Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

8.

4 Structural Analysis of Composites With Finite Element Codes: An Overview


of Commonly Used Computational Methods
Johannes Reiner and Reza Vaziri, The University of British Columbia, Vancouver, BC, Canada
r 2018 Elsevier Ltd. All rights reserved.

8.4.1 Introduction 62
8.4.2 FE Modeling Techniques 62
8.4.2.1 Spatial Discretization 63
8.4.2.1.1 Microscale – Representative volume element 63
8.4.2.1.2 Mesoscale – Ply-based modeling 63
8.4.2.1.3 Macroscale modeling 64
8.4.2.1.4 Multiscale modeling 65
8.4.2.2 Temporal Discretization 65
8.4.2.2.1 Implicit scheme 65
8.4.2.2.2 Explicit scheme 65
8.4.3 Computational Damage and Failure Modeling 66
8.4.3.1 CDM 66
8.4.3.1.1 Damage initiation 67
8.4.3.1.2 Damage evolution 67
8.4.3.1.2.1 Sudden degradation models 67
8.4.3.1.2.2 Gradual degradation models 67
8.4.3.2 Discrete Failure Representation 68
8.4.3.2.1 Cohesive Zone Method and Virtual Crack Closure Technique to model delamination 68
8.4.3.2.1.1 Virtual crack closure technique 68
8.4.3.2.1.2 Cohesive zone method 68
8.4.3.2.2 Partition of unity for matrix cracking 69
8.4.3.3 Calibration of Model Parameters 69
8.4.3.3.1 Material parameters 69
8.4.3.3.1.1 Intra-laminar modeling 69
8.4.3.3.1.2 Inter-laminar modeling 70
8.4.3.3.2 Numerical parameters 70
8.4.3.4 Numerical Issues 70
8.4.3.4.1 Mesh-dependency 70
8.4.3.4.2 Convergence in implicit schemes 71
8.4.3.4.3 Verification/validation 71
8.4.4 Commercially Implemented FE Modeling Tools 72
8.4.4.1 SIMULIA Abaqus 72
8.4.4.2 Ansys Composite PrepPost 73
8.4.4.3 LS-DYNA 73
8.4.4.4 AlphaStar GENOA 74
8.4.4.5 MSC Software 74
8.4.4.6 Autodesk Helius Composite 75
8.4.5 Trends, Challenges and Future Directions 75
8.4.5.1 Round Robin Exercises 75
8.4.5.1.1 WWFE 75
8.4.5.1.2 AQDTA 75
8.4.5.1.3 Composite Materials Handbook-17 76
8.4.5.2 Interaction of Failure Modes 76
8.4.5.3 Uncertainty Quantification 77
8.4.5.4 Multiscale Modeling 78
8.4.5.5 Process Modeling 79
8.4.6 Summary and Conclusions 79
References 80
Relevant Websites 83

Comprehensive Composite Materials II, Volume 8 doi:10.1016/B978-0-12-803581-8.10050-5 61


62 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

8.4.1 Introduction

The linear behavior of simple composite materials and structures can be analytically described by classical laminate theory (CLT).
By considering elastic mechanical properties (Young’s moduli, shear modulus, and Poisson’s ratios) as well as thermal and
moisture effects, CLT is able to effectively calculate stress–strain relations in composite laminates up to failure initiation. Through-
thickness normal and shear stresses are ignored in CLT thus assuming composite plates to be infinitely long and wide which leads
to neglecting edge effects.
Many composite structures sustain loads above their initial strength. CLT combined with first-ply-failure theories is only
capable of describing local failure of the first ply. However, local ply failure does not necessarily lead to immediate global laminate
failure. Damage tolerance is in fact an advantageous characteristic of composite materials and structural design solutions based on
first-ply-failure are generally overly conservative and do not utilize the full potential and weight saving benefits of the advanced
material. In addition, edge effects are commonly observed in, for instance, bolted or riveted composite structures which are not
considered in applying CLT.
Finite element (FE) modeling overcomes these drawbacks and is able to carry out the stress analysis of structures more
accurately. Computational methods and simulations are being routinely used in design and analysis of various composite
materials and structures. A major factor in increasing industrial competitiveness is in the reduction of design cycle time. The ability
to complete designs entirely on the computer reduces the reliance on time consuming and costly physical tests. Composites can be
virtually optimized to suit specific applications before time consuming and expensive manufacturing. Most commercial FE
programs have user-friendly composite modeling capabilities which are able to accurately capture the elastic response through
anisotropic or orthotropic constitutive modeling. In contrast to CLT, FE models enable accurate stress analyses leading to pre-
diction of stress concentrations and, through implementation of appropriate progressive damage/failure models, can predict the
nonlinear inelastic response of composite structures up to ultimate failure.
Precise computational failure analysis is a critical part and an ongoing active area of research for predicting the structural
response of composite materials. Unlike isotropic metallic structures, failure in heterogeneous and anisotropic composites occurs
on multiple scales with complex interaction between various failure modes which significantly increases the numerical complexity.
Coalescence of microcracks in the matrix material can lead to microscopic intra-laminar fiber/matrix debonding. Further growth of
matrix cracks and interaction with intra-laminar debonding results in transverse mesoscopic matrix cracks which eventually lead to
the onset of inter-laminar delamination. Combined with macroscopic fiber damage, such as buckling (kinking in compression) or
breakage (in tension), it becomes obvious that damage evolution in composites is far from being trivial.
In addition, experimental studies show the strong effect of size scaling in composite structures with various configurations and
under different loading scenarios.1,2 It is still an ongoing challenge to use materials (and damage) parameters obtained from time
and cost-effective small-scale tests in order to describe the behavior of full-scale composite structures.
Consequently, numerous computational methods for the structural analysis of composites have been developed over the past
five decades in order to describe the complex nature of composite damage with the goal of establishing general, robust,
repeatable and computationally-efficient techniques. The particular choice of the FE model depends on the purpose of the
analysis.
The aim of this chapter is to provide a comprehensive review and guidelines on a range of FE modeling techniques that are
currently being used for composite materials including damage initiation and evolution. Even though there is a wide range of
composite materials such as ceramic, metallic or polymeric matrix systems reinforced with short, long or woven fibers, the focus of
this chapter will be confined to long or continuous fiber reinforced polymers. Most modeling techniques discussed here are
applicable to other material configurations with slight adjustments.
The layout of this chapter is as follows: Section 8.4.2 provides guidelines for the modeling of composite structures over
different length and time scales. It is followed by the presentation of computational techniques for the introduction of material
damage or failure in FE models. Smeared and discrete damage models are discussed in combination with common difficulties and
challenges associated with their representations. Based on these findings, Section 8.4.4 summarizes the capabilities of commercial
FE software packages in regards to composite modeling and associated damage and failure implementation. Section 8.4.5
addresses current trends and outlines latest developments which are believed to contribute significantly to further improve the
structural analysis of composites within FE models. The last section concludes this chapter and provides a summary of suggested
guidelines to effectively establish nonlinear FE models for composite materials.

8.4.2 FE Modeling Techniques

This section will present guidelines on the choice of elements and spatial as well as temporal discretization of composite structures
within the FE modeling framework. Based on specific structural dimensions and loading scenarios, it is important to know what
element types are suitable and what their drawbacks are.
There are three different element types that are commonly being used:

• Conventional shell elements where only the shell reference surface is discretized and the thickness is defined by section
properties. This element type considers translational and rotational degrees of freedom (DoFs).
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 63

• Continuum or thick shell elements discretize a 3D volume, but the kinematic behavior of the element is based on shell theory.
The thickness is defined by nodal geometries. Only translational DoFs are considered in these types of elements.
• Solid elements discretize a 3D volume and the full state of stress including the out-of-plane (through thickness) stresses can be
computed.

Shell elements reduce the level of computational effort. They are valid for modeling thin structures such as hollow tubes, panels
for airframes, and wind turbine blades. In contrast to conventional thin shell elements, continuum shell elements consider
transverse shear deformation. In plate bending, shear deformation tends to be important when the span-to-thickness ratio is less
than 10:1 for isotropic materials. In composite laminates where the axial (bending) modulus can be an order of magnitude greater
than the transverse shear modulus, shear deformation becomes important even for thin, flexible plates where the above ratio can
be higher than 10:1.
Solid elements are normally used to model thicker structures such as gas turbine blades or stringers. In situations where loads
are applied in the thickness direction, or accurate contact definitions are needed and/or the structure undergoes large deforma-
tions, solid elements are favored over shell elements.
Composites are usually characterized by orthotropic or anisotropic elastic material properties such as Young’s modulus in the
longitudinal and transverse directions, shear moduli, and Poisson’s ratios. Most commercial software packages include built-in
composite modeling capability where elastic properties of angled plies are calculated based on unidirectional properties.
Section 8.4.4 will present more details on commercially available FE software programs.

8.4.2.1 Spatial Discretization


In order to accurately capture the structural response of large-scale composite components, FE models have to consider their
complex nature at each scale of spatial resolution: micro, meso, and macro. Since the computational cost to do so is still too high
for current computers, researchers and engineers typically target one of the above scales to study the damage behavior of the
material/structure. The particular choice of refinement level depends on the purpose of the analysis. In the following, each length
scale is presented in combination with typical applications.

8.4.2.1.1 Microscale – Representative volume element


As illustrated in Fig. 1, microscale models study the details of the interaction between the fiber and matrix at scales up to a few
hundred microns. Fig. 1(a) shows the separate modeling of matrix and reinforcement as deformable continua on the microscale
where each element is composed of single homogeneous set of material properties. Constituents are fiber, matrix and the
fiber–matrix interface.
Microscale FE models have been successfully implemented to predict the mechanical properties of undamaged laminae such as
the elastic moduli or coefficients of thermal expansion.4 Within the framework of damage and fracture modeling, there are two
major approaches. In the first approach, the building block of the model consists of a fiber at its center surrounded by the matrix
material. By assuming a periodic distribution of the reinforcements, this unit-cell approach can be applied to various loading
scenarios.5,6 The second approach takes into account the morphology and distribution of the fibers in the laminae. The repre-
sentative volume element (RVE) includes a number of fibers embedded in the matrix. It can represent up to a few hundred
microns.7,8
It has been shown that microscale models are able to successfully predict the elastic properties of undamaged composite
materials. Although damage or fracture have been successfully applied to RVEs,9 the models are unable to extrapolate the localized
damage all the way up to the complete composite structure. Damage in composites is inherently a nonlocal event and its evolution
depends on the interplay between the material and structural behavior. Consequently, bridging the scales by some averaging
scheme using a bottom-up approach becomes questionable as the characteristic material length, which is linked to the constitutive
model at the macroscale, cannot be built-up from lower scales in a hierarchical manner.10

8.4.2.1.2 Mesoscale – Ply-based modeling


The most popular approach utilized to numerically study composite structures is modeling their response at the ply-level. This also
forms the building block of CLT. Mesoscale models are composed of several layers of elements representing individual composite
plies as illustrated in Fig. 1(b). Material within each ply is defined by homogeneous equivalent properties consisting of matrix and
reinforcement. Layers are assumed to be perfectly bonded unless interface elements are inserted in between plies to model
delamination as explained in Section 8.4.3.2.1.
Mesoscale FE models have been extensively applied to composite structures under various loading conditions ranging from
quasi-static open-hole tensions tests11 to dynamic impact events.12–15
Ply-based models assume that the structural behavior within each ply is independent of its neighbors. Hence the interaction of
damage mechanisms is generally neglected although it is widely accepted that the orientation of neighboring plies plays a major
role in the structural response of a lamina. One method to account for such drawback is to experimentally measure and
computationally implement in-situ strength properties.16 Another method is to allow for interacting failure modes by sophisti-
cated FE modeling such as delamination migration17 or matrix cracking induced delamination (MCID).18–20
64 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

Fig. 1 Illustration of finite element (FE) modeling techniques of composite materials at the (a) microscale, (b) mesoscale, and (c) sublaminate
macroscale.3

Furthermore, a series of coordinated round-robin studies known as the World Wide Failure Exercise (WWFE) has been
conducted to identify reliable and efficient FE tools to predict damage progression in composite materials. WWFE is currently in its
third edition.21 Most of the proposed models are formulated on the mesoscale.
The results of WWFE I and II concluded that only a few models give acceptable correlation (within 50%) with test data for 75% of
the test cases used.21 In addition, it appears that there are no failure or damage criterion universally accepted by designers as being
adequate under general loading conditions.22 The main reason for this outcome is the strong scaling effect in composites mentioned
in Section 8.4.1. The structural response of bidirectional or multidirectional composite laminates cannot solely be described by the
mechanical behavior of unidirectional laminae. The quasi-brittle nature of composites implies that their general nonlinear response
to loads is size dependent and therefore cannot be scaled up from the knowledge of their small scale (coupon) behavior.

8.4.2.1.3 Macroscale modeling


Composites can also be modeled as an equivalent homogeneous material with stacked or single layer element configuration as
illustrated in Fig. 1(c). Macroscale models are used for the simulation of large-scale structures. Details such as accurate damage
representation and failure interaction in the layers are neglected. The overall response is governed by macroscale parameters.
In order to describe the previously mentioned scaling effect of composite structures, sublaminate-based FE models can be
applied. Initially introduced by Williams,23 this COMposite DAmage Model (CODAM) takes the sublaminate as the representative
volume and as the base level for introducing damage. The model is constructed at the scale of a block of layers which accounts for
the in-situ mechanical properties of adjacent plies due to damage. Damage causes an overall nonlinear response in regards to
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 65

stiffness, load-carrying capacity or post-peak behavior. In sublaminate-based models, damage is implemented in a smeared
manner without capturing the details of single damage events in the plies.24,25
The challenge associated with the sublaminate modeling is the calibration of the macroscopic material properties. Damage of
the sublaminate depends on orientation, sequence and thickness of its constituent layers. Each variant of the stacking sequence
must be regarded as a distinct material for which the material model assigned to the sublaminate needs to be recalibrated.26,27,124
Another problem of modeling large-scale structures on the macroscale is the level of discretization. For instance, computa-
tionally expensive fine meshing is needed to accurately simulate inter-laminar delamination. More details on interface modeling
are outlined in Section 8.4.3.2.1.

8.4.2.1.4 Multiscale modeling


As seen in previous sections, each modeling approach on different length scales has its own advantages and drawbacks. The
multiscale nature of composite structures has provided the impetus to establish FE models that contain various length scales and
to combine their benefits. Multiscale models can be categorized into two groups: hierarchical and concurrent.
Hierarchical models apply homogenization techniques to transfer the predicted behavior from a lower scale (e.g., microscale
RVE) to a higher length scale.28–30 Macroscopic behavior is controlled by the underlying micromechanical constitutive relation to
govern stress–strain behavior applied to an integration point of a macroscale FE model. They are able to predict the macroscopic
elastic response of a material consisting of complex fiber architectures on the microscale. However, hierarchical modeling is not
able to extrapolate damage or failure since it is based on homogenization.
In contrast to applied homogenization to bridge the scales, concurrent multiscale models increase model resolution in
damaged areas where meso-micro features are explicitly accounted for. An adaptive multilevel method has been successfully
applied to predict matrix cracking in composite laminates.31,32
The immense computational cost associated with the multiscale models has hindered their adoption in practical applications.
In addition, it is challenging to accurately characterize mechanical properties on small-scale models. Nevertheless, the idea of
incorporating several length scales in FE models is still believed to significantly contribute to the improvement of structural
analyses of composites including damage or failure.

8.4.2.2 Temporal Discretization


A similar approach to the spatial discretization is required to resolve the model in the time domain. The choice of a specific
numerical solution technique depends on the applied strain rate. Static loading scenarios are solved differently than high speed
dynamic impact problems. The temporal integration is carried out using either an implicit or explicit solution technique.

8.4.2.2.1 Implicit scheme


Implicit methods attempt to find a solution to the nonlinear system of equations iteratively by considering the current state of the
system as well as its subsequent (or previous) time state. Newton-Raphson method or its modified forms are used to enforce
equilibrium of the internal restoring forces with the externally applied loads. These incremental-iterative procedures are uncon-
ditionally stable which allows for using large increments. Hence, large incremental time steps can be applied in implicit schemes
while maintaining numerical accuracy. However, the application of Newton-Raphson method or other iterative schemes within
each time step increases the computational cost. While the approach works well for a broad range of problems, in severely
nonlinear cases, it may fail to converge.

8.4.2.2.2 Explicit scheme


In contrast to implicit methods, explicit solution techniques are only conditionally stable requiring very small time steps to obtain
a numerically stable solution. The critical time increment is governed by the highest frequency in the system which is generally
associated with the smallest element size as well as material properties that influence the wave speed. However, the solution is only
based on current quantities and hence a computationally costly iterative solution technique is not required. Since iterative solution
of a system of equations is not involved, the explicit approach is well suited for problems involving complex contact conditions
and severe nonlinearities.
The preferred solution scheme is the implicit method since it allows for large time increments and it is not sensitive to small
element sizes. In static or quasi-static problems, implicit solution techniques are the optimal choice since large time increments
can be applied. Furthermore, higher-order elements are more readily available in implicit solvers while explicit methods preclude
their use. However, depending on the applied strain rate, there are cases where explicit solving schemes are favorable. In high-
speed dynamic problems where small time increments are required and the total simulation time is short, an explicit scheme is
more efficient since it does not involve any iterative techniques. Examples of such high-speed problems are simulations involving
wave propagation, as well as high-speed impact/crash and blast events. In addition, severely nonlinear problems that involve
complex constitutive response, typically strain-softening material models, discontinuities, contact, and so forth lend themselves to
intensive computations (usually on massively parallel systems) using explicit solvers.
A large number of dynamic problems cannot be fully classified as either low-speed (implicit) or high-speed (explicit) such as
crash or drop-weight test simulations. Both solution methods can be applied in such cases.
66 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

8.4.3 Computational Damage and Failure Modeling

As outlined in Section 8.4.1, failure in composite structures occurs on multiple scales with various interactions among the modes
of failure. This section presents the most common techniques to numerically predict damage initiation and progression in
composite materials. Two distinct modeling approaches exist: continuum damage mechanics (CDM) and discrete crack simula-
tion. The former represents damage by means of smeared local stiffness reduction within the constitutive model whereas the latter
allows for simulating crack propagation through explicit representation of the crack geometry. The specific modeling techniques
are summarized in Fig. 2.

8.4.3.1 CDM
The basic idea of CDM is to replace the mechanical properties of a damaged material in the fracture process zone with those of an
equivalent homogeneous material by associating damage mechanisms with their overall effects on the mechanical material
parameters. Among the various approaches to model fracture and damage in solids, CDM have gained the most attention.4 This is
mainly due to the convenience of their implementation in general purpose FE software programs.
CDM originates from creep rupture analysis in homogeneous metals which was first studied in the late 1950s.33 Since then,
extensive research has led to numerous failure criteria to describe strength reduction of metals as a function of micromechanical
damage evolution. However, experimental observations of failure in composite materials increasingly put to question the merit of
such analysis in heterogeneous structures. Comprehensive literature reviews on damage modeling in composite laminates can be
found in Refs.34–37
In order to represent the failure types in composite laminates macroscopically, the loss of stiffness in one dimension can be
described by a phenomenological damage parameter D with 0rDr1 such that
s ¼ ð1  DÞCe ¼ Ce
where s is the Cauchy stress and e is the strain, C the initial (undamaged) stiffness and C ¼ ð1  DÞC the effective stiffness due to
damage. D¼0 represents undamaged material, whereas D¼1 denotes complete (saturated) damage. In multidimensional state of
stress, different damage parameters are applied in different directions to account for the orthotropic nature of damage in
composites.
This approach requires homogenization of a damaged structure and application of effective material properties such as C. Fig. 3
illustrates the method by smearing out voids and cracks in heterogeneous material and using effective material properties in a
homogenized structure.

Fig. 2 Classification of computational damage and fracture modeling techniques related to different length scales.

Fig. 3 Schematic description of continuum damage mechanics (CDM) by homogenization of cracked structure. Adapted from Voyiadjis, G.Z.,
Kattan, P.I., Yousef, M.A., 2014. Some basic issues of isotropic and anisotropic continuum damage mechanics. In: Voyiadjis Z.G. (Ed.), Handbook
of Damage Mechanics: Nano to Macro Scale for Materials and Structures. New York, NY: Springer New York. pp. 1–37.
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 67

It is important to describe initiation of failure and damage evolution to accurately formulate stiffness reduction. Over the past
five decades, countless failure criteria on the initial failure of composite laminates have been developed. The following section
presents the most successful and promising criteria.

8.4.3.1.1 Damage initiation


Strength-based failure criteria relate internal stresses and experimental measures of material strength to the onset of failure in
composites.38 The existing criteria can be classified as phenomenological or global, and non-interactive (independent) or inter-
active. Phenomenological criteria describe each failure mode in composites individually whereas global criteria do not distinguish
between different failure modes. Non-interactive approaches only compare individual stress or strain components with the
corresponding allowable strength values. In contrast, interactive failure criteria assume mode interaction to account for the
complex nature of failure in composites.
Some of the most popular criteria are: (i) non-interactive criteria, for example, (1) maximum stress and (2) maximum strain
criteria; (ii) global interactive criteria, for example, (3) Tsai-Hill,39,40 (4) Tsai-Wu,41 (5) Hoffman,42 (6) Yamada-Sun43; and
(iii) phenomenological interactive criteria which distinguish between fiber and matrix failure in tension and compression
respectively, for example, (7) Hashin,44 (8) Chang and Chang,45 (9) Puck,46 (10) Cuntze,47 (11) LaRC03,48,49 and (12)
SIFT.50

8.4.3.1.2 Damage evolution


As stated in Ref. [11] even an accurate failure criterion for elementary plies is not sufficient for the prediction of laminate failure.
Local failure of a ply does not necessarily lead to immediate global laminate failure. Damage tolerance is in fact an advantageous
characteristic of composite materials and structural design solutions based on damage initiation are generally overly conservative
and do not use the full potential of the advanced material. Increased loads from local failure are normally redistributed
to the remaining undamaged material. Hence, a progressive failure analysis (PFA) is required to properly characterise the
residual stiffness of the composite laminate. Failure criteria need to be extended with a theory on what happens after local loss of
integrity.
Fig. 4 illustrates stiffness degradation where E and Ed refer to undamaged and damaged (degraded) Young’s modulus,
respectively, strain at damage initiation, e1, strain at ultimate failure, e2, as well as degradation factor, df.
Either all or only selected elastic ply properties are degraded. For instance, ply properties such as the modulus perpendicular to
the fibers or the in-plane shear modulus are reduced in order to represent damage in the matrix material.

8.4.3.1.2.1 Sudden degradation models


Stiffness properties are reduced instantaneously to some fraction of the undamaged properties (path OBCD in Fig. 4). Early
application of this method, which is referred to as the ply discount method, was adopted by Chang and Chang.45

8.4.3.1.2.2 Gradual degradation models


Constitutive models with gradual softening laws follow the path OBD in Fig. 4. The nonlinear path BD is a function of some
evolving field variable. Initial formulations for anisotropic composite structures date back to early 1990's.52,53
A typical example for the nonlinear degradation as seen in Fig. 4 is given by Eduardo and Griffin54 where the longitudinal and
shear moduli were degraded due to fiber tensile failure using a degradation factor d1 such that

 b !
s
d1 ¼ exp 
s0

where s and s0 are current and initial (undamaged) stresses and b is a shape parameter. Rapid degradation is described by high
values of b whereas lower values represent gradual degradation.

Fig. 4 Illustration of degradation models.51


68 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

8.4.3.2 Discrete Failure Representation


One challenge with respect to the failure criteria in Section 8.4.3.1 was identified to be a physically accurate representation of
matrix cracks in composite laminates. Fig. 5 shows that homogenization in CDM can lead to nonphysical crack propagation.
A discrete crack representation can resolve this drawback by explicitly incorporating cracks in FE meshes. Fig. 6 compares CDM
with explicit discrete matrix cracking.11 It can be seen that discrete crack representation in FEM can lead to more accurate
prediction. In addition, the discrete method enables the modeling of failure interaction in composite laminates such as the
coupling of matrix cracks and intra- or inter-laminar damage. In order to avoid computationally expensive remeshing, a prescribed
crack path or adaptive mesh-independent enrichment methods based on partition of unity are required.
Delamination or debonding is commonly modeled by prescribed crack paths as the location of failure is known before simu-
lation. However, matrix cracks can occur anywhere in the matrix and hence enrichment methods as described in Section 8.4.3.2.2 are
used to incorporate explicit crack growth in FE meshes without a priori knowledge about crack location and propagation.

8.4.3.2.1 Cohesive Zone Method and Virtual Crack Closure Technique to model delamination
Strength-based failure criteria are not able to accurately predict delamination in composite laminates.22 An alternative method to
strength-based failure criteria in Section 8.4.3.1.1 is to use fracture energy-based criteria to predict the delamination onset and
propagation. Two of the most common and successful methods are presented in the following:

8.4.3.2.1.1 Virtual crack closure technique


Initially formulated by Rybicki and Kanninen56 to evaluate stress intensity factor in double cantilever beam (DCB) samples, virtual
crack closure technique (VCCT) has become a popular tool to computationally predict delamination. It assumes that the dis-
sipated energy due to crack growth is identical to the work required to close the crack.
Advantages of the VCCT method are36 that it is: (1) relatively simple to use, (2) possible to calculate fracture parameters at the
crack tip under mixed mode loading, (3) applicable to 3D cracks, and (4) not too demanding on mesh density requirement.
A major drawback of VCCT is that its FE implementation requires the presence of an initial crack prior to the analysis. This
makes the method useful for cases where the exact location of the delamination front is explicitly known. Since VCCT is based on
linear elastic fracture mechanics (LEFM), it is further limited to cases where the size of the fracture process zone is negligibly small
compared to other structural dimensions.

8.4.3.2.1.2 Cohesive zone method


The cohesive zone method (CZM) overcomes the previously mentioned drawbacks of VCCT. Dugdale and Barenblatt57,58
introduced a phenomenological description of discrete fracture as a material separation across a surface. CZM assumes a process
zone in front of the crack tip. Fracture properties are determined by a cohesive traction-separation relation of upper and lower
crack surfaces in the process (cohesive) zone. This nonlinear relation is mostly governed by experimentally determined strain

Fig. 5 Illustration of typical crack growth in homogeneous material and in fiber reinforced composites (FRC).55

Fig. 6 Comparison of peak load values of continuum damage mechanics (CDM) and discrete matrix cracking to experimental open hole test
results of composite laminates.11
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 69

energy release rates. Applied traction-separation laws depend on material and application. The most common forms are bilinear or
exponential relations.59
CZM is well established and a common tool for numerical nonlinear fracture modeling, primarily to model inter-laminar
delamination in composite laminates.60–62 A detailed overview of different implementation strategies and applications can be
found in Ref. [59]. A powerful mixed-mode cohesive formulation is presented by Camanho et al.63
In FEM, CZM can be implemented as cohesive surfaces (similar to VCCT) or as cohesive elements with minimal (or zero)
thickness. The main advantage of CZM is the capacity to model both the onset and propagation of fracture in the same analysis35
by defining a traction-separation law based on experimentally obtained strain energy release rates.
A drawback of CZM is their high computational cost since in explicit solutions, when incorporated into elements, their small
thickness results in very small time step requirement in order to satisfy the Courant stability condition for numerical time
integration.14 In addition, a fine FE mesh is needed to prevent numerical issues and to achieve physically meaningful results, see
Section 8.4.3.4. A maximum characteristic length is obtained by relating the elastic energy of the cohesive element with the energy
dissipated by fracture.14,64 Recently, a novel computational method has been developed to adaptively insert cohesive elements
within regions where inter-laminar delamination has the potential to initiate. This so-called Local Cohesive Zone method (LCZ) is
verified and validated against composite structures in Modes I, II and Mixed-Mode tests65 as well as under axial crushing and
transverse impact loading.66

8.4.3.2.2 Partition of unity for matrix cracking


Based on the partition of unity method,67 additional DoFs are used to numerically describe crack propagation in FE meshes. The
mesh-independent X-FEM method was developed by Moës et al.68 and has become a widely used tool for arbitrary modeling of
fracture with arbitrary orientations. In the context of matrix cracking in composite laminates, X-FEM has been applied by using
higher order shape functions by Iarve.69,70
A mathematically equivalent method to X-FEM is the phantom node method (PNM) initially formulated by Hansbo and
Hansbo.71 In contrast to X-FEM, PNM is a local method only operating at the element level and only using standard shape
functions. X-FEM requires transition elements around crack tips to model displacement jumps. The locality makes PNM con-
ceptually simpler and hence easier to implement and modify.
PNM has been applied to model open hole tension tests,11 progressive splitting72 or delamination migration.17 In Fig. 7, failure
interaction of matrix cracks and inter-laminar delamination under flexural four-point bending was successfully modeled by
combining PNM with breakable CZM interface elements.19

8.4.3.3 Calibration of Model Parameters


8.4.3.3.1 Material parameters
8.4.3.3.1.1 Intra-laminar modeling
The application of failure criteria and degradation models in CDM requires extensive experimental studies to obtain material
constants such as strength values in the fiber and transverse direction.73 Even though most of these values are easily determined by
standard tensile testing of unidirectional composites, some parameters such as the interaction terms in the Tsai-Wu failure criterion
require difficult biaxial testing.37 An additional challenge is the inherent uncertainty of strength characteristics of composite
materials. Different material characterization tests are required depending on whether the mode of loading is tensile or com-
pressive. All models require standard material characterization tests to extract the elastic properties; however, depending on the
specific model being used, some require customized test configurations to quantify the strength and fracture properties. The
amount of expensive and time-consuming experimental calibration can be significantly reduced by efficient virtual testing.

Fig. 7 Matrix cracking induced delamination (MCID) in composite laminates under flexural loading. Left: Comparison of experimentally measured
and numerically predicted local delamination ratio (LDR) induced by matrix cracks. Right: qualitative inspection of MCID.19
70 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

McGregor et al.125 successfully combine experimental and numerical calibration methodologies to establish a macro-mechanical
CDM model applied to axial crushing of braided composite tubes.

8.4.3.3.1.2 Inter-laminar modeling


Application of VCCT or CZM requires accurate quantification of strain energy release rates (fracture energy) Gc in Mode I
(opening), Mode II (in-plane shear), and Mode III (out-of-plane shear). Gc is identified as the work needed to create a unit area of
a fully developed crack.74 Sophisticated tests and measurement techniques are necessary to ensure accurate and reliable interface
quantifications in composite laminates. DCB and end-notched flexure (ENF) tests are commonly used to determine Gc in Mode I
and Mode II, respectively.75–77 Associated challenges are accurate measurement of crack propagation and prevention of stiffening
effects such as fiber bridging which leads to significantly higher fracture energies.
Another challenge associated with CZM is the determination of mixed mode parameters. For instance, in order to apply the
popular B-K mixed mode cohesive relation in Ref. [63] the interaction of Mode I and Mode II interface failure needs to be
determined according to Benzeggagh and Kenane.78

8.4.3.3.2 Numerical parameters


In addition to experimentally determined values, numerical parameters are often required in order to guarantee efficient and
robust FE models. Depending on the choice of the time integration scheme, whether implicit or explicit, different numerical
parameters may be required. For instance, the artificial damping factor in viscous regularization discussed in Section 8.4.3.4.2
should be chosen to be close to the minimum incremental time step in order to stabilize convergence and obtain physically
meaningful results.
In CZM, the choice of specific traction-separation laws such as initial stiffness or cohesive shape needs to be determined by the
user. The choice of the initial stiffness can be determined by relating the through-thickness Young’s modulus and the thickness of a
sublaminate according to Turon et al.79 As previously mentioned in Section 8.4.3.2.1 on CZM, the traction-separation relation is
governed by experimentally determined strain energy release rates. The area under the curve corresponds to these energies.
However, the shape of the softening curve can be chosen by the user. Most common shapes are a linear or exponential decrease of
traction between the two cohesive surfaces. Depending on the application, other shapes are favorable. A detailed overview on
shapes, material parameters and recommended use can be found in Ref. [59].
In implicit models, solution control parameters can significantly improve numerical performance. Depending on the problem, the
optimal choice of the minimum and maximum incremental time step can speed up the calculation. In the context of VCCT
simulations, an automated algorithm can instantly increase the incremental time steps to reach the point of interest quickly (e.g.,
delamination initiation). This is known as linear scaling. Another control parameter is the user-defined tolerance to enforce equili-
brium of the internal structural forces with the externally applied loads. This can help to improve convergence; however, care has to be
taken in defining the tolerance to maintain physically meaningful results. It is recommended to use the default settings in FE solvers.
As mentioned in the context of explicit schemes, very small time steps are required to obtain an accurate and numerically
stable solution. The choice of the critical time step is crucial and has to be critically analyzed. For non-dynamic problems, usually
mass scaling is used to increase the critical time step and hence make the run times more reasonable. Further efficiencies can be
gained by increasing the rate of loading reasonably and as long as the dynamic effects associated with such increases do not violate
the balance of energies.

8.4.3.4 Numerical Issues


PFA in FEM can cause numerical convergence problems due to the sudden stiffness reduction in CDM constitutive models,
nonlinear traction-separation relation at cohesive surfaces or due to incorporating discrete cracks in X-FEM or PNM. In the
following, common problems and challenges stemming from damage modeling in composite laminates are discussed.

8.4.3.4.1 Mesh-dependency
In CDM models, softening can lead to the local loss of positive definiteness of the material tangent stiffness matrix.80 The
mathematical formulation becomes ill-posed and numerical solutions are dependent on the FE mesh size and alignment,81 i.e.,
the amount of dissipated energy due to crack evolution vanishes upon mesh refinement.11,64 Nonlocal models are generally
required to overcome this problem. In these models, the local stress at an integration point not only depends on strain and other
state variables at that particular point, but also on the strains and state variables of the points in the surrounding area.82 The
constitutive behavior ceases to be independent of its neighborhood and local quantities are replaced by averaged values within a
certain prescribed radius of influence. This radius can be interpreted as a new material property, generally referred to as an internal
length that dictates the size of the damage zone.
Two of the most common regularization techniques to mitigate the effect of mesh-dependency in the structural analysis of
composite structures are presented in the following:

• Integral models
This method is based on spatial averaging of the state variables, typically strains, in a finite neighborhood of a certain point.80
A popular integral model to resolve mesh-dependency is the implementation of the crack band method which was developed
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 71

to describe fracture of heterogeneous concrete by Bažant and Oh64 and first applied to progressive damage modeling of
composite materials by Williams et al.23 In conjunction with critical energy release rates, the method introduces mesh-
independent damage evolution along a band defined by an internal length parameter. This allows for modeling of complex
loading scenarios such as tension-compression reversals. Another example is CODAM2, a nonlinear regularization scheme
implemented in the commercial FE software, LS-DYNA.4 The major drawback of integral models is their implementation in
commercial FE codes which necessitates access to information stored in neighboring elements to the element (or integration
point) in which the constitutive relation is being evaluated. Typically, the only accessible modules in commercial FE codes for
customization of material models are user material modules (UMAT) which are strain-driven routines that compute the
appropriate stresses at a given integration point (i.e., locally).4,25,83–85
• Gradient based models
In gradient damage models, the constitutive relation is enhanced by incorporation of strain gradients or by the introduction of
both strain gradients and their stress conjugates. The advantage over previously mentioned integral models is the fact that
gradient-based models are strictly local in a mathematical sense.86 Hence such models are easier to implement in commercial
codes (in contrast to the integral models which require information from neighboring elements) and computationally less
expensive.87–89

8.4.3.4.2 Convergence in implicit schemes


Incorporating multiple failure modes in FE analyses can cause ill-conditioning of the system of equations. Failed elements
with corresponding stiffness degradation in CDM lead to small eigenvalues which approach zero. Furthermore, failure
can cause collapse behavior where the load–displacement response shows negative stiffnesses or eigenvalues36 such as
snap-through and snap-back effects in the load–displacement response. Conventional Newton-Raphson algorithms in implicit
schemes are not able to compute such effects and hence adjustments are needed to resolve these issues, which are briefly
discussed below.

• Adaptable mass
An adaptable mass matrix can be introduced in a dynamic solution algorithm to prevent singularities in the iterative Newton-
Raphson scheme. Care has to be taken to guarantee that the artificial inertia effects in static nonlinear problems are small
enough.
• Viscous regularization
This method is used to damp sharp load drops due to sudden element failure or steep nonlinearities introduced in CZM,
X-FEM, or PNM. Viscous regularization can be applied to the entire FE problem (in step or solver definition) or directly applied
to the progressive damage evolution.90 Viscous parameters need to be small enough (approximately close to minimum
incremental time step) to stabilize convergence without introducing artificial damping effects.
• Arc-length algorithm
Another method to deal with convergence problems in damage analysis in composite laminates is to adjust the conventional
Newton-Raphson algorithm. Arc-length methods by Riks,91 Ramm,92 and Crisfield93 not only iterate through displacement, but
they also scale the load in the global load–displacement response to ensure accurate solution. These approaches ensure highest
solution precision. However, calculations are relatively time-consuming.36

8.4.3.4.3 Verification/validation
In order to develop and produce reliable numerical simulations in computational mechanics and physics, verification and
validation (V&V) processes are required.94
Verification addresses the accuracy of the numerical solution as compared to a known exact solution of the underlying
conceptual model by code verification and numerical error estimation. These errors can arise from spatial and temporal dis-
cretization; or iterative solution methods to mathematically describe a nonlinear partial differential equation (PDE) including
singularities, discontinuities, and complex geometries.95
In the validation process, the conceptual model is compared to physical events of a “real world” problem96 by quantitative
comparison between computational and experimental results as well as determination of whether there is an acceptable agreement
between the model and the experiment.95 Consequently, verification should be completed before model validation.
In regards to the analysis of composite structures by FE models, verification is required when novel theories or FE
implementation techniques are used. The validation against materials and structural test data is an important and challenging
aspect of developing computational models to predict the structural response of composite materials. The process can be
based on the building block approach where knowledge on the FE model is built step-by-step from the coupon level up to
the large-scale and configured composite structures. Predictions obtained by numerical simulations at each level of the
pyramid (building block) are compared with physical test data to determine the accuracy and confidence level of the computa-
tional tool. As mentioned in the introduction, it is still an ongoing challenge to describe the behavior of full scale composite
structures through numerical models due to their complex heterogeneous and anisotropic nature. A validation metric based on
the statistical concept of confidence intervals as suggested in Ref. [95] considers experimental uncertainty and can be used to
capture the overall trend in predicting the structural response of composite materials rather than achieving pointwise reproduction
of test data.
72 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

8.4.4 Commercially Implemented FE Modeling Tools

The following section evaluates capabilities of commercially available software within built-in structural FE modeling tools of
composites, in particular, damage and fracture modeling tools. Although there is a wide range of available FE programs, this
section focuses on the most popular software packages. Additional information on commercial FE simulation tools can be found
in Ref. [97].
As previously mentioned in Section 8.4.2, built-in or add-on composite modeling tools are available in all FE software
packages. This includes in particular, different types of shell and solid element formulations, constitutive modeling on any length
scale as well as incorporating implicit and explicit solution procedures in combination with some built-in stabilization methods as
outlined in Section 8.4.3.4. The software tools differ in damage and failure modeling capacity.
In most academic publications on damage modeling using commercial software, user-defined codes are implemented to apply
specific failure criteria and evolution by modifying constitutive material behavior. The capability to integrate user-defined tools is
only briefly mentioned since implementation can be very specific and complex.
Optimization tools are finding increasing interest in modeling of composite structures, for example, to maximize the specific
load carrying capacity of a structure by optimizing the stacking sequence of the laminate or by changing local geometry to reduce
the maximum principal stress. These tools are available in all software packages below. As the current focus is on damage modeling
capabilities, no further details on optimization are discussed.
One of the biggest challenges in modeling composites is the variability in material properties. This is caused by the complex
multiscale morphology of composites, introduced uncertainty in manufacturing processes and inconsistent testing methods.
Instead of the traditional method of applying high safety coefficients (knock-down factors on strength), probabilistic modeling of
composite structures can be used to design less conservative structures. Some of the commercial FE software packages contain
probabilistic modeling capability as discussed below.
This review studies availability of standard FE modeling tools outlined in Section 8.4.2, failure criteria mentioned in
Section 8.4.3.1.1, damage progression in Section 8.4.3.1.2 as well as capabilities to incorporate discrete failure in Section 8.4.3.2.
Table 1 at the end of the section summarizes the findings outlined below.

8.4.4.1 SIMULIA Abaqus


Abaqus is part of SIMULIA family of codes which is a multiphysics modeling and simulation software. Other subprograms for
instance are Isight and Tosca for optimization or fe-safe for advanced fatigue analysis. Abaqus Standard is used for problems
solved by implicit schemes and Abaqus Explicit for high speed dynamic problems. As one of the major commercial FE software
programs, Abaqus is compatible with a lot of other in-house or commercially available FE codes.
Damage initiation can be controlled by maximum stress/strain criterion or two-dimensional Hashin’s criteria. Material
degradation due to damage is governed by linear or exponential softening as illustrated in Fig. 4.
CZM is available in the form of cohesive surfaces or elements in two- and three-dimensional applications. Viscous regular-
ization can be directly added to cohesive formulation in order to improve convergence. Creating zero-thickness cohesive elements
is significantly simplified in the latest 2016 version of Abaqus. VCCT is implemented as a special crack feature “Debond using
VCCT” in combination with several linear, bilinear, and mixed-mode fracture criteria. Another implemented crack feature in
implicit applications is X-FEM. A pre-crack (if existing) and a potentially cracked domain need to be defined by the user a priori. It
can be coupled to thermo-mechanical problems. Comprehensive tutorials can be found here.
User-defined constitutive models can be implemented in C or FORTRAN through USDFLD or UMAT subroutines for implicit
solvers and VUSDFLD and VUMAT in explicit problems.

Table 1 Summary of commercial finite element method (FEM) software package and their damage modeling capabilities. CDM criteria refer to
Section 8.4.3.1.1

Failure mode Delamination Ply failure Discrete cracks

Numerical method CZM VCCT CDM X-FEM PNM Statistics Multiscale


Global Phenomenological

1 2 3 4 5 6 7 8 9 10 11 12

Abaqus ☑ ☑ ☑ ☑ ☑ ☑
Ansys ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑
LS-DYNA ☑ ☑ ☑ ☑ ☑ ☑ ☑
AlphaStar GENOA ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑
MSC ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑ ☑

Source: CDM criteria: (1) Maximum Stress, (2) Maximum Strain, (3) Tsai-Hill, (4) Tsai-Wu, (5) Hoffman, (6) Yamada-Sun, (7) Hashin, (8) Chang & Chang, (9) Puck, (10) Cuntze,
(11) LaRC03, (12) SIFT
CDM, continuum damage mechanics; CZM, cohesive zone method; PNM, phantom node method; VCCT, virtual crack closure technique; X-FEM, extended finite element method.
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 73

Fig. 8 Screenshot of probabilistic parameter definition using Ansys Six Sigma.

As mentioned in Section 8.4.3.4.2, arc-length methods can significantly improve convergence. A built-in RIKS solution scheme
is available in Abaqus which can be simply used by changing the solver settings.

8.4.4.2 Ansys Composite PrepPost


Ansys covers a wide range of products to simulate multiphysics problems. It mainly consists of subprograms to model fluids,
structures or electronics. All subprograms can be called and combined in one platform Ansys Workbench. One of those sub-
programs is ANSYS Composite PrepPost which will be referred to in the following.
Various failure criteria are available: maximum stress/strain, Tsai-Wu, Tsai-Hill, Hashin, LaRC, Cuntze and Puck. Degradation is
coupled with the Chang and Chang softening model.45
CZM is available with linear and exponential softening where the traction-separation law can be defined in terms of maximum
allowable displacement jump across the interface or critical strain energy release rate. Ansys includes the same VCCT capabilities as
Abaqus.
Material behavior or modified failure criteria can be implemented in user-programmable features (UPFs) written in C or
FORTRAN.
One powerful Ansys tool to effectively model composite structures is Six Sigma – a probabilistic input tool for mechanical
properties to account for material variability. Fig. 8 shows an example on how to incorporate the statistical variation in Six Sigma
into Ansys Workbench.

8.4.4.3 LS-DYNA
LS-DYNA is available both as a stand-alone explicit FE code and as part of Ansys. Originally developed by John Hallquist at
Lawrence Livermore National Laboratory in 1976, and later commercialized by Livermore Software Technology Corporation
(LSTC) in the late 1980s, LS-DYNA is commonly applied to highly nonlinear transient FE problems. In the following, only features
relevant to damage initiation and propagation of composites are outlined. As part of Ansys, previous characteristics covered in
Section 8.4.4.2 can be coupled with LS-DYNA.
Initiation and progression of damage in composites is available through Chang and Chang’s failure criteria with associated
strength and stiffness degradation.45 In addition, other failure criteria such as Tsai-Wu or maximum strain criterion are also
available. Strain rate effects can additionally be considered. LS-DYNA offers a comprehensive database of built-in material models
for damage simulation in composite materials. The material library is frequently updated in order to provide most recent
developments such as the nonlocal sub-laminate based CDM approach (MAT219) or orthotropic continuum damage models
(MAT261 and 262). A complete list with a brief explanation of each available material model can be found here. More details on
74 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

Fig. 9 Multiscale modeling for composite structures in GENOA.98

composite modeling in LS-DYNA are available in S. Hartman’s presentation here. Similar to Abaqus, X-FEM is also available as a
built-in tool in LS-DYNA.
Cohesive surfaces/elements are available in combination with mixed mode or trilinear traction-separation law. VCCT imple-
mentation follows previous descriptions in other software packages.
Similar to the previously discussed FE software packages, user-defined constitutive relations or boundary conditions can be
implemented in LS-DYNA through FORTRAN based codes. For instance, subroutines to implement structural materials are called
usrmat, user-defined loading is known as loadud or user-defined adaptive methods as useradap. LSTC also offers specialised
features developed for the automotive and aerospace industries such as specific tools for crashworthiness simulation, blade
containment or bird strike analysis.

8.4.4.4 AlphaStar GENOA


AlphaStar GENOA is designed for advanced multiscale PFA of composites.
GENOA supports the following strength-based failure criteria: maximum stress/strain, Tsai-Wu, Tsai-Hill, Puck, Hoffman and
Hashin. Various specific micromechanical degradation approaches are available to model matrix defects or residual stresses due to
curing and manufacturing effects. Fiber failure in tension and compression can be modeled as a gradual “rope effect” with
probabilistic Weibull strength distribution.
Similar to Abaqus and Ansys, standard CZM and VCCT features are implemented in GENOA. In addition, a built-in tool to
incorporate interface mechanics such as fiber bridging is available.
User-defined constitutive models in FORTRAN, C/C þ þ , or PYTHON subroutines.
As illustrated in Fig. 9, AlphaStar GENOA has multiscale capabilities which allows for implementing micromechanical effects to
solve problems on the meso or macroscle effectively. In addition, probabilistic property input is available similar to Six Sigma in Ansys.

8.4.4.5 MSC Software


MSC Software includes the following packages.
Patran: Pre-processing modeler
Marc: Advanced nonlinear simulation
Nastran: Post-processing structural analysis
In Patran/Nastran, damage initiation can be determined by maximum stress/strain, Hill, Hoffman, Tsai-Wu, Hashin and Puck
criteria. There are two different progressive degradation options: immediate or gradual, similar to what was discussed in
Section 8.4.3.1.2.
CZM and VCCT follow previous findings in other commercial software packages. Initiation and progression of interface failure
along predefined paths can be described by bilinear, exponential or linear-exponential formulations. Instead of applying X-FEM to
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 75

incorporate explicit cracking without a priori knowledge of its path, the package Marc can be used to model crack propagation
through VCCT and computationally expensive global remeshing.
User-defined failure criteria can be incorporated through UFAIL subroutines.
Digimat is able to account for multiscale material and structural modeling. Coupled to Marc, it is possible to link micro-
mechanical properties to their macroscopic performance. Note that Digimat can also be used in combination with Abaqus, Ansys,
LS-DYNA, and Radioss.

8.4.4.6 Autodesk Helius Composite


Autodesk helius composite is a standalone package that offers a collection of analytical tools for conceptual design of composite
materials. Users can select from an expandable library of materials to build lamina and laminates, and then calculate their material
properties using fundamental tools such as micromechanics and Composite Laminate Theory.
Helius PFA is an add-on for commercial finite element analysis (FEA) programs and provides multiscale simulation cap-
abilities. It is compatible with Abaqus, Ansys and MSC products, multiple failure criteria are available to enable users to view
damage effects in the fiber and matrix and identify multiple modes of failure. It supports multiple analysis conditions, including
dynamic and fatigue loading in combination with multiple material types including chopped and continuous fibers, unidirec-
tional lamina and various woven fabrics.
Specific capabilities in regards to CDM, discrete failure implementation and additional features in Autodesk Helius products
have not been fully explored as part of this review.
Other commercial software packages worth mentioning are Altair Hyperworks with the efficient structural optimization tool
Optistruct and a powerful explicit solver for highly nonlinear dynamic problems Radioss, several software packages by the ESI
group such as PAM-Composites for manufacturing and PAM-Crash for full-car crash simulations as well as AnalySwift SwiftComp
for general-purpose multiscale constitutive modeling with multiphysics capability.

8.4.5 Trends, Challenges and Future Directions

8.4.5.1 Round Robin Exercises


As mentioned in Section 8.4.3.4.3, one of the challenges associated with the analysis of composite materials and structures is the
accurate validation and verification of computational models and how these models can be used to predict unknown material
configurations or load cases. A number of coordinated studies including round-robin testing have been initiated to assess existing
computational composite models. A brief summary of these studies are presented in the following.

8.4.5.1.1 WWFE
The WWFE was introduced in the mid-1990s in order to close the gap between the academic world and the design world. It is
currently in its third edition.21 The first edition was focussed on failure criteria in composite structures under 2D state of stresses
whereas the second version evaluated existing failure criteria under triaxial stresses. Objectives of the current edition are damage
evolution under uniaxial, biaxial, bending and loading/unloading scenarios in combination with thermal effects and interacting
failure modes.
Key findings are published in Refs. [21,34] the first two editions concluded that only a few failure theories give
acceptable correlation (within 50%) with test data for 75% of the test cases used. In addition, it does not appear that there is any
criterion universally accepted by designers as being adequate under general load conditions.22 Nevertheless, WWFE is considered a
success where typical weaknesses were identified and most theories could be significantly improved. It also provides designers with
guidelines on accuracy and bounds of applicability for the current failure theories.

8.4.5.1.2 AQDTA
In 2014, the Air Force Research Laboratory (AFRL) initiated a study to assess and quantify the benefits of applying damage tolerant
design principles to advanced composite aircraft structures (AQDTA). The goal of AQDTA was to evaluate the state of the art in
progressive damage analysis (PDA). Blind and recalibration benchmarking exercises were performed on un-notched and notched
composite coupons under both static and fatigue loading.126 Most of the participants used a micromechanics-based multiscale
approach. CDM as well as advanced discrete failure modeling techniques were used. For the static results, stiffness and strength
blind predictions differed from experimental data by 19% on average while recalibrations reduced the difference to 8% on average.
In qualitative comparisons to X-ray measurements from experimental test coupons, discrete damage models (based on partition of
unity and CZM) performed best by accurate representation of matrix cracks and delamination. It was found that all presented
simulation tools require very high fidelity modeling and significant FEA expertise.126 Many of the participants needed additional
input data which was typically estimated from literature. Follow-on programs will address these issues to further evaluate existing
PDA methods.
76 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

8.4.5.1.3 Composite Materials Handbook-17


The Composite Materials Handbook-17 (CMH-17) is an organization consisting of industry, academia and regulators, and
supported by the Federal Aviation Administration (FAA) in the United States, whose primary purpose is to support the devel-
opment and use of composite materials and structures through publication and maintenance of proven and reliable guidelines/
standards for design and fabrication of composite materials and structures. As part of the CMH-17 crashworthiness and energy
management activities, a round-robin exercise was initiated in 2008 with the aim of assessing the performance of various
commercial FE codes, primarily explicit codes, in predicting the crush response of a number of composite components with
structural details that represent typical crash absorbing structures employed in aerospace industry. Commercial code vendors,
academia and industry participants willing to participate in this exercise have attempted to use the current capabilities of explicit
FE codes, LS-DYNA, Abaqus/Explicit, PAM-Crash, and Radioss, to tackle these challenging problems. This is an on-going simu-
lation effort backed by parallel testing program for their validation that is progressing from the base level of the building-block
consisting of small scale and simple coupons through to subassemblies of configured composite components that are repre-
sentative of crash absorbing structures.

8.4.5.2 Interaction of Failure Modes


As mentioned in Section 8.4.3.1, CDM originates from failure analyses in homogeneous metal structures. Since fatigue in metals is
mostly governed by a single dominant crack, most of the failure criteria and modeling concepts do not consider explicit failure
interaction which is required for realistic damage modeling in composite laminates. In addition, damage due to residual stresses
needs to be considered.22
Continuum models in Fig. 3 are not appropriate for the simulation of all failure modes in composites.11 In homogeneous
structures, cracks propagate normal to the maximum principal stress direction as shown in Fig. 5 (left). However, matrix cracks in
composites are influenced by the reinforcement architecture and morphology, i.e., they grow, for example, along the fiber direction
as illustrated in Fig. 5 (right). Hence it is necessary to enforce the orientation of matrix cracks in order to describe failure
mechanisms and their interaction realistically.11 X-FEM and PNM in combination with cohesive interface elements provide a
promising framework with FEM to model failure interaction on the micro- and mesoscale such as coalescence of multiple MCID as
shown in Fig. 7. Nonlocal CDM models with orthotropic averaging schemes have also been shown to be capable of tracking the
path of crack propagation in orthotropic composites.25 Recent studies20 have shown that a combination of the continuum and
discrete approaches can efficiently be used to capture the interacting effect of delamination (modeled with discrete cohesive zone
interface) and matrix cracks/splits (using the nonlocal continuum damage model, CODAM2). Through judicious placement of the
discrete delamination interface and synchronizing the onsets of delamination and splitting, the computational effort is markedly
reduced relative to a fully discrete approach that employs pre-inserted cohesive zones both within the plies and their interfaces.99
Since matrix cracks initiate other failure modes in composites, modeling of failure interaction is essential. Note that one key
objective of WWFE III is the prediction of matrix-driven delamination.21 Hence, explicit crack modeling through X-FEM or PNM
needs to be coupled to other failure mechanisms.
As earlier presented in Fig. 7,19 combine discrete matrix cracking with cohesive interface failure in composite laminates under
flexural loading. It is shown that MCID can be accurately predicted by PNM and breakable CZM cohesive elements.
De Carvalho et al.17 combine PNM with VCCT to model delamination migration in cross-ply laminates. Delamination
migration is another damage mechanism characterized by the interaction between delamination and matrix cracking. It is defined
as the process by which delamination propagation at a given interface relocates to a different interface via a matrix crack.
Numerical results show good qualitative and quantitative agreement with experiments.
Advanced three-dimensional failure modeling of open-hole composite laminates is presented in Ref. [11]. By applying PNM to
model matrix cracking, CZM in cohesive interface elements as well as CDM-based maximum strain criterion with exponential
degradation, failure patterns due to damage interaction can be accurately represented. Fig. 10 shows a comparison between

Fig. 10 Damage inspection of open-hole IM7/8552 composite laminate by (a) C-scan, (b) X-ray, and (c) simulation.11
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 77

optically determined damage and numerically predicted failure. Specifically, failure in the X-ray image shown in Fig. 10(b) is well
predicted by the computational model considering matrix, fiber and interface failure. Previously described crack band method in
CDM and a dissipation-based arc-length solution method are applied.

8.4.5.3 Uncertainty Quantification


One of the biggest challenges in modeling composites is the variability in material properties. This is caused by the complex
multiscale morphology of composites, introduced uncertainty in manufacturing processes and inconsistent testing methods. For
instance, small voids can initiate different failure modes such as matrix cracking or inter-laminar delamination and hence can lead
to a variation of material properties for identical composite laminates. Probabilistic testing methods 100 have been proposed to
account for material variability.34
Only the previously mentioned failure criteria by Cuntze 47 in Section 8.4.3.1.1 incorporates probabilistic effects which
achieved good results in the WWFE exercises. However, introducing uncertainties in computational composite models increases
the number of material input parameters whose determination requires extensive testing.
Fig. 11 summarizes possible uncertainties in composites. From a damage modeling standpoint, uncertainties in determination
of the fiber and matrix strength as well as structural uncertainties due to failure interaction are relevant.
In order to achieve safe and cost-effective design of composites, reliability analyses can evaluate the probability of failure.
Instead of the traditional method of applying high safety coefficients, statistical simulations can be used to design less conservative
structures.
However, accurate statistical methods such as Monte Carlo simulation may require a million FE runs.102 Reasonably accurate
and efficient statistical methods in conjunction with failure criteria are an ongoing area of research. For mathematical analysis, it is
necessary to describe the failure domain in an analytical form. This so-called limit state function (LSF) in composites reliability
analyses usually consists of the same failure criteria as those used in a deterministic approach such as WWFE methods in
Section 8.4.3.1. Hence, accuracy of a reliability analysis critically depends on the appropriate choice of failure criteria.103
Two methods to reduce computational cost associated with uncertainty quantification in composites are briefly discussed. A
comprehensive review on current reliability methods in composites can be found in Ref. [103]. Previous efforts on integrating
deterministic process models with reliability models to address the virtual processing of composites with probabilistic measures of
process outcomes can be found in Refs. [104–106].
Surrogate models are simplified representations of complex mechanical models which are used in reliability
analysis to approximate the LSF. Haeri and Fadaee102 apply an efficient surrogate model in combination with the Tsai-Wu failure
criterion as LSF to composites under various load cases such as uniaxial tension, bending or composite shells under uniform
pressure.
Multilevel Monte Carlo (MLMC) methods can be used where the construction of accurate surrogate models is impractical or
too complex. As opposed to traditional Monte Carlo analyses, this method only requires a handful of costly fine scale simulations
where the missing exploration of the variability is handled by a large number of coarse computations. Butler et al.107 successfully
apply MLMC to composite wing skin panels containing random manufacturing defects such as ply angle perturbations and
localized fiber waviness.

Fig. 11 Probabilistic design assessment of composite structures.101


78 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

8.4.5.4 Multiscale Modeling


Analyzing composite laminates on all length scales mentioned in Section 8.4.2.1 requires a very fine discretization and the
resulting FE mesh is highly heterogeneous. Even with ever increasing computational power through multicore or GPU-accelerated
analyses, alternative computational strategies have to be found in order to lower computational costs while improving numerical
robustness.30 Fig. 12 illustrates the complexity of multiscale modeling by considering the entire scaling range from the nanos-
tructure to the macroscopic response of fiber reinforced composites (FRC).
The ultimate modeling of composites modeling is the spatial as well as temporal scaling of problems for virtual testing. Fig. 13
illustrates this multiscale approach applied to a composite structure on the macroscale. The space domain is defined by con-
sidering laminae on the mesoscale and RVE on the microscale. Time scaling is necessary in fatigue analyses with a high number of
cycles.
Ladevèze30 formulates such an approach by splitting the time space into subintervals and by constructing substructures in the
spatial space. The discontinuous Galerkin method is used to handle possible discontinuities. A non-incremental iterative strategy
solves the “macro” problem at each iteration for the entire time interval along with independent linear problems in specific
substructures. The scheme is suitable for parallel multigrid computation.
A more recent example of multiscale modeling in composites is presented in Ref. [109]. By considering bending of hetero-
geneous plates, the macroscale is represented by the boundary element method whereas RVEs on the microscale are solved
by FEM.
Leading to the following section on process simulation,110 present a bottom-up multiscale approach. Fiber-matrix decohesion
in a RVE on the microscale is modeled by cohesive interface elements. Required in-situ nanomechanical measurement of matrix
and interface properties is further described. The structural response on the mesoscale is either represented by CDM or by pre-
inserted cohesive elements along the fiber direction to model matrix cracking. In addition, cohesive elements model inter-laminar
delamination. The multiscale formulation is applied to high velocity bird impact on composite structures. Challenges and future
developments associated with multiscale modeling are also outlined. It is concluded that a unified multiscale framework for
virtual composite processing is the next big challenge.

Fig. 12 High fidelity multiscale modeling of fiber reinforced composite (FRP) under flexural four-point bending.108

Fig. 13 Spatial and temporal multiscale modeling of macroscopic structural response in composites.108
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 79

Fig. 14 Multiscale simulation of multifunctional composites by considering virtual processing (left) and testing (right).110

8.4.5.5 Process Modeling


In order to achieve the long-term goal of an integrated design strategy for composites, accurate processing simulations are
required. This can enable design, testing and optimization of components before they are actually manufactured. The complete
computational description of processing is very challenging due to its multiphysics and multiscale nature. It has to take into
account mold filling, resin flow through fiber-bed, fabrics as well as within each fiber bundle, generation and transfer of heat
due to chemical reactions during curing and the effect of consolidation pressure.110,111 Fig. 14 illustrates how multiscale
methods in virtual processing and testing could be used and combined to model multifunctional composites in real
applications.
Currently, virtual processing is used to understand, refine and optimize manufacturing processes such as induction welding of
thermoplastic carbon fiber composites.112 Process simulation begins with material characterization by defining and measuring
mechanical and thermal behavior experienced under specific conditions during manufacturing. In most cases, temperature,
pressure, and time are the key parameters.
Process simulation or virtual manufacturing has gained popularity in recent years as an effective and robust approach to
minimize production risk and mitigate manufacturing-induced defects (e.g., Refs. [113–115]). Aside from certification concerns
regarding defects such as dimensional changes,116 porosity,117,118 under-cured and overcured parts,119–121 recently attention has
been paid to the effect of defects such as residual stresses on mechanical performance of composites including micro-crack
formation.111,122,123 It is shown, for example, by modifying process parameters such as cure cycle, the overall residual stress level
in a laminate may be reduced to increase the mechanical performance of 90 degree layers significantly.111
The current widely used approach is to employ an integrated model where the complexity of the model is broken into several
sub-models. For example, in a general approach, process simulation may be broken into three steps of thermochemical, flow-
compaction and stress-deformation simulations.123 For each of these simulations, based on the complexity of the utilized
constitutive model such as elastic models, pseudo-viscoelastic models, thermo-viscoelastic models or nonlinear viscoelastic
models,111 a relevant multiscale material characterization approach is required.

8.4.6 Summary and Conclusions

Several techniques can be employed for the structural analysis of composites through FE modeling. This chapter discussed a range
of FE modeling methods to guide potential users to efficient and accurate structural modeling of composites.
80 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

Based on specific structural dimensions and loading scenarios, the user chooses suitable element types as well as an appropriate
spatial representation and solution technique as introduced in Section 8.4.2. In addition, accurate structural modeling includes
computational damage or failure representation.
Section 8.4.3 presented two different FE techniques: smeared local stiffness reduction through CDM and discrete crack
simulation by CZM, VCCT, XFEM, or PNM. The associated critical calibration of input parameters was discussed alongside typical
problems and challenges. Several methods to overcome these difficulties were presented such as viscous regularization or
application of arc-length solution algorithms.
The most commonly used commercial FE software packages were then analyzed in regards to numerical modeling of
damage initiation and propagation. The different programs were: Simulia Abaqus, Ansys, LS-DYNA, MSC Software, AlphaStar
GENOA and Autodesk Helius Composite as well as Altair, ESI and AnalySwift. It was concluded that all packages contain standard
tools to model pre-inserted delamination such as CZM or VCCT. However, efficient crack modeling without a priori knowledge
on crack location and propagation are, to the best of the authors’ knowledge, only available in Abaqus, LS-DYNA, and AlphaStar
GENOA.
Furthermore, recent developments in academic research and future directions of structural modeling in composites were
studied in Section 8.4.5. Three round-robin validation challenges were presented to evaluate the progress on structural modeling
of composites: WWFE, AQDTA and CMH-17. It was followed by recent findings on modeling of failure interaction by combining
PNM with CZM or VCCT. With increasing computational power, uncertainty quantifications and multiscale methods in com-
posites become more realistic. General approaches and recent publications were presented which could lead to virtual processing
and testing of multifunctional composites in the future.

See also: 8.1 Designing Composite Materials Having Structural Integrity. 8.2 Conceptual Design of Composite Structures. 8.8 Analysis of
Delamination Damage in Composite Structures Using Cohesive Elements. 8.11 Foreign Object Impact on Composite Materials and the Modeling
Challenges. 8.12 Multiscale FE Modelling and Design of Composite Laminates Under Impact. 8.18 Design of Fiber Composite Aerospace
Structures Under Vibration Loads

References

1. Wisnom, M.R., Khan, B., Hallett, S.R., 2008. Size effects in unnotched tensile strength of unidirectional and quasi-isotropic carbon/epoxy composites. Composite
Structures 84 (1), 21–28. Available at: http://dx.doi.org/10.1016/j.compstruct.2007.06.002
2. Lavoie, J.A., Soutis, C., Morton, J., 2000. Apparent strength scaling in continuous fiber composite laminates. Composites Science and Technology 60 (2), 283–299.
Available at: http://dx.doi.org/10.1016/S0266-3538(99)00124-4
3. SIMULIA. 2017. Dassault Systemes. Available at: http://www.3ds.com/products-services/simulia/.
4. Forghani, A., Shahbazi, M., Zobeiry, N., Poursartip, A., Vaziri, R., 2015. An overview of continuum damage models to simulate intra-laminar failure mechanisms in
advanced composite materials. In: Camanho, P., Hallet, S. (Eds.), Numerical Modelling of Failure in Advanced Composite Materials. New York: Elsevier Ltd.
5. Xia, Z., Chen, Y., Ellyin, F., 2000. A meso/micro-mechanical model for damage progression in glass-fiber/epoxy cross-ply laminates by finite-element analysis.
Composites Science and Technology 60 (8), 1171–1179. Available at: http://dx.doi.org/10.1016/S0266-3538(00)00022-1
6. Zhang, Y., Xia, Z., Ellyin, F., 2005. Viscoelastic and damage analyses of fibrous polymer laminates by micro/meso-mechanical modeling. Journal of Composite Materials
39 (22), 2001–2022. doi:10.1177/0021998305052024.
7. Grufman, C., Ellyin, F., 2007. Determining a representative volume element capturing the morphology of fibre reinforced polymer composites. Composites Science and
Technology 67 (3–4), 766–775. Available at: http://dx.doi.org/10.1016/j.compscitech.2006.04.004
8. Melro, A.R., Camanho, P.P., Andrade Pires, F.M., Pinho, S.T., 2013. Micromechanical analysis of polymer composites reinforced by unidirectional fibres: Part I –
Constitutive modelling. International Journal of Solids and Structures 50 (11–12), 1897–1905. Available at: http://dx.doi.org/10.1016/j.ijsolstr.2013.02.009
9. Han, B., Veidt, M., Reiner, J., Dargusch, M., 2016. Application of augmented finite element and cohesive zone modelling to predict damage evolution in metal matrix
composites and aircraft coatings. Journal of Mechanical Engineering Research 8 (1), 1–17.
10. Bazant, Z.P., 2010. Can multiscale-multiphysics methods predict softening damage and structural failure? International Journal for Multiscale Computational Engineering
8 (1), 61–67. doi:10.1615/IntJMultCompEng.v8.i1.50.
11. Van Der Meer, F.P., Sluys, L.J., Hallett, S.R., Wisnom, M.R., 2012. Computational modeling of complex failure mechanisms in laminates. Journal of Composite Materials
46 (5), 603–623. doi:10.1177/0021998311410473.
12. Williams, K.V., Vaziri, R., 2001. Application of a damage mechanics model for predicting the impact response of composite materials. Computers & Structures 79 (10),
997–1011. Available at: http://dx.doi.org/10.1016/S0045-7949(00)00200-5
13. Perillo, G., Vedivik, N.P., Echtermeyer, A.T., 2015. Damage development in stitch bonded GFRP composite plates under low velocity impact: Experimental and numerical
results. Journal of Composite Materials 49 (5), 601–615. doi:10.1177/0021998314521474.
14. González, E.V., Maimí, P., Camanho, P.P., Turon, A., Mayugo, J.A., 2012. Simulation of drop-weight impact and compression after impact tests on composite laminates.
Composite Structures 94 (11), 3364–3378. doi:10.1016/j.compstruct.2012.05.015.
15. Reiner, J., Torres, J.P., Veidt, M., Heitzmann, M., 2016. Experimental and numerical analysis of drop-weight low-velocity impact tests on hybrid titanium composite
laminates. Journal of Composite Materials 50 (26), 3605–3617. doi:10.1177/0021998315624002.
16. Sebaey, T.A., Costa, J., Maimí, P., et al., 2014. Measurement of the in situ transverse tensile strength of composite plies by means of the real time monitoring of
microcracking. Composites Part B: Engineering 65, 40–46. Available at: http://dx.doi.org/10.1016/j.compositesb.2014.02.001
17. De Carvalho, N.V., Chen, B.Y., Pinho, S.T., et al., 2015. Modeling delamination migration in cross-ply tape laminates. Composites Part A: Applied Science and
Manufacturing 71, 192–203. doi:10.1016/j.compositesa.2015.01.021.
18. Chen, B.Y., Pinho, S.T., De Carvalho, N.V., Baiz, P.M., Tay, T.E., 2014. A floating node method for the modelling of discontinuities in composites. In: Engineering
Fracture Mechanics, 127. . pp. 104–134. National University of Singapore.
19. Reiner, J., Veidt, M., Dargusch, M., Gross, L., 2017. A progressive analysis of matrix cracking-induced delamination in composite laminates using an advanced phantom
node method. Journal of Composite Materials. doi:10.1177/0021998316684203.
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 81

20. Shahbazi, M., Vaziri, R., Zobeiry, N., 2016. An efficient virtual testing framework to simulate the interacting effect of intra-laminar and inter-laminar damage progression in
composite laminates. Paper presented at the American Society for Composites, 31st Technical Conference, Williamsburg, Virginia.
21. Kaddour, A.S., Hinton, M.J., Smith, P.A., Li, S., 2013. The background to the third world-wide failure exercise. Journal of Composite Materials 47 (20–21), 2417–2426.
doi:10.1177/0021998313499475.
22. Camanho, P.P., 2002. Failure Criteria for Fibre-Reinforced Polymer Composites. Porto: DEMEGI, FEUP.
23. Williams, K.V., Vaziri, R., Poursartip, A., 2003. A physically based continuum damage mechanics model for thin laminated composite structures. International Journal of
Solids and Structures 40 (9), 2267–2300. Available at: http://dx.doi.org/10.1016/S0020-7683(03)00016-7.
24. Forghani, A., 2011. A nonlocal approach to simulation of damage in composite structures. PhD Thesis, The University of British Colombia, Vancouver, Canada.
25. Forghani, A., Zobeiry, N., Poursartip, A., Vaziri, R., 2013. A structural modelling framework for prediction of damage development and failure of composite laminates.
Journal of Composite Materials 47 (20–21), 2553–2573. doi:10.1177/0021998312474044.
26. McGregor, C.J., Vaziri, R., Poursartip, A., Xiao, X., 2007. Simulation of progressive damage development in braided composite tubes under axial compression.
Composites Part A: Applied Science and Manufacturing 38 (11), 2247–2259. Available at: http://dx.doi.org/10.1016/j.compositesa.2006.10.007.
27. Zobeiry, N., Vaziri, R., Poursartip, A., 2015. Characterization of strain-softening behavior and failure mechanisms of composites under tension and compression.
Composites Part A: Applied Science and Manufacturing 68, 29–41. Available at: http://dx.doi.org/10.1016/j.compositesa.2014.09.009
28. Chamis, C.C., Murthy, P.L.N., Gotsis, P.K., Mital, S.K., 2000. Telescoping composite mechanics for composite behavior simulation. Computer Methods in Applied
Mechanics and Engineering 185 (2–4), 399–411. Available at: http://dx.doi.org/10.1016/S0045-7825(99)00268-6
29. Fish, J., Yu, Q., 2001. Multiscale damage modelling for composite materials: Theory and computational framework. International Journal for Numerical Methods in
Engineering 52 (1–2), 161–191. doi:10.1002/nme.276.
30. Ladevèze, P., 2004. Multiscale modelling and computational strategies for composites. International Journal for Numerical Methods in Engineering 60 (1), 233–253.
doi:10.1002/nme.960.
31. Raghavan, P., Li, S., Ghosh, S., 2004. Two scale response and damage modeling of composite materials. Finite Elements in Analysis and Design 40 (12), 1619–1640.
Available at: http://dx.doi.org/10.1016/j.finel.2003.11.003
32. Ghosh, S., Bai, J., Raghavan, P., 2007. Concurrent multi-level model for damage evolution in microstructurally debonding composites. Mechanics of Materials 39 (3),
241–266. Available at: http://dx.doi.org/10.1016/j.mechmat.2006.05.004
33. Kachanov, L.M., 1958. Time of the rupture process under creep conditions. Izvestiia Akademii Nauk SSSR, Otdelenie Teckhnicheskikh Nauk 8, 26–31.
34. Soden, P.D., Kaddour, A.S., Hinton, M.J., 2004. Recommendations for designers and researchers resulting from the world-wide failure exercise. Composites Science and
Technology 64 (3–4), 589–604. doi:10.1016/s0266-3538(03)00228-8.
35. Orifici, A.C., Herszberg, I., Thomson, R.S., 2008. Review of methodologies for composite material modelling incorporating failure. Composite Structures 86 (1–3),
194–210. doi:10.1016/j.compstruct.2008.03.007.
36. Liu, P.F., Zheng, J.Y., 2010. Recent developments on damage modeling and finite element analysis for composite laminates: A review. Materials and Design 31 (8),
3825–3834. doi:10.1016/j.matdes.2010.03.031.
37. Rohwer, K., 2016. Models for intra-laminar damage and failure of fiber composites – A review. Facta Universitatis, Series: Mechanical Engineering 14 (1),
1–19.
38. Dávila, C.G., Camanho, P.P., Rose, C.A., 2005. Failure criteria for FRP laminates. Journal of Composite Materials 39 (4), 323–345.
39. Hill, R., 1948. A theory of the yielding and plastic flow of anisotropic metals. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering
Sciences 193 (1033), 281–297. doi:10.1098/rspa.1948.0045.
40. Tsai, S.W., 1965. Strength characteristics of composite materials, NASA Contractor Report CR-224. Washington, DC: National Aeronautics and Space Administration.
41. Tsai, S.W., Wu, E.M., 1971. General theory of strength for anisotropic materials. Journal of Composite Materials 5, 58–80.
42. Hoffman, O., 1967. The brittle strength of orthotropic materials. Journal of Composite Materials 1 (2), 200–206. doi:10.1177/002199836700100210.
43. Yamada, S.E., Sun, C.T., 1978. Analysis of laminate strength and its distribution. Journal of Composite Materials 12 (3), 275–284. doi:10.1177/002199837801200305.
44. Hashin, Z., 1980. Failure criteria for unidirectional fiber composites. Journal of Applied Mechanics, Transactions ASME 47 (2), 329–334.
45. Chang, F.K., Chang, K.Y., 1987. A progressive damage model for laminated composites containing stress concentrations. Journal of Composite Materials 21 (9),
834–855. doi:10.1177/002199838702100904.
46. Puck, A., Schürmann, H., 1998. Failure analysis of FRP laminates by means of physically based phenomenological models. Composites Science and Technology 58 (7),
1045–1067. doi:10.1016/s0266-3538(96)00140-6.
47. Cuntze, R.G., Freund, A., 2004. The predictive capability of failure mode concept-based strength criteria for multidirectional laminates. Composites Science and
Technology 64 (3-4), 343–377. doi:10.1016/s0266-3538(03)00218-5.
48. Pinho, S.T., Davila, C.G., Camanho, P.P., Iannucci, L., Robinson, P., 2005. Failure models and criteria for FRP under in-plane or three-dimensional stress states
including shear non-linearity. Available at: http://hdl.handle.net/2002/15158
49. Dávila, C.G., Jaunky, N., Goswami, S., 2003. Failure criteria for FRP laminates in plane stress. Paper presented at the 44th AIAA/ASME/ASCE/AHS/ASC Structures,
Structural Dynamics, and Materials Conference.
50. Gosse, J.H., Christensen, S., 2001. Strain invariant failure criteria for polymers in composite materials. Paper presented at the Collection of Technical Papers - AIAA/
ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference.
51. Garnich, M.R., Akula, V.M.K., 2009. Review of degradation models for progressive failure analysis of fiber reinforced polymer composites. Applied Mechanics Reviews
62 (1), 1–33. doi:10.1115/1.3013822.
52. Ladeveze, P., LeDantec, E., 1992. Damage modelling of the elementary ply for laminated composites. Composites Science and Technology 43 (3), 257–267. doi:10.1016/
0266-3538(92)90097-m.
53. Matzenmiller, A., Lubliner, J., Taylor, R.L., 1995. A constitutive model for anisotropic damage in fiber-composites. Mechanics of Materials 20 (2), 125–152. Available at:
http://dx.doi.org/10.1016/0167-6636(94)00053-0
54. Eduardo, M., Griffin, Jr., O., 1997. Progressive failure analysis of laminated composite structures. In: 38th Structures, Structural Dynamics, and Materials Conference.
American Institute of Aeronautics and Astronautics.
55. Van Der Meer, F.P., Sluys, L.J., 2009. Continuum models for the analysis of progressive failure in composite laminates. Journal of Composite Materials 43 (20),
2131–2156. doi:10.1177/0021998309343054.
56. Rybicki, E.F., Kanninen, M.F., 1977. A finite element calculation of stress intensity factors by a modified crack closure integral. Engineering Fracture Mechanics 9 (4),
931–938. doi:10.1016/0013-7944(77)90013-3.
57. Dugdale, D.S., 1960. Yielding of steel sheets containing slits. Journal of the Mechanics and Physics of Solids 8 (2), 100–104.
58. Barenblatt, G.I., 1962. The mathematical theory of equilibrium cracks in brittle fracture. Advances in Applied Mechanics 7, 55–129.
59. Shet, C., Chandra, N., 2002. Analysis of energy balance when using cohesive zone models to simulate fracture processes. Journal of Engineering Materials and
Technology, Transactions of the ASME 124 (4), 440–450. doi:10.1115/1.1494093.
60. Turon, A., Camanho, P.P., Costa, J., Dávila, C.G., 2006. A damage model for the simulation of delamination in advanced composites under variable-mode loading.
Mechanics of Materials 38 (11), 1072–1089. doi:10.1016/j.mechmat.2005.10.003.
61. Cox, B., Yang, Q., 2006. In quest of virtual tests for structural composites. Science 314 (5802), 1102–1107. doi:10.1126/science.1131624.
82 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

62. Hallett, S.R., Wisnom, M.R., 2006. Numerical investigation of progressive damage and the effect of layup in notched tensile tests. Journal of Composite Materials 40 (14),
1229–1245. doi:10.1177/0021998305057432.
63. Camanho, P.P., Dávila, C.G., De Moura, M.F., 2003. Numerical simulation of mixed-mode progressive delamination in composite materials. Journal of Composite
Materials 37 (16), 1415–1438.
64. Bažant, Z.P., Oh, B.H., 1983. Crack band theory for fracture of concrete. Matériaux et Constructions 16 (3), 155–177. doi:10.1007/bf02486267.
65. Shor, O., Vaziri, R., 2015. Adaptive insertion of cohesive elements for simulation of delamination in laminated composite materials. Engineering Fracture Mechanics 146,
121–138. Available at: http://dx.doi.org/10.1016/j.engfracmech.2015.07.044
66. Shor, O., Vaziri, R., 2017. Application of the local cohesive zone method to numerical simulation of composite structures under impact loading. International Journal of
Impact Engineering 104, 127–149.
67. Melenk, J.M., Babuška, I., 1996. The partition of unity finite element method: Basic theory and applications. Computer Methods in Applied Mechanics and Engineering
139 (1–4), 289–314.
68. Moës, N., Dolbow, J., Belytschko, T., 1999. A finite element method for crack growth without remeshing. International Journal for Numerical Methods in Engineering
46 (1), 131–150.
69. Iarve, E.V., 2003. Mesh independent modelling of cracks by using higher order shape functions. International Journal for Numerical Methods in Engineering 56 (6),
869–882. doi:10.1002/nme.596.
70. Mollenhauer, D., Iarve, E.V., Kim, R., Langley, B., 2006. Examination of ply cracking in composite laminates with open holes: A moiré interferometric and numerical
study. Composites Part A: Applied Science and Manufacturing 37 (2), 282–294. doi:10.1016/j.compositesa.2005.06.004.
71. Hansbo, A., Hansbo, P., 2004. A finite element method for the simulation of strong and weak discontinuities in solid mechanics. Computer Methods in Applied
Mechanics and Engineering 193 (33-35), 3523–3540.
72. van der Meer, F.P., Sluys, L.J., 2009. A phantom node formulation with mixed mode cohesive law for splitting in laminates. International Journal of Fracture 158 (2),
107–124. doi:10.1007/s10704-009-9344-5.
73. Talreja, R., 2008. Damage and fatigue in composites – A personal account. Composites Science and Technology 68 (13), 2585–2591. doi:10.1016/j.
compscitech.2008.04.042.
74. Girão Coelho, A.M., 2016. Finite element guidelines for simulation of delamination dominated failures in composite materials validated by case studies. Archives of
Computational Methods in Engineering 23 (2), 363–388. doi:10.1007/s11831-015-9144-1.
75. de Moura, M.F.S.F., Campilho, R.D.S.G., Amaro, A.M., Reis, P.N.B., 2010. Inter-laminar and intra-laminar fracture characterization of composites under mode I loading.
Composite Structures 92 (1), 144–149. Available at: http://dx.doi.org/10.1016/j.compstruct.2009.07.012
76. Sager, R.J., Klein, P.J., Davis, D.C., et al., 2011. Inter-laminar fracture toughness of woven fabric composite laminates with carbon nanotube/epoxy interleaf films. Journal
of Applied Polymer Science 121 (4), 2394–2405. doi:10.1002/app.33479.
77. Reiner, J., Torres, J.P., Veidt, M., 2017. A novel top surface analysis method for mode i interface characterisation using digital image correlation. Engineering Fracture
Mechanics. 0013-7944), Available at:http://dx.doi.org/10.1016/j.engfracmech.2016.12.022
78. Benzeggagh, M.L., Kenane, M., 1996. Measurement of mixed-mode delamination fracture toughness of unidirectional glass/epoxy composites with mixed-mode bending
apparatus. Composites Science and Technology 56 (4), 439–449. doi:10.1016/0266-3538(96)00005-x.
79. Turon, A., Davila, C.G., Camanho, P.P., Costa, J., 2007. An engineering solution for solving mesh size effects in the simulation of delamination with cohesive zone
models. Engineering Fracture Mechanics 74 (10), 1665–1682.
80. Putar, F., Soric, J., Lesicar, T., Tonkovic, Z., 2016. Damage modelling using strain gradient based finite element formulation. Paper presented at the ECCOMAS Congress,
Crete Island, Greece.
81. Li, F., Xingwen, D., 2010. Mesh-dependence of material with softening behavior. Chinese Journal of Aeronautics 23 (1), 46–53. Available at: http://dx.doi.org/10.1016/
S1000-9361(09)60186-2
82. Bazant, Z.P., Jirásek, M., 2002. Nonlocal integral formulations of plasticity and damage: Survey of progress. Journal of Engineering Mechanics 128 (11), doi:10.1061/
(ASCE)0733-9399(2002)128:11(1119).
83. Bazǎ nt, Z. k. P., Pijaudier-Cabot, G., 1988. Nonlocal continuum damage, localization instability and convergence. Journal of Applied Mechanics 55 (2), 287–293.
doi:10.1115/1.3173674.
84. Jirásek, M., Marfia, S., 2005. Non-local damage model based on displacement averaging. International Journal for Numerical Methods in Engineering 63 (1), 77–102.
doi:10.1002/nme.1262.
85. Korsunsky, A.M., Nguyen, G.D., Houlsby, G.T., 2005. Analysis of essential work of rupture using non-local damage-plasticity modelling. International Journal of Fracture
135 (1), L19–L26. doi:10.1007/s10704-005-4391-z.
86. Bongers, C., 2011. A stress-based gradient-enhanced damage model. Master, Delft University of Technology.
87. Aifantis, E.C., 1992. On the role of gradients in the localization of deformation and fracture. International Journal of Engineering Science 30 (10), 1279–1299. Available at:
http://dx.doi.org/10.1016/0020-7225(92)90141-3.
88. de Borst, R., Pamin, J., Geers, M.G.D., 1999. On coupled gradient-dependent plasticity and damage theories with a view to localization analysis. European Journal of
Mechanics - A/Solids 18 (6), 939–962. Available at: http://dx.doi.org/10.1016/S0997-7538(99)00114-X.
89. Peerlings, R.H.J., De Borst, R., Brekelmans, W.A.M., De Vree, J.H.P., 1996. Gradient enhanced damage for quasi-brittle materials. International Journal for Numerical
Methods in Engineering 39 (19), 3391–3403. doi:10.1002/(SICI)1097-0207(19961015)39:19o3391::AID-NME743.0.CO;2-D.
90. Duvaut, G., Lions, J.L., 1976. Inequalities in Mechanics and Physics. Berlin: Springer-Verlag.
91. Riks, E., 1979. An incremental approach to the solution of snapping and buckling problems. International Journal of Solids and Structures 15 (7), 529–551. doi:10.1016/
0020-7683(79)90081-7.
92. Ramm, E., 1981. Strategies for Tracing the Nonlinear Response Near Limit Points. Berlin, Heidelberg: Springer.
93. Crisfield, M.A., 1983. Arc-length method including line searches and accelerations. International Journal for Numerical Methods in Engineering 19 (9), 1269–1289.
94. Babuska, I., Oden, J.T., 2004. Verification and validation in computational engineering and science: Basic concepts. Computer Methods in Applied Mechanics and
Engineering 193 (36–38), 4057–4066. Available at: http://dx.doi.org/10.1016/j.cma.2004.03.002.
95. Oberkampf, W.L., Barone, M.F., 2006. Measures of agreement between computation and experiment: Validation metrics. Journal of Computational Physics 217 (1), 5–36.
Available at: http://dx.doi.org/10.1016/j.jcp.2006.03.037.
96. Roache, P.J., 1998. Verification and Validation in Computational Science and Engineering. Albuquerque, NM: Hermosa..
97. Miot, S., 2017. Review of composites simulation tools. Available at: https://www.nafems.org/about/technical-working-groups/composites/
a_review_of_composites_simulation_tools/
98. AlphaStar. 2017. GENOA. Available at: http://www.alphastarcorp.com/sections/Products/GENOA/index.jsp
99. Hallett, S.R., Green, B.G., Jiang, W.G., Wisnom, M.R., 2009. An experimental and numerical investigation into the damage mechanisms in notched composites.
Composites Part A: Applied Science and Manufacturing 40 (5), 613–624. Available at: http://dx.doi.org/10.1016/j.compositesa.2009.02.021
100. Lin, K.Y., Du, L.J., Rusk, D., 2000. Structural design methodology based on concepts of uncertainty. NASA_techdocs.
101. Chamis, C.C., 2004. Probabilistic simulation of multi-scale composite behavior. Theoretical and Applied Fracture Mechanics 41 (1–3), 51–61. doi:10.1016/j.
tafmec.2003.11.005.
Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods 83

102. Haeri, A., Fadaee, M.J., 2016. Efficient reliability analysis of laminated composites using advanced Kriging surrogate model. Composite Structures 149, 26–32.
doi:10.1016/j.compstruct.2016.04.013.
103. Chiachio, M., Chiachio, J., Rus, G., 2012. Reliability in composites – A selective review and survey of current development. Composites Part B: Engineering 43 (3),
902–913. doi:10.1016/j.compositesb.2011.10.007.
104. Li, H., Foschi, R., Vaziri, R., Fernlund, G., Poursartip, A., 2002. Probability-based modelling of composites manufacturing and its application to optimal process design.
Journal of Composite Materials 36 (16), 1967–1991. doi:10.1177/0021998302036016241.
105. Bebamzadeh, A., Haukaas, T., Vaziri, R., Poursartip, A., Fernlund, G., 2009. Response sensitivity and parameter importance in composites manufacturing. Journal of
Composite Materials 43 (6), 621–659. doi:10.1177/0021998308101299.
106. Bebamzadeh, A., Haukaas, T., Vaziri, R., Poursartip, A., Fernlund, G., 2010. Application of response sensitivity in composite processing. Journal of Composite Materials
44 (15), 1821–1840. doi:10.1177/0021998310366062.
107. Butler, R., Dodwell, T.J., Haftka, R.T, et al., 2015. Uncertainty quantification of composite structures with defects using multilevel Monte Carlo simulations. In: 17th AIAA
Non-Deterministic Approaches Conference. American Institute of Aeronautics and Astronautics.
108. Vanderbilt University. 2016. Multiscale modeling of random nano- and micro-fiber reinforced cementitious composites. Available at: https://my.vanderbilt.edu/mcml/cv/
multiscale-modeling-of-random-nano-and-micro-fiber-reinforced-cementitious-composites-2/
109. Fernandes, G.R., Pituba, J.J.C., De Souza Neto, E.A., 2015. Multi-scale modelling for bending analysis of heterogeneous plates by coupling BEM and FEM. Engineering
Analysis with Boundary Elements 51, 1–13. doi:10.1016/j.enganabound.2014.10.005.
110. Llorca, J., González, C., Molina-Aldareguía, J.M., et al., 2011. Multiscale modeling of composite materials: A roadmap towards virtual testing. Advanced Materials 23 (44),
5130–5147. doi:10.1002/adma.201101683.
111. Zobeiry, N., Forghani, A., Li, C., et al., 2016. Multi-scale characterization and representation of composite materials during processing. Philosophical Transactions of the
Royal Society of London A: Mathematical, Physical and Engineering Sciences A374 (20150278),
112. Duhovic, M., L’Eplattenier, P., Caldichoury, I., Mitschang, P., Maier, M., 2014. Advanced 3D finite element simulation of thermoplastic carbon fiber composite induction
welding. Paper presented at the 16th European Conference on Composite Materials, ECCM 2014.
113. Johnston, A., Vaziri, R., Poursartip, A., 2001. A plane strain model for process-induced deformation of laminated composite structures. Journal of Composite Materials
35 (16), 1435–1469. doi:10.1106/YXEA-5MH9-76J5-BACK.
114. Fernlund, G., Rahman, N., Courdji, R., et al., 2002. Experimental and numerical study of the effect of cure cycle, tool surface, geometry, and lay-up on the dimensional
fidelity of autoclave-processed composite parts. Composites Part A: Applied Science and Manufacturing 33 (3), 341–351. Available at: http://dx.doi.org/10.1016/S1359-
835X(01)00123-3.
115. Fernlund, G., Osooly, A., Poursartip, A., et al., 2003. Finite element based prediction of process-induced deformation of autoclaved composite structures using 2D process
analysis and 3D structural analysis. Composite Structures 62 (2), 223–234. Available at: http://dx.doi.org/10.1016/S0263-8223(03)00117-X.
116. Albert, C., Fernlund, G., 2002. Spring-in and warpage of angled composite laminates. Composites Science and Technology 62 (14), 1895–1912. Available at: http://dx.
doi.org/10.1016/S0266-3538(02)00105-7
117. Kay, J., Farhang, L., Hsiao, K., Fernlund, G., 2011. Effect of process conditions on porosity in out-of-autoclave prepreg laminates. Effect of process conditions on porosity
in out-of-autoclave prepreg laminates. Paper presented at the 18th International Conference on Composite Materials, Jeju Island, Korea.
118. Fernlund, G., Wells, J., Fahrang, L., Kay, J., Poursartip, A., 2016. Causes and remedies for porosity in composite manufacturing. Paper presented at the In IOP
Conference Series: Materials Science and Engineering.
119. Fabris, J., Lussier, D., Zobeiry, N., Mobuchon, C., Poursartip, A., 2014. Development of standardized approaches to thermal management in composites manufacturing.
Paper presented at the SAMPE Conference, Seattle, WA.
120. Fabris, J., Mobuchon, C., Zobeiry, N., et al., 2015. Introducing thermal history producibility assessment at conceptual design. Paper presented at the SAMPE Conference,
Baltimore, MD.
121. Fabris, J., Mobuchon, C., Zobeiry, N., Poursartip, A., 2016. Understanding the consequences of tooling design choices on thermal history in composites processing.
Paper Presented at the Sampe Conference, Long Beach, CA.
122. Li, C., Zobeiry, N., Chatterjee, S., Poursartip, A., 2014. Advances in the characterization of residual stress in composite structures. Paper presented at the SAMPE
Conference, Seattle, WA.
123. Zobeiry, N., Poursartip, A., 2015. 3 – The origins of residual stress and its evaluation in composite materials. In: Beaumont, P.W.R., Soutis, C., Hodzic, A. (Eds.),
Structural Integrity and Durability of Advanced Composites. Cambridge: Woodhead Publishing, pp. 43–72.
124. McGregor, C., Vaziri, R., Xiao, X., 2010. Finite element modelling of the progressive crushing of braided composite tubes under axial impact. International Journal of
Impact Engineering 37 (6), 662–672.
125. McGregor, C., Zobeiry, N., Vaziri, R., Poursartip, A., Xiao, X., 2017. Calibration and validation of a continuum damage mechanics model in aid of axial crush simulation
of braided composite tubes. Composites Part A: Applied Science and Manufacturing 95, 208–219.
126. Engelstad, S.P., Stover, R.J., Action, J.E., et al., 2015. Air Vehicle Integration and Technology Research (AVIATR), Task Order 0037: Assessment, Quantification, and
Benefits of Applying Damage Tolerant Design Principles to Advanced Composite Aircraft Structure, AFRL-RQ-WP-TR-2015-0068, February 2015.

Relevant Websites

http://www.ansys.com/
ANSYS.
http://www.alphastarcorp.com/sections/Products/GENOA/index.jsp
Alphastar.
http://www.altairhyperworks.com
Altair.
http://analyswift.com/
Analyswift.
https://www.cmh17.org/HOME/PolymerMatrix/Crashworthiness.aspx
CMH-17.
http://www.3ds.com/products-services/simulia/
Dassault Systemes.
https://www.esi-group.com/
ESI.
http://www.autodesk.com/products/helius-composite/overview
Helius Composite.
84 Structural Analysis of Composites With Finite Element Codes: An Overview of Commonly Used Computational Methods

http://www.autodesk.com/products/helius-pfa/overview
Helius PFA.
http://www.lstc.com/
Livermore Software Technology Corporation.
http://www.mscsoftware.com/
MSC.

You might also like