
The maximum principal direction of the second-order orientation tensor is used as the
1-direction of the local coordinate system for the inclusion.
In composites where the fiber material is linear elastic with uniform moduli across all
pseudo-grains, if we assume uniform matrix strain inside the composite, homogenization and
strain partitioning can be simplified. Instead of computing the integrals numerically, Abaqus/Standard uses the following equation to compute the orientation averaging terms, :
where , , , , and are the five constants of the tensor that is transversely isotropic. ODF
reconstruction is not necessary and is not performed.
Input File Usage
Use the following option to specify the fiber orientation using a second-order
tensor:
Specifying a Second-Order Orientation Tensor in Abaqus/Explicit
In Abaqus/Explicit, when a second-order orientation tensor, , is used to specify the dispersion of the fiber material, the direct
Mori-Tanaka approach is used. Because the fiber material can be only linear elastic with
uniform moduli across all pseudo-grains, ODF reconstruction is not performed and
orientation averaging terms are computed directly.
The second-order orientation tensor is specified in the orientation definition of the
section where the multiscale material is assigned.
Input File Usage
Use the following option to specify the fiber orientation using a second-order
tensor:
In the presence of thermal strains due to a temperature change, the averaged strain/stress
in each constituent can be written as:
The relationship between and can be established through superposition of the two problems to achieve
the condition with (3).
Problem I solves a uniform stress field under the condition while assuming . The solution to problem I gives
where .
Problem II is purely mechanical assuming . Solving gives the solution
By superimposing the constituent strain solutions from problem I and problem II, we get
which results in the relationship between and below
Thermal expansion is supported only for plane strain and three-dimensional solid
elements.
Composites with Viscoelasticity
Composites with viscoelasticity are supported only in Abaqus/Standard.
When one or more constituents have viscoelasticity defined in the time domain, the
homogenization approach described above is no longer appropriate because it does not take
into account the relaxation of stress with time. However, similar homogenization can be
performed in the frequency domain, and homogenized properties of the composite in the time
domain can be computed using homogenized properties at different frequencies. This
homogenization approach is applicable only when the viscoelasticity in the constituents is
defined in the time domain using a Prony series expansion and includes the following steps:
The shear and bulk moduli of all the constituents are transformed to the frequency
domain using Fourier transforms (see details in Prony Series Parameters).
The mean-field homogenization is performed in the frequency domain the same way as
discussed above, with the exception that the moduli are complex numbers instead of real
numbers. The real part of the composite modulus is the storage modulus, , and the imaginary part is the loss modulus, .
Assuming the time response of the composite can be written in a similar form as that in
the constituents where viscoelasticity is defined using a Prony series expansion
where and is the instantaneous elastic modulus of the composite, the homogenized
storage modulus of the composite, , can be written in the frequency domain as follows:
The parameters are selected to be equally spaced numbers on a logarithmic scale in the
range . and are the minimum and maximum frequencies used to define the
viscoelasticity in all constituents. The parameters are computed through a least-squares curve fit at frequencies .
You can only specify the thermorheologically simple temperature effect in the matrix
material. The composite is assumed to inherit the same temperature effect as the matrix
material.
Other limitations include:
The inverse Mori-Tanaka and balanced homogenization formulations are not supported when
you use the orientation tensor to define the direction of the inclusion.
When you use the orientation tensor to define the direction of the inclusion, the
elasticity specified in the inclusion material must satisfy the transversely isotropic
condition, and no other inclusions can be defined in the composite.
You cannot use the orientation tensor to define the direction of the inclusion in which
viscoelasticity is defined.
Distributions other than distributions of the orientation tensor are not allowed in the
material definition.
Temperature and field variable dependencies are allowed to define the elasticity of the
matrix material only, and only one inclusion is allowed in this case.
Other than viscoelasticity, only the following properties are supported in the inclusion material:
Conductivity
Density
Elasticity
Thermal expansion
Specific heat
Other than viscoelasticity, only the following properties are supported in the matrix material:
Conductivity
Density
Elasticity
Thermal expansion
Specific heat
Cure properties
Because the multiscale material is replaced by a composite with prehomogenized properties,
the microlevel output is not available. Other than viscoelasticity, Abaqus/Standard also computes the homogenized values of all the other supported properties.
Composites Curing
Mean-field homogenization can be used to model composite behavior during a curing process
(Curing Processes in Polymers). Similar to composites with viscoelasticity,
a multiscale material is replaced by a composite with prehomogenized properties. The same
limitations described in Composites with Viscoelasticity
also apply.
The composite inherits the following properties from the matrix material definition:
Cure kinetics
Cure max conversion
Cure glass transition temperature
Mean-field homogenization is used to compute the following cure properties for the composite:
Cure heat generation. The specific heat of the reaction is computed using mass-averaged
value of the constituents.
Cure shrinkage. Levin's theorem is used to compute the effective thermal expansion
coefficient (Composites with Thermal Expansion) and to compute the
effective cure shrinkage coefficient of the composite.
Composites with Damage
When damage and failure are defined at the constituent level, the damage in each
constituent contributes to the overall damage in an indirect way through stress averaging
and strain partitioning. As the stiffness of the constituent decreases, the strain increment
in this constituent is likely to increase based on the formulation of the homogenization. In
reality, the damage behavior of each constituent is likely to interact with each other and
affect the overall damage behavior of the composite; therefore, you might need to specify an
additional damage variable, , for the composite. At any time during the analysis, the stress tensor of
the composite is given by
where , and is the effective stress of the constituent.
You can choose to combine the damage variables, , in all constituents to get the overall damage behavior of the composite.
The overall damage variable, , can be computed as the maximum of :
or you can choose to combine the damage variables in a multiplicative sense:
You can also specify using user subroutine UDAMAGEMF. In addition, you can
specify a separate damage variable, , for each constituent. This variable is used to account for the additional
contribution to the overall damage behavior due to interaction (for example, a fiber
material might buckle more easily if cracks start to develop in the matrix material) as
follows:
Input File Usage
Use the following option to specify that contributes to in a maximum sense:
This simple example illustrates how to define a multiscale material.
The name of the multiscale material is COMPOSITE. The material
consists of one inclusion material MAT2 embedded in the matrix
material MAT1. The shape of the inclusion is prolate. The Mori-Tanaka
homogenization method is used. The inclusion has a volume fraction of and an aspect ratio of . The direction of the inclusion is fixed and defined with a vector . The name of the matrix is MATRIX_MAT1, and the
name of the inclusion is FIBER_MAT2.
In Abaqus/Standard the multiscale material model can be used with
three-dimensional solid elements, including stress/displacement elements, diffusive heat
transfer elements, and coupled temperature-displacement elements;
two-dimensional solid elements (stress/displacement only);
membrane elements;
three-dimensional conventional shell elements (stress/displacement only); and
continuum shell elements (stress/displacement only).
In Abaqus/Standard additional plane stress iterations are included for two-dimensional plane stress elements
and shell elements.
In Abaqus/Explicit the multiscale material model can be used only with three-dimensional solid
stress/displacement elements.
Hybrid elements and incompatible mode elements are not supported.
Output
By default, the output variables are macrolevel results. In addition, you can also request
microlevel output. The name of the constituent is appended to the end of the output
variable.
For composites reinforced with dispersed short fibers, the fiber output variables are
computed in the local system with the local fiber direction as the maximum principal
direction of the orientation tensor. The matrix output is available only if the direct
Mori-Tanaka homogenization approach is used. Otherwise, only averaged stress and strain are
available. If plasticity is specified in the matrix material, averaged plastic strains are
also available. The same matrix output limitation also applies to composites reinforced with
multiple fiber families.
Microlevel strain output is not available in Abaqus/Explicit.
Input File Usage
Use the following options to request field output for the constituents:
Advani, S., and C. Tucker, “The Use of Tensors to Describe and
Predict Fiber Orientation in Short Fiber Composites,” Journal of Rheology, vol. 31 (8), pp. 751–784, 1987.
Levin, V.M., “On the Coefficients of Thermal
Expansion of Heterogeneous Materials,” Mechanics of
Solids, vol. 2, pp. 58–61, 1967.
Benveniste, Y., and G. J. Dvorak, “On a Correspondance Between Mechanical
and Thermal Effects in Two-Phase Composites,” Micromechanics and
Inohomogeneity, pp. 65–81, 1990.
Benveniste, Y., G. J. Dvorak, and T. Chen, “On Diagonal and Elastic Symmetry of
the Approximate Effective Stiffness Tensor of Heterogeneous Media,” Journal of the Mechanics and Physics
of Solids, vol. 39(7), pp. 927–946, 1991.
Chamis, C.C., “Mechanics of Composite Materials:
Past, Present and Future,” Journal of Composites Technology and
Research, vol. 11, pp. 3–14, 1989.
Christensen, R.M., “Solutions for Effective Shear
Properties in Three Phase Sphere and Cylinder Models,” Journal of the Mechanics and Physics
of Solids, vol. 27, pp. 315–330, 1979.
Cintra, J., and C. Tucker, “Orthotropic Closure Approximations for
Flow-Induced Fiber Orientation,” Journal of
Rheology, vol. 39, pp. 1095–1122, 1995.
Doghri, I., I. Adam, and N. Bilger, “Mean-Field Homogenization of
Elasto-Viscoplastic Composites Based on a General Incrementally Affine Linearization
Method,” International Journal of
Plasticity, pp. 219–238, 2010.
Doghri, I., and A. Ouaar, “Homogenization of Two-Phase
Elasto-Plastic Composite Materials and Structures. Study of Tangent Operators, Cyclic
Plasticity and Numerical Algorithms,” International Journal
of Solids and Structures, vol. 40(7), pp. 1681–1712, 2003.
Eshelby, J.D., “The Determination of the Elastic Field
of an Ellipsoidal Inclusion and Related Problems,” Proceedings of the Royal Society of
London, pp. 376–396, 1957.
Kammoun, S., L. Brassart, G. Robert, I. Doghri, and L. Delannay, “Micromechanical Modeling of Short
Glass-Fiber Reinforced Thermoplastics-Isotropic Damage of Pseudograins,” American Institute of Physics
Conference Proceedings 1353, pp. 974–977, 2011.
Lielens, G., “Micro-Macro Modeling of Structured
Materials,” PhD Thesis Universite Catholique de
Louvain Belgium, 1999.
Mura, T., Micromechanics of Defects in Solids, MARTINUS NIJHOFF Publishers, 1987.
Onat, E.T., and F.A. Leckie, “Representation of Mechanical Behavior
in the Presence of Changing Internal Structure,” Journal of Applied Mechanics, vol. 55(1), pp. 1–10, 1988.
Pierard, O., and I. Doghri, “An Enhanced Affine Formulation and the
Corresponding Numerical Algorithms for the Mean-Field Homogenization of
Elasto-Viscoplastic Composites,” International Journal
of Plasticity, pp. 131–157, 2006.
Selmi, A., I. Doghri, and L. Adam, “Micromechanical Simulations of Biaxial
Yield, Hardening and Plastic Flow in Short Glass Fiber Reinforced Polyamide,” International Journal of Mechanical
Sciences, vol. 53(9), pp. 696–706, 2011.
VerWeyst, B.E., “Numerical Predictions of Flow-Induced
Fiber Orientation in 3-D Geometries,” Ph.D. thesis,
University of Illinois at Urbana-Champaign, 1998.