You are on page 1of 100

Submitted by

Andreas Peter Pröll

Submitted at
Institute of Structural
Lightweight Design

Supervisor
Univ.-Prof. Dipl.-Ing. Dr.
Martin Schagerl

Co-Supervisor
Asst.-Prof. Adi

Fatigue Damage Models Adumitroaie, Ph.D.


Dipl.-Ing. Lukas
Retschitzegger

for Laminated July 2017

Composite Structures

Master’s Thesis
to obtain the academic degree of

Diplom-Ingenieur
in the Master’s Program

Polymer Technologies and Science

JOHANNES KEPLER
UNIVERSITY LINZ
Altenbergerstraße 69
4040 Linz, Österreich
www.jku.at
DVR 0093696
Statutory Declaration i

Statutory Declaration
I hereby declare that the thesis submitted is my own unaided work, that I have not used other than
the sources indicated, and that all direct and indirect sources are acknowledged as references. This
printed thesis is identical with the electronic version submitted.

place and date Signature


Abstract ii

Abstract
Many cyclic loaded structures show damage after a certain number of cycles even though the max-
imum stress in a cycle is far below static strength. This phenomenon is called fatigue. It is a critical
criterion and has to be considered for appropriate dimensioning of engineering structures, which
are in many cases subjected to repeated loadings. Especially in the field of laminated composite
materials, fatigue is still content of extensive research due to their complex damage mechanisms.
The present work focuses onto the investigation of the current state of finite elemente analysis
(FEA) software packages in the field of fatigue of laminated composites. Due to the motivation of
a possible application to a composite rim, where problems with fatigue delaminations occur, the
focus of the assessment lies on interlaminar fatigue damage.

To achieve these objectives, a certain theoretical basis is needed. Therefore, the first part of the
thesis contains a summary of fundamentals in fracture mechanics, laminated composites, fatigue
modeling in general and state of the art fatigue methods for laminated composites. In the second
part, extensive reviews of theories of selected FEA software manufacturers are given, namely Siemens
Samtech Samcef and 3DS Abaqus. Former manufacturer, which meanwhile integrated the software
package into their product lifecycle management (PLM) environment NX, the theory, which is based
on continuum damage mechanics (CDM), focuses on intralaminar fatigue damage. For interlam-
inar damage, a cohesive zone model is suggested, but no specific fatigue theory is developed until
now. Since the fatigue model was still not implemented into the software, no assessment could be
done. Abaqus implemented a low cycle fatigue tool for interlaminar crack growth, which is based
on linear elastic fracture mechanics (LEFM) and Paris law for fatigue crack growth. Furthermore,
the onset of a crack is considered in an additional criterion. However, after an extensive practical
assessment, it was concluded that the method is still very limited in its capabilities and shows some
unreasonable behavior. Accordingly, its applicability onto complex structural components such as
a composite rim is not recommended and hence was not done.

In conclusion, the tools for calculating the fatigue behavior of composite materials mentioned in
this work are not yet fully applicable for evaluating practical problems. Based on the knowledge
obtained in the reviews and the assessments, a proposal for the treatment of fatigue in laminated
composite materials is given for future work.
Table of Contents iii

Statutory Declaration i

Abbreviations v

1 Introduction 1
1.1 Overview and state of need . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Research goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2 Fundamentals in fracture mechanics 2


2.1 History of fracture mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2 Crack definition, damage modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.3 Linear elastic fracture mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.3.1 Stress based approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.3.2 Energy based approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.4 Mode mixture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

3 Fundamentals of laminated composite materials 6


3.1 Material-composition and macroscopic structure . . . . . . . . . . . . . . . . . . . . 6
3.1.1 Fiber-materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1.2 Fiber configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.1.3 Matrix-materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2 Damage mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.2.1 Intralaminar damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.2.2 Interlaminar damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.3 General 3D modeling approach in Finite Element Analysis . . . . . . . . . . . . . . . 13
3.3.1 Element types for laminate modeling . . . . . . . . . . . . . . . . . . . . . . . 13
3.3.2 Techniques for interface damage modeling . . . . . . . . . . . . . . . . . . . . 15

4 Fundamentals in fatigue 19
4.1 Fatigue: History and general definition . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.2 Design philosophies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4.3 Loading conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.4 Fatigue damage evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.4.1 Damage initiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.4.2 Onset of propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.4.3 Damage propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.5 Treatment of general load spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.5.1 Classification of general load spectra . . . . . . . . . . . . . . . . . . . . . . . 24
4.5.2 Damage accumulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.6 Comparison of fatigue behavior of metals and laminated composites . . . . . . . . . 25

5 State-of-the-art fatigue damage modeling techniques of laminated composites 26


5.1 Laminate and lamina fatigue life estimation . . . . . . . . . . . . . . . . . . . . . . . 26
5.2 Progressive damage models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3 Interlaminar fatigue damage models . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3.1 LEFM methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
5.3.2 Cohesive zone methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
Table of Contents iv

6 Theories behind selected fatigue models in FEA-packages 29


6.1 Siemens Samtech Samcef: intralaminar fatigue damage modeling of woven and UD
FRP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
6.1.1 General modeling approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
6.1.2 Identification of the material parameters needed . . . . . . . . . . . . . . . . 31
6.1.3 Current status, known advantages and drawbacks of the model . . . . . . . . 32
6.2 3DS Simulia Abaqus: interlaminar fatigue damage modeling using VCCT low cycle
fatigue analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
6.2.1 General modeling approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
6.2.2 Identification of the material parameters needed . . . . . . . . . . . . . . . . 35

7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D
elements 37
7.1 Development of a suitable reference case . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.1.1 End notch flexure (ENF) FE model . . . . . . . . . . . . . . . . . . . . . . . 38
7.1.2 Description of the relevant Abaqus keywords . . . . . . . . . . . . . . . . . . 42
7.1.3 Example input file . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
7.1.4 Parameters and options used for increasing accuracy and convergence . . . . 45
7.1.5 Simulations and determination of the reference case . . . . . . . . . . . . . . 46
7.2 Assessment of fatigue delamination under cyclic loading . . . . . . . . . . . . . . . . 54
7.2.1 Treatment of the stress ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7.2.2 Treatment of mixed-mode loading conditions . . . . . . . . . . . . . . . . . . 57
7.2.3 Damage accumulation at the crack front in 3D simulations . . . . . . . . . . 58
7.2.4 Impact of the crack onset criterion . . . . . . . . . . . . . . . . . . . . . . . . 60
7.2.5 Behavior at phase-shifted and non-sinusoidal cyclic loadings . . . . . . . . . . 64
7.2.6 CPU-parallelization, cycle limit . . . . . . . . . . . . . . . . . . . . . . . . . . 69
7.3 Guidelines for implementation and input parameters . . . . . . . . . . . . . . . . . . 71
7.4 Summary of the assessment regarding fatigue delamination under cyclic loading . . . 72

8 Conclusions and future work 73


8.1 Research goals and performed work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
8.2 Conclusions and recommendations for future work . . . . . . . . . . . . . . . . . . . 73

Literature 75

Appendix 78
Abbreviations v

Abbreviations

Abbr. Abbreviation

3D three dimensional

CFRP Carbon Fibre Reinforced Polymer/Plastic

CLD Constant Life Diagram

CZM Cohesive Zone Modelling

DCB Double Cantilever Beam

ENF End Notch Flexure

EPFM Elastic Plastic Fracture Mechanics

FE Finite Element

FEA Finite Element Analysis

FF Fiber-Failure

FRP Fibre Reinforced Polymer/Plastic

FSDT First order Shear Deformation Theory

GFRP Glass Fibre Reinforced Polymer/Plastic

HCF High Cycle Fatigue

iFF inter Fiber Failure

IT Information Technology

LCF Low Cycle Fatigue

LEFM Linear Elastic Fracture Mechanics

MMR Mixed Mode Ratio

PAN Polyacrylnitrile

PLM Product Lifecycle Management

PSS Ply Stacking Sequence

SHM Structural Health Monitoring

UD Unidirectional

VCCT Virtual Crack Closure Technique


1 Introduction 1

1 Introduction
1.1 Overview and state of need
In many engineering applications, lightweight design is of special interest to increase the perform-
ance, efficiency, the environmental sustainability or even to enable new technologies. This can be
obtained by the optimization of the material selection, construction and the global system itself [1].
In the last decades, laminated composite materials exhibit rising production volumes due to their
immense potential in lightweight design. This class of materials shows very good fatigue properties
and damage tolerance, in particular compared to aluminum [2, 3]. Furthermore, other than isotropic
materials, the coupling between the individual stresses and strains can be tailored by optimizing
the ply stacking sequence (PSS), which is unique at the material level and opens up completely
new design possibilities. This property is for example used for implementing swept forward wings
at a jet fighter for increased maneuverability [4, 5], which would be very difficult with isotropic ma-
terials. However, laminated composites have quite complex mechanical behavior, which is a result
of their heterogeneous structure and the used polymer matrices. Compared to metals, until now
there is no general calculation technique to predict the damage behavior of laminated composites
including all damage mechanisms, especially under fatigue loadings. Therefore, the full lightweight
potential of laminated composites still cannot be obtained in an efficient, fast way for structural
components in most cases up to now. This leads to denial of the material in many applications
for the reason of an iterative, very time consuming design processes and the need of an excessive
amount of component testing besides high material and manufacturing costs. To overcome these
problems, advanced theories and calculation methods are needed.

1.2 Research goals


In the last couple of years, new methods for calculating fatigue of composite structures were proposed
by researchers and some of them were implemented in finite element analysis (FEA) packages
recently. Thus, the aim of the following work is to investigate selected finite element analysis (FEA)
packages regarding their capabilities to simulate fatigue of laminated composite structures. For
the reason of a possible application on a composite rim, where problems with fatigue delamination
occur, the main focus of the following assessments lies on methods for modeling interlaminar damage
under fatigue loading.

1.3 Thesis outline


Before selected methods can be tested, extended knowledge about laminated composites, funda-
mentals in fatigue and fracture mechanics theories is required, which is acquired and summarized
in a first step (sections 2-4). In the next chapter, current available techniques to simulate damage
and fatigue in composite materials are reviewed (section 5). After that, theories of selected fatigue
theories of FEA software manufacturers are analyzed and discussed (section 6). In the last chapter
these models are tested in the corresponding software package, if already implemented in the FEA
software (section 7).
2 Fundamentals in fracture mechanics 2

2 Fundamentals in fracture mechanics


This section provides a brief description of common fracture mechanics approaches and their history.

2.1 History of fracture mechanics


Fracture mechanics describes the influence and growth of existing cracks and flaws in a structure
in an analytical way using solid mechanics principles. It goes back to the beginning of the 20th
century when Griffith found an energy based approach for describing the strength of ideal-brittle
materials such as glass containing artificial flaws. This theory bases on a critical release rate of
elastic energy when introducing a growing crack to form new surfaces [6]. However as a result
of fully neglecting ductile behavior this theory is not suitable for metals due to a plastic zone
around the crack tip. Therefore in 1948 Orowan extended the theory by considering the effect of
additional energy dissipation as a result of plastic deformation [7]. The next major developments
were performed by Irwin around 1960: he introduced a method of calculating the asymptotic stress
field around a crack tip by formulating a stress intensity factor K. Therefore this approach is called
the K-concept. In addition he was able to connect his stress-based approach to the energy-based
approach of Griffith and Orowan.

2.2 Crack definition, damage modes


Cracks can be considered as cuts into a body. Figure 2.1 depicts a crack with its crack surfaces.
These are typically in tensionless state and touch each other at the crack front (red colored in
figure 2.1) [8].

crack front
crack surfaces

Figure 2.1: cracked body

Cracks can open in three different modes, depicted in figure 2.2. Thereby Mode 1 shows symmetric
crack opening with respect to the crack mid surface caused by normal forces in opening direction.
Mode 2 describes crack growth caused by anti symmetric sliding of the crack surfaces. This is
induced by in-plane shear forces which are perpendicular to the crack front. The last mode - Mode
3 - depicts crack growth due to out-of-plane shear forces. These generate relative displacements
tangential to the crack front - similar to an opening scissor.

Mode 1: Mode 2: Mode 3:


opening sliding tearing

Figure 2.2: the three fracture modes of cracked surfaces


2 Fundamentals in fracture mechanics 3

2.3 Linear elastic fracture mechanics


Linear elastic fracture mechanics (LEFM) act on the assumption of small scale yielding which has
to be evaluated with special care for the corresponding material, structure and loading case. For
metals and brittle materials this assumption is fulfilled in most cases. More information can be
found in literature and [9], [10].

2.3.1 Stress based approach


In the stress based LEFM approach, a crack will propagate, when a critical stress intensity factor
KC , which is considered as a material property, is reached. Figure 2.3 shows a large panel with
a transverse crack of the length of 2a subjected to uniform tensile stress σ0 , which corresponds to
pure Mode 1 crack opening. In addition, the asymptotic stress field on the crack tip is depicted.
This is derived from the stress field around the crack tip in polar coordinates. Since the stresses
become infinite at the crack tip, a scaling factor for the stress field around the crack tip is induced,
which in fact is the stress intensity factor K. Equation 2.1 shows the stress intensity factor for the
example above. More derivations for specific crack cases can be found in [8, p.82 ff.].
σ0

σy

2a
x, r

z x

σ0
Figure 2.3: cracked infinite plate with asymptotic stress field at the crack tip


K1 = σ0 πa (2.1)
The critical value KC , where a crack starts to propagate, is dependent on the constraint. In plane
stress condition, higher values are achieved than in plane strain condition, where it converges to
a minimum value after a certain specimen thickness. This plateau-value is called the plane strain
fracture toughness KC and the value sought for. For characterization of the individual fracture
toughnesses of materials, standardized testing methods are provided by well-established institutes
for standardization [9, p.131], [8, p.103 f.].

2.3.2 Energy based approach


As mentioned in section 2.1 already, Griffith formulated a criterion for crack propagation based on
an energy balance which was augmented by Orowan. In this concept, the infinitesimal decrease of
potential energy Π of the specimen with respect to an increase in crack surface A, which is scaled
by the crack length a in the two dimensional case, is considered. This term is called the energy
2 Fundamentals in fracture mechanics 4

release rate G, depicted in equation 2.2, which can be seen as material parameter similarly to KIC
in the stress-based concept. Thereby Π is the sum of outer (external energies) and inner potential
energy (elastic strain energy), depicted in equation 2.3 [8, p.105].

∂Π
G=− (2.2)
∂a

Π = Πa + Πi (2.3)
In the case of a linear elastic material, the K-concept can be inserted which leads to a quadratic
connection between K and G, depicted in equation 2.4. Thereby E 0 corresponds to the Young’s
modulus in plane stress state and E 0 = 1−νE
2 in the case of plane strain condition. This assumption
is valid for pure Mode 1 as well as pure Mode 2 and pure Mode 3 loadings [8, p. 108].

KI2
G= (2.4)
E0
Similar to the stress-based approach, standardized testing-methods are provided by well-established
institutes for standardization. Comparing the deviations of these two approaches, one can imply
that the stress-based one is a more local method where the energy-based concept represents a global
definition. For the reason of easier calculation at discretized models such as finite element models,
the energy based approach is commonly preferred.

2.4 Mode mixture


In the general loading case more than one damage mode may occur simultaneously. Therefore the
behavior under mixed mode condition has to be covered by a rule since they influence each other.
This is particularly interesting in adhesive or interface layers as a result of a pre-defined crack
surface.
Many adhesive bonds or interfaces between laminated composites exhibit different fracture tough-
nesses in the individual damage modes and nonlinear mixed mode behavior.

A simple criterion is the Power-Law fracture criterion which is depicted in equation 2.5. The
exponents n1 , n2 and n3 are material constants which have to be obtained by mixed mode material
tests. When the fracture variable f reaches 1, the crack propagates. In addition, the exponents n1
to n3 can be unified to one uniform exponent α = n1 = n2 = n3 , which is done frequently. Figure 2.4
shows some 2D fracture curves for the Power-law approach. Thereby for α = 1, the criterion reduces
to the linear fracture criterion and for α = 2, an elliptical fracture curve is obtained.
n1 n2 n3
Gequiv G1 G2 G3
  
f= = + + (2.5)
GC,equiv G1C G2C G3C
2 Fundamentals in fracture mechanics 5

1 α = n1 = n2 = 1
α = n1 = n2 = 2
n1 = 2, n2 = 1
0.8

0.6
G2C
G2

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1
G1
G1C

Figure 2.4: example for the BK criterion

Another criterion which is very popular for delamination of laminated composite materials is the
BK-fracture criterion. It was developed by Benzeggagh and Kenane in 1996 [11]. Thereby an
equivalent critical energy release rate is formed and compared to the sum of energy release rates
GT , shown in equations 2.6 and 2.7. MMR indicates the mixed mode ratio between mode 1 and
mode 2. Additionally the mode 3 fracture toughness G3C is neglected since its value is similar to
G2C in the general case. The exponent η is a material constant which has to be obtained by mixed
mode tests. Similar to the Power-Law fracture criterion the crack propagates when the fracture
variable f reaches 1. Figure 2.5 shows the appearance of the BK criterion with realistic values,
normalized to G2C , which is usually higher than G1C .

G2 + G3
GT = G1 + G 2 + G 3 MMR = (2.6)
GT

GT GT
f= = (2.7)
GT C G1C + (G2C − G1C ) · MMRη

0.8
Gequiv
G2C

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1
MMR

Figure 2.5: example for the BK criterion


3 Fundamentals of laminated composite materials 6

3 Fundamentals of laminated composite materials


As the name implies, laminated composites consist of two or more different components which form
a heterogeneous structure. In general, the aim of composites is to obtain a material with better char-
acteristics than the individual components would have. Due to the different material properties and
high aspect-ratio (length to diameter) of the components in case of laminated composites, namely
fibers and matrix, they show highly anisotropic material properties. The fibers are responsible for
the high strength and stiffness of the material while the matrix acts as power transfer between the
individual fibers [12]. In nature this combination can be found in many bionic structures such as
wood or bones due to the superior strength- and stiffness-to-weight ratio which satisfies the main
target of bionic design: efficient material usage.

3.1 Material-composition and macroscopic structure


In the following section different materials and possible configurations of the individual components
are depicted.

3.1.1 Fiber-materials
Materials show highly increased material properties in fiber form than compared to the correspond-
ing bulk material, as depicted in figure 3.1. This is due to the sharp reduction of inner defects in
the fiber form. In many cases, another major contribution to these significantly higher properties
is the fact of molecular orientation. A good example are polymer fibers, which have randomly en-
tangled polymer chains aligned in fiber direction when produced in a spinning process and/or when
stretched, as shown in figure 3.2. This effect also leads to an increase in stiffness in fiber direction.

extrapolates to 11GPa
tensile strength / MPa

3000

2000

1000 extrapolates to approximate


strength of bulk glass (170MPa)

0 10 20
fiber thickness / µm
Figure 3.1: diameter-strength relationship by the example of glass [13]
3 Fundamentals of laminated composite materials 7

Figure 3.2: left: fiber pierced out of a polymer block; right: spinned fiber

The most important properties of fibers are high specific mechanical properties and compatibility
of its surface with the corresponding matrix. This is why e.g. Polyethylene fibers are not used in
composite laminates due to their highly inactive surface in untreated condition besides their rather
high specific properties at low cost.
In the following the most common fibers are reviewed in a short manner.

Carbon fibers

Carbon fibers consist of graphene layers in turbostratic


configuration, which are mainly oriented in fiberdirec-
tion as depicted in figure 3.3. The fact that graphene
contains delocalized electrons leads to electrical con-
ductivity, which can lead to corrosion when in con-
tact with metals and shielding of radio waves. Two
major precursors are available for producing carbon
fibers, Polyacrylnitrile-fibers (PAN) on one hand and
pitch-fibers on the other. Pitch fibers are less expens-
ive in production but show significantly lower tensile
strength. The properties of carbon fibers can be varied
in a broad range by thermal treatment which influ- Figure 3.3: turbostatic molecular struc-
ences carbon content and orientation of the strongest ture of carbon fibers [13]
carbon links along fiber direction [12, p.31]. Carbon
fibers show high-end properties at a rather high price.

Glass fibers

Glass fibers are drawn from specific glass melts - a blend of sand, limestone and other oxidic
compounds [12]. Their molecular structure is still amorphous and they show the glass-specific
properties such as inertness, high corrosion resistance and hardness. Several types of glassfibers
are existing with different strength but similar stiffness. The most common type thereby is E-glass
which shows good mechanical properties at favourable price.
3 Fundamentals of laminated composite materials 8

Aramid (DuPont Kevlar®) fibers

These fibers are based on spinned and stretched Aramid which is a liquid crystalline polymer
containing a stiff molecular structure as a result of many aromatic rings. Kevlar fibers show high
specific tensile strength and toughness but absorb moisture and degrade at UV-exposure.

Comparison of the most common fibers

Figure 3.4 shows stress-strain curves of high modulus- (HM-), high strength- (HS-) carbon, E- and
S-glass, boron and Kevlar fibers. It can be seen that the increased modulus of the HM-carbon
fiber leads to major drawbacks in strength versus the HS-carbon fibers. S-glass leads to overall
better performance compared to E-glass but at much higher price due to low production volumes.
Taking the density into account Kevlar fibers which seem to be unobtrusive in this chart become
very competitive to carbon-fibers. Summarized the design engineer has to keep the fiber types’
mechanical and economical advantages and drawbacks in mind for proper material selection at the
respective case of application.

tensile strength / GPa


HS
on
rb

4 ass
-gl
Ca

R)
49 S(
ron

r
la
ev ss
Bo
M

K la
E-g
nH
rb o
Ca

1 2 3 4
/%
Figure 3.4: comparison of selected fiber types [13]

3.1.2 Fiber configurations


The fibers can be embedded in the matrix in many different ways. This work is focused on unidirec-
tional (UD) and woven continuous fiber fabrics as shown in figure 3.5. Thereby bunches of these
individual fibers which have a thickness of only a few microns are aligned to rovings or twisted to
yarns. In woven fabrics these fiber-bundles are woven to fabrics similar to cloth. One individual
layer of fibers is called a ply. These are stacked further together layer-by-layer over each other to
a laminate. Thus, using different ply orientations with respect to the global laminate coordinate
system, the laminate can be tailored to the specific load case. Using UD fabrics, fiber contents of
70% can be reached in the laminate. Due to the warpage of fibers in woven fabrics, fiber contents
are limited to about 60% [14].
3 Fundamentals of laminated composite materials 9

Figure 3.5: left: UD fabric; right: woven fabric [14]

3.1.3 Matrix-materials
The main purpose of the matrix material is to form a strong bond with the fiber surface to hold
them together and transfer the loads to the fibers. It also has to carry transversal and interlam-
inar stresses. For manufacture, the matrix material has to be low-viscous with good wettability.
In general there are many different options for matrix materials such as thermoplastic polymers,
thermosets and other specialities. The most common matrix material for laminated composites are
epoxy resins due to their favourable properties in processing and good mechanical performance.
However as with other thermosets, recyclability is still a major drawback hence they cannot be
brought back to some sort of deformable state again without destroying them.

3.2 Damage mechanisms


The heterogeneity of composites leads to many different damage mechanisms. Failure can occur
intralaminar - this means inside of an individual ply - or interlaminar - damage between plies which
means that they separate from each other. Tree figure 3.6 visualizes the possible damage modes:

damage mechanism

intralaminar interlaminar

matrix damage

tension/shear
fiber damage matrix damage
delamination
tension compression tension compression/shear

rupture micro-buckling/shear-off cracking shear-off

Figure 3.6: damage mechanisms in laminated composites

There we can see that in the case of intralaminar failure fracture is caused by the fiber or matrix.
Furthermore the intralaminar damage mode is highly dependent on the direction of normal stress.
One has to keep in mind that the different damage modes can influence each other in progressive
failure. In the following paragraph the individual fracture modes on the basis of a laminate consisting
of UD plies are discussed. Thereby an individual coordinate-system is assigned to every ply which
3 Fundamentals of laminated composite materials 10

is aligned along the fiber direction (1), perpendicular to the fibers (2) and the thickness direction
(3), as shown in figure 3.7.

Figure 3.7: ply coordinate system [2]

As depicted, there are different nomenclatures available for the individual directions. In this work
the number-based one is used due to the fact that it is commonly used in anglo-saxon countries.

3.2.1 Intralaminar damage


Intralaminar failure can occur in the fiber, matrix or in the interface between them. In contrast to
metals, tension and compression loads must be treated differently. This is due to the nature of a
fiber, which cannot carry compressive loads on its own, similar to a rope or a cable.

Fiber failure

In the static tension loading case, fiber failure mainly occurs when the ply is loaded in 1-direction.
For the reason of a statistical fiber strength distribution, which commonly follows a Weibull-shape
[12, p.28], damage occurs gradually. This means that fracture starts at the weakest individual
fibers which causes stiffness-degradation. Furthermore at higher loads this leads to failure of whole
fiber-bundles. Therefore fiber-failure is not fully brittle but quasi-ductile [2, p.346].

When applying compressive loading in 1-direction, failure is triggered by the loss of stability similar
to the buckling of rods but on a micro-mechanical level. In contrast to rods where buckling occurs
in bending mode in FRP materials, it takes place in shear mode due to the low shear-stiffness of
the material. This mechanism is called micro-buckling, which is depicted in figure 3.8.

Figure 3.8: micro-shear-buckling of fibers under compression load in 1-direction [2, p.351]
3 Fundamentals of laminated composite materials 11

This form of failure can occur in in-plane direction but also in out-of-plane direction at certain
layups. Due to the high dependency of the compressive strength on the fiber-angle, the fibers have
to be aligned in loading-direction as perfect as possible without any waviness to assure maximum
strength values. To avoid this failure mode high accuracy as well as exact and careful handling
during the whole production process is demanded.

In-plane shear-stress also reduces the compressive strength in 1-direction, R11 . But one has to
keep in mind that most composite structures are thin-walled and therefore global loss of stability
generally occurs before the described micro-buckling. Critical areas for micro-buckling are edges
of holes where the fibers buckle into the hole due to stress increase near the hole or boundaries
at bending-beams which are bended by transversal loads due to the shear-stresses and the lack of
support on the edges [2, p.362].

Matrix failure

Matrix failure or in-plane inter-fiber-failure (iFF) is a more complex phenomenon. Generally one
has to separate between the action plane which is the plane of maximum load in the material and
the fracture plane. These planes may not fall together in some cases as A. Puck discovered in 1992
in UD layers [2].

In the case of tension in 2-direction, action- and fracture plane fall together which means that the
fracture plane is parallel to the fibers in the matrix. This occurs already at low strains of about
0.5% due to the high difference of matrix- and fiber-modulus. This for example can occur in a cross
ply laminate, which consists of only 0° and 90° plies. When applying tensional loading in global
1-direction thus along the fibers of the 0° ply micro cracks will occur in the 90° at low strains long
before the 0° ply will fail [15, p.68].
Under pure compression load in 2-direction, the fracture plane is slightly above 45°, similar to
concrete. Facing in-plane shear loading τ12 , there are two possible action planes due to the associated
shear stresses which are the 13-plane or the 23-plane. Due to the much lower fracture resistance of
the matrix in comparison to the fibers, fracture occurs parallel to the fibers as depicted in figure 3.9.
Same behavior can be obtained from out-of-plane shear in the 13-plane.

Figure 3.9: fracture-plane under in-plane shear loading [2, p.366]

Finally shear in the 23-plane causes fracture under 45° normal to maximum principal stress direction
due to matrix tension failure.

More information about this topic can be found in literature [15] and [2].

3.2.2 Interlaminar damage


In layered composite structures debonding of the individual plies can take place. This damage mode
is induced by interlaminar stresses such as normal stress in thickness direction σ3 and out-of-plane
3 Fundamentals of laminated composite materials 12

shear stresses τ13 and τ23 . It is similar to intra-laminar matrix fracture but with planar propagation
resulting due to the absence of crack arresters such as fibers. As a consequence they can impact
the laminate properties significantly due to high spreading. In many cases delamination is caused
by iFF-damage (from an impact for example) or resulting from manufacturing imperfections such
as voids. Drilling also produce delaminations between the top- and the bottom layers as a result of
peel-up and push-out mechanisms as it is shown in figure 3.10 [16].

Figure 3.10: left: peel-up delamination; right: push-out delamination [16]

Another causes of delamination can be ply-dropoffs, free edges or curved sections which induce
interlaminar stresses. In the case of free edges at the example of a cross-ply (consists of 0°- and
90°-plies) under tension in laminate 1-direction, contraction of the 90°-layers is restrained due to the
stiff fibers. This leads to peel-stresses at the outer edges of the 0°-layers as we can see in figure 3.11.
In curved sections tension stresses in radial direction appear when the specimen is bended contrary
to the curvature, which lead to delamination([2, 17]), depicted in figure 3.12.

Figure 3.11: interlaminar stresses caused by restrained transverse contraction [2, p.388]

σr

F F

Figure 3.12: interlaminar radial stresses in bended curved beams

Considering an existing interlaminar pre-crack in a laminate, all of these phenomena can be divided
into three damage modes, as mentioned in section 2.2. These are the opening mode 1, the in-plane
3 Fundamentals of laminated composite materials 13

shear mode 2 and the out-of-plane shear mode 3.


Contrary to metals where crack growth occurs primarily perpendicular to the maximum principal
stress direction (mode 1), in composite materials a damage propagates along the interface layer
due to the heterogeneous, layered structure. Therefore in realistic load cases mode mixing of the
individual crack modes occurs which leads to the question of a proper mathematical description in
fracture mechanic approaches. Further details are described in section 3.3.2.

3.3 General 3D modeling approach in Finite Element Analysis


For modeling laminated composites, several possibilities are available. One can model the whole
laminate, ply-stacks or individual plies with or without interfaces in between. Higher modeling
accuracy results in more precise representation of the real structure but significantly increases
calculation time. Therefore one has to find a compromise between good representation of the
structure and needed simplifications which leads to proper modeling strategy selection for the present
problem. The work focuses on three-dimensional (3D) modeling techniques for the reason of the
applicability of techniques for interface damage modeling. Therefore, solid and continuum shell
elements will be used.

3.3.1 Element types for laminate modeling


In this section, suitable element types for 3D-modeling of laminated composites are reviewed. In
general, elements can use linear shape functions resulting in one node at each corner or they can
use quadratic shape functions which results in an additional node at all edges. Thereby former
elements give a linear displacement distribution over the edge and latter elements give quadratic
displacement distributions over the edge which improves their accuracy substantially.
Stresses are calculated in integration points. In addition there is an option of reduced integration
resulting in a decrease in calculation time but generating the problem of hour-glassing at first-
order elements in some cases. Additional information can be found in literature and FEM-package
manuals [18].

Shell elements

When modeling with shell elements, a thin-walled structure is reduced to a two-dimensional refer-
ence surface - the mid surface in the common case. Figure 3.13 shows a structural body and the
corresponding discretization to a shell with linear quadrilateral elements. Additionally triangular
elements are available for complex shapes. However these need higher mesh refinement for the same
level of accuracy. This results from the natural shearlocking in triangular structure statically de-
termined shape (which is used in frameworks for example) and that leads to an undesired increase
in element stiffness.
3 Fundamentals of laminated composite materials 14

structural body shell discretization

Figure 3.13: shell discretization of a structural body

The thickness of a shell element is defined through the section property definition, thereby the
positive thickness direction has to be determined. In general shell elements have displacement
and rotational degrees of freedom. For calculating the plate behavior, these kind of elements have a
selectable number of section points over the thickness. In addition, some of them include deformation
induced by transverse shear stresses, which is important at laminated composite structures for
example or can be used for large deformations. In fact shell elements lead to efficient calculation
due to a high level of simplification. However, using them for discretizing more complex structural
parts can lead to high errors between simulation and reality as a result of oversimplification. This
is caused by reducing the problem to plane stress or plane strain condition in thickness direction.

Solid elements

For full three dimensional discretization of complex structural bodies solid elements are available.
Because they are defined in all three coordinate directions, three dimensional stress states can
be calculated properly. However analyses with solid elements are much more time consuming,
especially with quadratic formulation, since significantly more elements are needed for discretization.
Furthermore in the case of bending problems more than one element should be used over the
thickness for proper stress calculation. For more complex shapes tetrahedral elements are available.
However as well as in the two dimensional case these need higher mesh refinement as a result of
these elements’ natural stiffening. Figure 3.14 shows a structural body and the corresponding solid
discretization with linear hexaedral elements.

structural body solid discretization

Figure 3.14: solid discretization of a structural body

Continuum shell elements

Continuum shell elements, which are available in Abaqus for example, base on a three-dimensional
body. These elements have displacement degrees of freedom only. They look like three-dimensional
3 Fundamentals of laminated composite materials 15

solid elements but behave similar to conventional shell elements which leads to reduced calculation
time compared to solid elements. In addition the bending behavior is included analogous to shell
elements so only one element over the thickness is needed. Shear deformation is covered by first
order shear deformation theory (FSDT), which makes them suitable for laminated composites since
they exhibit low shear stiffness through the thickness. Furthermore they include thickness change.
Other than shell elements, they can be stacked to provide a more refined through-thickness response
[18].

3.3.2 Techniques for interface damage modeling


For including interlaminar damage behavior when modeling laminated composite structures, the
interfaces between the layers have to be modeled separately. Therefore the two most common
techniques for modeling interface damage in finite element models are presented in the following.

Virtual Crack Closure Technique

The virtual crack closure technique (VCCT) is based on linear elastic fracture mechanics (LEFM,
see section 2.3) and can be used for the simulation of delamination phenomena in layered structures
such as laminated composite materials. Figure 3.15 depicts a 2D finite element model using linear
quad elements with two partially connected layers, an upper layer u and a lower layer l. Additionally
there is a cracktip at the overlapping of the connected nodes n of each layer.
a ∆a ∆a ∆a

lu Znu
mu
n o p z
∆wm Xnl
Xnu
ml Znl
ll
∆um x

Figure 3.15: VCCT for four-noded 2D elements

In general the assumption of crack closure techniques is that the energy ∆E released when opening
a crack by the length of ∆a is equal to the energy for closing the same crack. Therefore the acting
forces Zn and Xn (Znu = Znl equilibrium, same with Xn ) on each layer’s node n are calculated with
a finite element analysis. It is obvious that Zn induces delamination in mode 1 and Xn in mode 2.
In general a second calculation is needed where the bonding of the nodes n is released in order to
calculate the opening displacement of the crack tip.
This is done in the crack closure method [19, p.3]. However, in the virtual crack closure technique
(or modified crack closure method), it is assumed that a crack extension of ∆a from a + ∆a to
a + 2∆a (from node n to node o) does not significantly alter the state at the crack tip. Therefore
the displacements at separation of the nodes n is assumed to be equal to the displacements of nodes
3 Fundamentals of laminated composite materials 16

mu and ml . This leads to the following equation for the work required to close the crack for the
two-dimensional case:
1
∆E = · [Xn · ∆um + Zn · ∆wm ] (3.1)
2
For calculating the energy release rates G = ∆E
∆A , different formulas are obtained for different element
types [19]. In the case of linear 2D quad elements which are used above in figure 3.15, G1 and G2
are calculated as shown in equations 3.2. For other element types the author refers to the literature,
in particular [19].

1 1
G1 = − · Zn · ∆wm G2 = − · Xn · ∆um (3.2)
2∆a 2∆a
The calculated energy release rates are compared to the corresponding fracture toughnesses to verify
any crack onset or crack propagation, see sections 2.4 and 4.4. If the criterion is fulfilled, the node
is released and the procedure repeats for the next crack tip node. Alternatively, an appropriate
number of cycles until the crack has propagated to the next node of a finite element mesh can be
calculated.

Cohesive Zone Modeling

Cohesive zone models (CZM) base on independent works of Dugdale [20] and Barenblatt [21], who
augmented linear elastic fracture mechanics by nonsingular solutions for the opening displacement
and traction fields of the crack faces at the crack tip. Therefore Dugdale used elastic perfectly plastic
material behavior, which generates a plastic zone in front of the crack tip, that virtually elongates
the crack tip by a certain distance d, dependent on a critical crack tip opening displacement δC
[8, p.164]. Barenblatt looked for specific shapes of the crack tip, in which the singularity vanishes,
hence the stresses on the crack tip become finite. Despite the different initial approaches, the same
solution is obtained.

Figure 3.16 shows a cohesive zone model applied on a mode 1 crack and figure 3.17 a simple, bilinear
cohesive law with linear stiffness degradation after a certain σ0 , which corresponds to the damage
of the cohesive zone. There, the elastic region in former figure corresponds to the linear rising slope
at the beginning in latter, until δ0 is reached. After exceeding δ0 , the cohesive zone is damaged
and the stiffness degrades linearly, which is described by K(D). Thereby the red line in figure 3.17
indicates the loading and unloading path of a damaged area cohesive zone. When δC is reached,
the damage reaches 1 and the cohesive stiffness drops to 0, which means that the bond separates.
Comparison of this theory to elastic plastic fracture mechanics (EPFM) shows that the area under
the traction law represents the critical strain energy release rate GC (see [8, 168 ff.]). When δC is
equal to δ0 , which corresponds to ideal elastic material followed by a instant stiffness drop to 0, the
inelastic damage process zone in figure 3.16 vanishes and hence, an ideal brittle material response
as in energy based LEFM (see section 2.3.2) is obtained.
3 Fundamentals of laminated composite materials 17

σ(x)

δC

a d z
true crack damage process zone elastic region
D=1 0<D<1 D=0

Figure 3.16: cohesive zone at a mode 1 crack

D=0
D>0
σ
σ0
GC

K(D) = (1 − D)K0 D=1


K0
1
1 δ
δ0 δC
traction separated
Figure 3.17: bilinear equivalent one dimensional cohesive traction law

Many other different traction laws were created to represent specific material behavior, such as
the exponential law depicted in figure 3.18. In general, the point of onset of damage as well as
the unloading behavior of the damaged material can be set individually. The red line depicts a
suggestion for the loading and unloading behavior in damaged state at the exponential traction law.
However, the unloading behavior can differ from the loading behavior, which induces hystereses.
3 Fundamentals of laminated composite materials 18

D=0

σ D>0
σ0

GC

δ
δ0 δC
traction separated
Figure 3.18: exponential one dimensional cohesive traction law

Mixed mode is treated by interpolating between traction laws of the same type for mode 1 and
mode 2/3, as it can be seen in figure 3.19, where simple quadratic interpolation is used:

δ1C
δ1

δmC
δ2C
δ2 δm

Figure 3.19: mixed mode treatment for cohesive zones [22, p. 2-805]

As a conclusion, cracks are not modeled as discontinuities due to the reason that the model is
based on continuum damage mechanics. Therefore, other than in VCCT, no initial crack is needed.
This means that the undamaged interface and the crack initiation can be modeled. However, an
appropriate cohesive traction law with stiffness and (un-) loading behavior under damaged condition
has to be determined. Thus, for simple bilinear traction laws, Turon et al. created guidelines for
obtaining proper parameters for numerical FEM simulations in [23].
4 Fundamentals in fatigue 19

4 Fundamentals in fatigue
4.1 Fatigue: History and general definition
In the 19th century when railways came up, unexpected failure of wagon axles after a certain time
of use lead to first fatigue experiments by August Wöhler. This lead to the famous S-N or "Wöhler-"
curves, where the applied loading was described over the number of cycles to failure for metals. In
general, fatigue means the impairment, crack initiation and propagation under repeated loading.
Thereby the load level is lower then in the static case. The loading can be periodic, aperiodic,
deterministic or even stochastic. In high cycle fatigue (Number of cycles to failure over 1 · 104 ),
rupture commonly occurs without high plastic deformation even at ductile materials. Figure 4.1
shows an example catastrophic failure of a turbine shaft due to high cycle fatigue loading, where
crack propagation occurred from the inside. Therefore fatigue is a critical failure mode in cyclic
loaded structural parts.
Additional information can be found in literature and [24–26].

Figure 4.1: turbine shaft of a 300MW steam turbine set [27]

4.2 Design philosophies


When designing a structure, the engineer has to choose between two fundamental design philosophies
[28]:

ˆ safe-life concept: This concept acts on the restriction of no damage in the whole lifetime in
a structural part. Therefore, high safety factors and over-dimensioning is needed in general.

ˆ fail-safe concept: In the fail-safe or damage-tolerant concept, the structural component is


designed as a redundant system, in which flaws in a structure do not induce failure of the
whole component to a certain extent over the lifetime. This means, that cracks are allowed
to a certain size.

The fail-safe concept is of special interest in composite structures, since other than in metals, the
stage of damage initiation is rather short compared to the stage of propagation (see section 4.6).
This results in excessive over-dimensioning when a safe-life concept is used in the dimensioning of
composite structures.
4 Fundamentals in fatigue 20

4.3 Loading conditions


For the characterization of a certain periodic loading, significant variables are introduced. In fig-
ure 4.2, a sinusoidal loading σ(t) with the cycle duration T and specific levels of maximum and
minimum stress σmax and σmin is shown. From these, further parameters, namely the mean and
amplitude stress σm and σa can be derived, as depicted in equations 4.1:

cyclic stress σ(t) σmax

σm

σmin

0 T 2T
time t

Figure 4.2: example for a periodic sinusoidal loading

σmin + σmax σmax − σmin


σm = σa = (4.1)
2 2
Furthermore another representative variable for describing the loading condition - the stress ratio
R - can be calculated as shown in equation 4.2:
σmin
R= (4.2)
σmax
Thereby specific values are obtained for different loading conditions:

ˆ between 0 < R < 1, the specimen is subjected to tension-tension loading

ˆ between 1 < R < ∞, the specimen is subjected to compression-compression loading

ˆ between -∞ < R < 0, the specimen is subjected to tension-compression loading

ˆ at R=0, the specimen is subjected to pulsating tension loading

ˆ at R=±∞, the specimen is subjected to pulsating compression loading

These five variables are dependent on each other, leaving two degrees of freedom and represent a
certain fatigue loading condition. Classifications of more general loadings, which usually occur in
real applications, are primarily based on these (see section 4.5.1).
4 Fundamentals in fatigue 21

4.4 Fatigue damage evolution


In metals, fatigue crack growth starts with microscopic cracks, which emerge for example from
dislocations. Multiple microscopic cracks form a macroscopic, sharp crack of a length of about
1mm. This procedure is called damage or crack initiation.
Dependent on the level of the applied cyclic loading, an existing macroscopic crack is faced to
a certain rate of crack growth then, when a certain threshold level is reached. Under moderate
loadings, it grows to a critical length until the crack growth becomes unstable and final rupture
occurs. Figure 4.3 shows these individual stages of fatigue damage evolution at the example of a
simple metal component, which is subjected to a cyclic fatigue loading F (t):
F(t) F(t)

crack
initiation

F(t) F(t)

on
g ati
a
prop
F(t)

final stable
rupture propagation unstable
propagation

F(t)

Figure 4.3: stages of fatigue damage evolution

As depicted, in metals, one or only a few dominant macroscopic cracks control the fatigue life of
a structural component in general. Therefore, e.g. in aeronautical structures, emerging cracks are
monitored to estimate the remaining fatigue life. However, this is not the case in every material, as
it will be shown for laminated composites in section 4.6.

4.4.1 Damage initiation


In general, damage initiation, also called damage nucleation, is treated by S/N-curves, in which the
maximum allowed stress amplitude over the number of cycles is depicted for specific levels of the
mean stress. For statistical reasons, the curve has to be considered as a band with high probability
of failure on the upper border and low probability of failure on the lower boarder, which follows a
logarithmic normal distribution in general case. Therefore, every S/N-curve has a certain survival
probability, which is 95% in most cases [24, p.44 ff.]. Regarding the shape of the S/N-curve, two
different behaviors can be observed:
ˆ Type 1 behavior, in which after a certain number of cycles, which is 1 · 106 to 1 · 107 cycles
in normal case, the maximum allowed stress amplitude reaches a plateau, which is the fatigue
endurance limit. This occurs in low-alloy steels and titanium for example.
4 Fundamentals in fatigue 22

ˆ Type 2 behavior, in which the degradation is also reduced after a certain number of cycles,
but does not reach a plateau. However, to define an endurance limit with high reliability, a
ultimate cycle number of 2 · 106 to 1 · 109 cycles is commonly taken [24, p.20].

Figure 4.4 shows these two types of S/N-curves with the regions of low cycle fatigue (LCF) to about
1 · 104 and high cycle fatigue (HCF) from there on:

log(σA ) log(σA )

σE σE
LCF HCF LCF HCF
Ne log(N ) Ne log(N )
Figure 4.4: types of S/N-curves: Type 1 (left), Type 2 (right) [25, p.361]

For low cycle fatigue applications of ductile metals, which is up to about 5 · 105 cycles, /N-curves
are frequently used since they better represent the occuring elasto-plastic deformations [24, p.33
ff.].
For the numerical description of S/N-curves, many researchers such as Basquin, Stromeyer, Palmgren,
Bastenaire or Stüssi, proposed suitable models. Equation 4.3 shows Basquin’s law, which represents
the polynomial decrease of the maximum stress amplitude over the cycle number in the region of
finite life fatigue strength at a given mean stress. Thereby σe indicates the fatigue strength of the
amplitude whereas Ne expresses the endurance limit and k controls the slope of the curve.
1
Ne

k
σa = σe · (4.3)
N
Multiple S/N-curves with different mean stresses can be assembled together to constant life diagrams
(CLD), for example according to Smith or Haigh, then. Figure 4.5 shows a qualitative example of
a constant life diagram according to Haigh with a linear curve according to Goodman and equal
behavior on positive and negative mean stress:

(R=-1) stress amplitude σa


σy

σe N=1
N=104
N=105
N=Ne =107
(R=1) −σy 0 σy (R=1)
mean stress σm

Figure 4.5: linear constant life diagram according to Haigh


4 Fundamentals in fatigue 23

Thereby the line for N=1 indicates the static strength σy whereas Ne shows the endurance limit
with σa = σe at σm = 0 (R=-1). Other curve shapes are proposed by Soderberg, Gerber or Morrow
for example. More information can be found in literature and [24–26].

4.4.2 Onset of propagation


For the onset of propagation of an existing flaw, Murri, Salpekar and O’Brien proposed a method
for determining the timespan for the formation of a macroscopic, sharp crack for mode 1 crack
growth [29]. Similar to a S/N-curve, dependent on the number of cycles, a certain maximum strain
energy release rate is needed for crack onset, as shown in equation 4.4:

G1,max = c · N d (4.4)
Thereby the constants c and d are material constants. A standardized testing method for the
measurement of the onset of propagation in mode 1 is presented in ASTM D 6115 - 97. As stated
in [30], the initial condition of the crack tip is essential for the onset of damage. This means,
that the onset differs significantly, if e.g. a specimen is pre-cracked or not. A similar crack onset
detection criterion is implemented in the Abaqus low cycle fatigue criterion, which is described in
section 6.2.1.

4.4.3 Damage propagation


After the formation of a sharp, macroscopic crack while being subjected to a sub-critical fatigue
loading, crack propagation occurs. Figure 4.6 depicts the rate of crack growth over the effective
stress intensity factor ∆K = Kmax − Kmin on a double logarithmic plot. It can be seen that the
curve can be partitioned into three significant regions:

ˆ region I, in which crack growth slowly starts at ∆Kth = Kth − Kmin , with a rising rate of
crack growth with increasing ∆K
ˆ region II, where stable crack growth occurs
ˆ region III, where accelerated crack growth occurs until reaching the relative static critical
strain energy release rate ∆KC = KC − Kmin

 
da
log dN

I II III

∆Kth ∆KC log(∆K)

Figure 4.6: qualitative crack growth curve of a macroscopic crack [25, p.356]
4 Fundamentals in fatigue 24

For region II, Paris and Erdogan [31] found a power law relationship, in which the crack growth
rate is dependent on the relative stress intensity factor ∆K = Kmax − Kmin in a cycle, depicted in
equation 4.5. There, C and m are material constants.

da
= C · (∆K)m (4.5)
dN
Since the so-called Paris law is only valid for a given stress ratio R, the model was expanded
by several researches to cover the effect of the stress ratio, for example by Walker for aluminum
alloys [32]. In addition, many other proposals and modifications are given by researchers to cover
additional effects such as the non-polynomic regions near the threshold and critical value, load
frequency, temperature and so on, which can be found in literature.

4.5 Treatment of general load spectra


In most real applications, a structural element is subjected to variable amplitude loadings, which
can be periodic, aperiodic or even random. In addition, the mean loadings can be variable as well.
Therefore, a load-time function has to be classified to load spectra of constant mean and amplitude
loading and their individual contributions to the damage have to be considered to accurately estim-
ate the fatigue life. In the following, common techniques for classification and damage accumulation
are discussed.

4.5.1 Classification of general load spectra


For the classification of general load time functions, many methods were developed. One-parameter
classification methods focus onto the classification into certain load regions with respect to a certain
attribute such as the peak value or when the function passes a load region. However, since stress
ratio effects are not covered by these methods, two-parameter classification methods were created.
Other than in one-parameter methods, mean and amplitude loadings can be transformed back from
the classification results. The most popular of them is the Rainflow-counting method for the reason
of a physical background in closed stress-strain hysteresis loops. With it, a transition matrix is
obtained, which can be formed into a more demonstrative amplitude-mean value matrix, containing
the relative occurrences. With the help of a Haigh diagram and the Miner’s rule for linear damage
accumulation (see section 4.5.2), the individual fractions of damage can be calculated and added
together. When subjected to subsequent stress patterns, these matrices can be added together.
Several researchers proposed modifications to overcome drawbacks such as sequence effects or open
hysteresis loops. Additional information can be found in literature, in particular [24].

4.5.2 Damage accumulation


To obtain the total damage to a given number of cycles, the individual damages of the load spectra
have to be estimated and added together. The most popular damage accumulation rule is the
Palmgren-Miner rule, where the incremental damage, ∆D, is estimated by comparing the number
of loading cycles, N , to the number of cycles to failure according to the corresponding S/N-curve,
Nf , as done in equation 4.6:

∆N
∆D = (4.6)
Nf
P
Fracture occurs, when ∆D = 1. As it can be seen, the Palmgren-Miner rule is linear. However, it
does not account for sequence effects and interactions between the damage and the fatigue strength
obtained from the S/N-curve. Therefore, many researchers provided proposals for nonlinear damage
accumulation rules, which however need additional parameters to be identified [24, p.293 ff.].
4 Fundamentals in fatigue 25

4.6 Comparison of fatigue behavior of metals and laminated composites


In metals, the crack initiation phase is usually the dominant timespan in fatigue. After crack
initiation, one or only a few cracks dominate the crack propagation phase. These macroscopic
cracks commonly propagate normal to principal stress direction at the crack tip, which is mode 1
crack propagation (compare section 2.2). In laminated composites, many microscopic cracks emerge
from voids and weak fiber-matrix bonds in the matrix already at low cycle numbers, stopped by
neighboring fibers. This leads to local stiffness degradation, which causes redistribution of load
paths and further stiffness degradation in other regions. When the matrix is saturated with a
certain amount of microscopic cracks, macroscopic cracks accumulate, which cause final failure of
the structural part [25]. Since the formation of these macroscopic cracks occurs over a wide range
of the fatigue life, crack propagation is the dominant mechanism in laminated composites. In
addition, interlaminar matrix damage - namely delamination (compare section 3.2.2) - can emerge
from intralaminar matrix cracks. These propagate at rather high speed, since in between the layers,
there are usually no crack arresters such as fibers. Furthermore delamination and intra-laminar
fiber-matrix debonding decrease the supporting effect of the fibers in compression loading, which
causes micro-buckling. Thus, interlaminar damage is of particular interest in fatigue.
5 State-of-the-art fatigue damage modeling techniques of laminated composites 26

5 State-of-the-art fatigue damage modeling techniques of lamin-


ated composites
To overcome the peculiar, different fatigue behavior of laminated composite materials, many ap-
proaches were already created. Since these overlap in many cases, differentiation into certain classes
is a difficult task. Therefore, a rather general classification is presented in the following.

5.1 Laminate and lamina fatigue life estimation


The first and oldest class of composite fatigue models are leaned on the same treatment which is
usually taken in metal fatigue. There, S/N-curves and constant life diagrams (CLD) are measured
for a certain laminate with a defined ply stacking sequence (PSS). Since laminated composites show
highly nonlinear dependency of the stress ratio, as depicted in figure 5.1, piecewise linear CLD using
S/N-curves for several stress ratios lead to a better representation. For a continuous estimation over
the whole stress ratio spectrum for the reduction of experimental tests, some researchers tried to
describe the effect of stress ratio by fitting functions for the Haigh diagram. However, they are not
consistent over several material combinations and therefore, piecewise linear CLD with sufficient
amount of stress ratio data is the most accurate choice in general [26, p. 127f]. Since these CLDs
are only valid for a certain laminate with a specific PSS and material combination, a huge amount
of material testing is needed.

Figure 5.1: piecewise CLD with interpolated lines for a [90/0/±45/0]S E-glass/polyester laminate [33, p.16]

Other approaches for estimating fatigue life are based on residual values of the laminate strength
or stiffness, as depicted in figure 5.2:
5 State-of-the-art fatigue damage modeling techniques of laminated composites 27

Figure 5.2: schematic degradation of strength and stiffness during constant amplitude fatigue loading [33,
p.10]

In the case of residual strength models, the specimen has to break for derivation of the residual
strength, which induces a high amount of destructive testing needed, similar to CLDs. This results
from the need of loading the specimens to certain number of cycles, followed by testing the remaining
fatigue strength for every single point of the curve. Furthermore, residual strength remains rather
constant until the end of fatigue life in laminated composite materials, where it suddenly decays
rapidly (sudden-death phenomenon). Residual stiffness usually changes significantly in earlier stages
of fatigue life and can be measured by non-destructive testing as it can be seen in figure 5.2. The
curve can be divided into three regions: significant initial stiffness drop in early fatigue life, followed
by little decrease over a wide range, until it drops significantly again in the region of final frac-
ture. Furthermore, stiffness measurements show less scatter than strength based data. In fatigue
tests with several specimen, for certain residual stiffnesses, probabilities of failure can be alloc-
ated. In [34], these residual stiffness curves were used to derivate S/N-curves. These "Sc/N-curves"
significantly reduce testing effort which collaborating well with conventionally measured S/N-curves.

Since multi-axial stress states are not covered by classical S/N approaches and the derivation of
equivalent stresses according to e.g. Mises or Tresca is not permissible for anisotropic materials,
many researchers proposed multi-axial fatigue failure criteria. There, many of them, e.g. Hashin-
Rotem [35], are based on S/N curves. However, most of them are limited to specific laminate types
such as UD, cross-ply or angle-ply.
Other approaches are extensions of static intralaminar failure criteria such as Tsai-Wu [36] or base
on strain energy density (e.g. Plumtree and Cheng [37]). Summaries of the wide spectrum of fatigue
life failure criteria can be found in literature, in particular [38, 39].
5 State-of-the-art fatigue damage modeling techniques of laminated composites 28

5.2 Progressive damage models


Since stiffness degrades significantly over fatigue life as shown in section 5.1, stress redistributions
occur in a structural component. Critical regions are partially unloaded as a result of fatigue
induced stiffness degradation. Consequently, incremental approaches with recalculation of the finite
element model after a certain increase in damage lead to better fatigue life estimation in complex
structural components. Therefore, many researchers proposed models which correlate the damage
growth with residual mechanical properties. These are based on continuum damage mechanics,
thermodynamics, micro-mechanical failure criteria or discrete damage characteristics such as crack
spacing. A summary of progressive damage models can be found in [38].

5.3 Interlaminar fatigue damage models


In cases where delamination is the critical damage mechanism or in advanced, ply based fatigue
approaches, interlaminar fatigue damage models are used frequently. These models count to the
progressive damage models, since the gradual increase of a discrete damage is monitored. In the
following, the most common techniques for interface fatigue damage modeling are presented.

5.3.1 LEFM methods


Since LEFM methods can only be applied on an existing crack, delamination initiation cannot be
simulated. For the calculation of crack propagation, the stress intensity factors or strain energy
release rates obtained from analytical or numerical calculations (see section 3.3.2), the rate of crack
propagation is generally calculated with a Paris relationship (see section 4.4.3). For the onset
of propagation of an existing flaw, some researchers proposed crack onset criteria to estimate the
number of cycles needed for the formation of a sharp crack, which is called onset of crack propagation
(see section 4.4.2).

5.3.2 Cohesive zone methods


In the field of cohesive zone modeling, two general approaches are used: the load envelope models
and the loading-unloading hysteresis damage models. In former, only the maximum load of a
load cycle is considered. Load variation is implemented by pre-defined scalars and the number of
cycles is interpreted as a continuous, differentiable variable as well as the damage, which develops
continuously. Damage evolution dN dD
is summed up from a quasi-static cohesive law dDdN and a fatigue
s

dDf
damage rate dN and integrated over ∆N , which however is not trivial. Since the fatigue damage
induces stiffness reduction in the damage process zone, the crack elongates by inducing quasi-static
damage ahead of the crack front to restore static equilibrium again. Therefore, the development
of the quasi-static damage is unknown over the cycle jump, which is a source of error in finite
element simulations using load envelope models. In the loading-unloading hysteresis methods,
the complete cyclic variation is modeled and the damage increases steadily and unrecoverable,
which induces stiffness degradation. For the reason of point-wise formulation, these models show
easy implementation in finite element codes. In general, these models have different constitutive
stiffnesses for opening and closing displacement. If the quasi-static cohesive law, which can be
dependent on the load history, is reached, the traction develops according to it. To reduce calculation
times in finite element models, damage is extrapolated after simulating a few loading cycles. Detailed
descriptions of these models can be found in [40].
6 Theories behind selected fatigue models in FEA-packages 29

6 Theories behind selected fatigue models in FEA-packages


Since fatigue damage modeling is still at the very beginning in composite materials, some of the
most promising approaches proposed by FEA developers were chosen in this work for examina-
tion regarding their theories behind and tested in the corresponding finite element packages - if
implemented in it already. The following section focuses on the theory part.

6.1 Siemens Samtech Samcef: intralaminar fatigue damage modeling of woven


and UD FRP
Samcef’s high cycle fatigue model is focused on intralaminar fatigue damage only (see [41] and
[42]). It is based on the work of Wim Van Paepegem [43], who developed a fatigue model based on
phenomenological residual stiffness of the individual plies. In addition, plastic deformations from
fatigue loadings are taken into account. Variable amplitude and multi-axial loading is treated based
on Brokate’s damage hysteresis operator (see [44]), which is connected to the Rainflow-counting
method and Palmgren-Miner damage accumulation.

6.1.1 General modeling approach


The model was originally developed for biaxially woven textiles, since their response is equal in
both principal directions. There, the linear stress-strain relationship, depicted in equation 6.1, is
extended by a damage tensor H containing the individual damage variables D1 , D2 and D12 , which
ensures degradation of the in-plane stiffness matrix, C. In addition, a plastic strain vector ~p is
applied for taking plastic deformations, which result from fatigue loadings, into account.

~σ = H · C · H · (~ − ~p ) (6.1)

√ 
1 − D1 0 0 0 0 0

1 − D2
 

 0 0 0 0 0 

 0 0 1 0 0 0 
H=
 


 0 0 0 1 0 0 

 0 0 0 0 1 0 
 √ 
0 0 0 0 0 1 − D12
 
C11 C12 C13 0 0 0
 
C21 C22 C23 0 0 0 
 
C
 31 C32 C33 0 0 0 
C=


 0
 0 0 C44 0 0 

 0
 0 0 0 C55 0 
0 0 0 0 0 C66

For the estimation of the damage growth, damage dependent failure indices were created, which
describe the relative loading. These base on the Tsai-Wu static failure criterion for composites
[36] to cover multiaxiality. The 2D failure indices are derived by replacing one σi entrance with
σ̃i σi
Σi , where σ̃i = 1−Di is the effective stress due to damage, in the Tsai-Wu criterion. Solving the
equation gives the corresponding failure index Σ2Di . Since all Tsai-Wu based failure indices rise close
to 1 in the region of the failure envelope - independent of the main stress applied - the final failure
indices Σi are modified by the one dimensional failure indices Σ1D i , which are the fraction between
6 Theories behind selected fatigue models in FEA-packages 30

effective stress in a certain direction over the corresponding static strength, hence base on a simple
maximum stress criterion. The resulting definition is shown in equation 6.2:

Σ2D
i
Σi = (6.2)
1 + (Σ2D 1D
i − Σi )
Figure 6.1 shows a visualization of the procedure for obtaining the 1D and 2D damage dependent
failure indices:
σ̃22
Σ1D
22

σ̃22
Σ2D
2

σ22 σ̃11
Σ2D
11 σ̃11
(σ̃11 , σ̃22 )
Σ1D
11

(σ11 , σ22 )

σ11

Figure 6.1: damage dependent failure indices [45]

The fatigue damage laws were originally created for plain woven fabrics, which simplifies the problem
since these exhibit the same behavior in both principal directions. Equations 6.3 show the fatigue
damage laws in 1- (i=1) and 2-direction (i=2) from [46]. It represents a continuous curve with a
shape that is leaned on an inverted residual stiffness curve in a composite (see figure 5.2). Thereby,
c1 and c2 form damage initiation, which induces the first stiffness drop. The constant c3 describes
damage propagation, which is the region of nearly linear stiffness degradation. Finally the constants
c4 and c5 are responsible for highly accelerated crack growth, when final failure due to the formation
of macroscopic cracks occurs. These deliberations are based on physical phenomena such as micro-
crack saturation.
 !
2 D i
c1 · (1 + D12 ) · Σi · exp −c2 √ +


2 )




 Σi · (1 + D12 for σi ≥ 0

c3 · Di · Σ2i · [1 + exp (c5 (Σi − c4 ))]



dDi  " !#1+2·exp(−D12 )
= Di (6.3)
dN 
 2
c1 · (1 + D12 ) · Σi · exp −c2 √ +

 Σi · (1 + D122 )


 for σi < 0
c5

   
c3 · Di · Σ2i · 1 + exp


 (Σi − c4 )
3
6 Theories behind selected fatigue models in FEA-packages 31

In the case of shear damage, which is depicted in equation 6.4, the propagation and fracture terms
were removed due to experimental observations [43, p. 333ff]:

dD12 D12
 
= c1 · Σ12 · exp −c2 √ (6.4)
dN 2 Σ12
Permanent strains in principal directions are accounted for by equation 6.5, in which they are
dependent by the shear damage evolution function, scaled by a constant c6 :
(
dD12
dp,i c6 · i · dN for σi ≥ 0
= (6.5)
dN 0 for σi < 0
For efficient calculation in FEA, a damage dependent algorithm for jumping several cycles is in-
cluded. To obtain the most suitable number of cycles to jump Njump in an element, a relative
damage is calculated, dependent on the damage in the corresponding cycle, as depicted in equa-
tion 6.6, and linearly extrapolated to an appropriate number of cycles, shown in equation 6.7. These
Njump values are classified in a cumulative relative frequency distribution and a certain percentile,
e.g. 10%, is taken as global Njump for further analysis.

−9 for Dij = 0
10


∆Dij = 0.5 · Dij for 0 < Dij ≤ 0.2 (6.6)


0.1 for Dij > 0.2
∆Dij
Njump = dDij
(6.7)
dN |N
For UD layers, Carrella-Payan et al. [41] extended the model by using slightly modified versions of
equation 6.3 for 1- and 2-direction as well as for shear loading, but with different sets of the constants
c1 to c5 each, resulting in 15 constants. In addition, the tensional and compressive damage is split

to d+
i and di , which was also done in the last model in [43] already. Therefore, the fatigue failure
indices, which are the ratio between the effective stress and the ultimate strength in this case, are

also divided into Σ+ i and Σi . Permanent strains are covered for in-plane shear and dependent on
the shear damages. Initial degradation due to the first static loading is also considered by starting
the fatigue simulation with the initial static damage obtained from a nonlinear static pre-simulation.
The cycle jump algorithm is generalized to a damage jump to enable treatment of variable amplitude
loadings. This is done by a damage operator approach based on Brokate [44], which is similar to
Rainflow counting but able to calculate damage at defined time increments.

6.1.2 Identification of the material parameters needed


Generally, the individual parameters are obtained by FE-based optimization. This means, that the
constants c1 to c5 , which are considered intrinsic material constants, are obtained by optimizing
them in a non-linear FE simulation with cycle jump algorithm to match a displacement controlled
one-sided bending experiment. Therefore, the out-of-plane displacement profile and reaction force
is recorded and compared. Since each constant is responsible for a certain region in the stiffness
degradation curve obtained from the experiment, the constants can be found step-by-step [43, p.235
ff.]. Furthermore the constant for permanent strain in woven fabrics, c6 , is determined by optimizing
to the experimental plastic deformation obtained from the [#45°]8 [46].
In the case of UD layers, tensile tests on 5 different layups and on 5 different load levels or one-sided
3-point bending tests with 5 layups at only one load level are needed for parameter identification
[41]. In addition, the constant for permanent strain accumulation is estimated by crosschecking the
raw data of the unloading of some of the specimens.
6 Theories behind selected fatigue models in FEA-packages 32

6.1.3 Current status, known advantages and drawbacks of the model


The current definition provides an intralaminar progressive damage model, which is based on residual
stiffness degradation. It covers stress ratio, multiaxiality and permanent strains for woven fabrics.
Furthermore in the newest definition, the model is extended to UD layers and variable amplitude
loading. However, frequency and environmental effects as well as statistical analysis are still missing.
In addition, interlaminar damage is still neglected at the moment. In their latest conference paper
[42], a cohesive approach with a bilinear cohesive law, as suggested by Turon, is considered for future
work. Parameter identification is done by a straightforward procedure, needing a moderate amount
of tensile and bending fatigue tests. In addition it is mentioned that the model is already validated
for several layups, in which good agreement is obtained but further improvement is suggested.
Eventually, the procedure was patented in 2014 by Bruyneel et al. [47].

Discussion

Despite the physical considerations in the development of this theory, the model seems to be em-
pirical since discrete physical phenomena are not modeled. However, the stiffness based damage
calculation shows to be promising for endless fiber reinforced polymers since the overall behavior of
the rather complicated fatigue mechanisms in the matrix can be treated in an easy, descriptive way.
For the reason that the theory was not implemented in Samcef V17.1 (September 2016), it was not
tested in this work.
6 Theories behind selected fatigue models in FEA-packages 33

6.2 3DS Simulia Abaqus: interlaminar fatigue damage modeling using VCCT
low cycle fatigue analysis
In this section, the general assumptions and techniques used in the Abaqus low cycle fatigue analysis
are described. Furthermore it contains a brief description of the identification of the parameters
needed. Additionally the corresponding keywords and solver parameters which are used in the
Abaqus input file are specified.

In general, this technique focuses on interlaminar fatigue damage modeling of laminated composite
structures. Thereby fatigue crack onset and growth can be simulated over the loading cycles due
to the fact that it is a cycle-by-cycle based approach. As a result of the VCCT theory used (see
section 3.3.2), a predefined crack is required, which implies that crack initiation cannot be simulated.

6.2.1 General modeling approach


The present analysis technique starts with extracting appropriate energy release rates at the crack
tip over a single loading cycle. With the help of fatigue criteria, the speed of crack propagation is
estimated node-by-node at the crack front then. By taking the crack tip element lengths in propaga-
tion direction into account, the number of cycles needed for crack propagation can be extrapolated.
This automatically implies a cycle jump approach, which is important for computation efficiency in
cycle-by-cycle based analyses (see section 5.2)

For the first step which is the identification of the relevant strain energy release rates at the crack
tip, the Abaqus Direct Cyclic approach is used. After that the minimum cycle jump where at least
one node pair of the VCCT interaction debonds is calculated with the low cycle fatigue fracture
criterion. Thereby the VCCT interaction represents the modeled interface between the plies. These
steps are described in the following.

Abaqus Direct Cyclic approach

With the Abaqus Direct Cyclic approach, the stabilized state response regarding stresses and strains
can be calculated directly for periodically loaded structures.
Therefore the basic approach is to search for an appropriate displacement function in the form of a
Fourier series, depicted in equation 6.8, where k indicates the amount of Fourier terms, ω expresses
the circular frequency and ui are the unknown coefficients.
n
X
u(t) = u0 + [usk sin(kωt) + uck cos(kωt)] (6.8)
k=1

First, the residual vector in Fourier series representation R(t), shown in equation 6.9, is calcu-
lated. Therefore the coefficients needed are obtained incrementally on an element-by-element basis,
depicted in equations 6.10.
n
X
R(t) = R0 + [Rks sin(kωt) + Rkc cos(kωt)] (6.9)
k=1

Z T Z T Z T
1 2 2
R0 = R(t)dt Rks = R(t) sin(kωt)dt Rkc = R(t) cos(kωt)dt (6.10)
T 0 T 0 T 0

Second the corrections of the displacement coefficients ck are calculated using equation 6.11. Thereby
i shows the current number of iteration and K indicates the elastic stiffness matrix, which only needs
6 Theories behind selected fatigue models in FEA-packages 34

to be calculated once in the procedure. Hence, this leads to reduced calculation time, especially in
large problems.
(i)
Kci+1
k = Rk (6.11)
Last, the displacement coefficients are modified by the corrections, shown in equations 6.12.

(i+1) (i) (i+1) c(i+1) c(i) c(i+1) c(i+1) c(i) c(i+1)


u0 = u0 + c0 uk = uk + ck uk = uk + ck (6.12)

This process is repeated until convergence is obtained, which means that all Rk and ck are suffi-
ciently small. As a result of the linear treatment of the stiffness which can be seen in equation 6.11,
geometric nonlinearities and such that are resulting from interfaces are not taken into account. A
more detailed description can be found in the Abaqus Theory Guide chapter 2.2.3 [18].

The resulting stabilized solution is needed then for estimating the occuring strain energy release
rates at the crack front of the interface via VCCT (see section 3.3.2). However, the manual does
not describe how the strain energy release rates are extracted for a loading cycle.

Low cycle fatigue fracture criterion

Equally to the traditional treatment of the fatigue damage evolution (see section 4.4), the Abaqus
low cycle fatigue criterion is divided into crack onset, which is ∆G-N-curve based, and crack propaga-
tion, which is based on Paris law (compare section 4.4.3). Figure 6.2 depicts the fatigue crack growth
regimes with the linear Paris regime in the middle, between Gthresh and Gpl . Under Gthresh , no crack
onset or growth will occur and between Gpl and GC , crack growth occurs at dramatically increased
velocity. In Abaqus, Gthresh and Gpl are set in relation to GC in Abaqus which is calculated by the
corresponding mixed mode criterion used, which is BK for example (see section 2.4). Thereby in
equations 6.13, the default values are shown.

Figure 6.2: Fatigue crack growth regimes [18]


6 Theories behind selected fatigue models in FEA-packages 35

Gthresh Gpl
r1 = = 0.01 r2 = = 0.85 (6.13)
GC GC
In the low cycle fatigue criterion, a modification of the onset criterion described in section 4.4.2 is
implemented, which can be seen in equation 6.14. Thereby c1 and c2 are material constants and N
is the cycle number. Other than in the original equation, ∆G is used instead of G1,max .

N
f= ≥1 (6.14)
c1 ∆Gc2
If GGTTC from the BK criterion is in between r1 and r2 and the onset criterion is fulfilled, the
propagation according to Paris law is calculated. In Abaqus, this is driven by ∆G and controlled
by the material constants c3 and c4 , shown in equation 6.15.

da
= c3 ∆Gc4 (6.15)
dN
The existent modification of the original stress intensity factor based Paris law (see section 4.4.3) is
valid for the reason of the polynomial connection between ∆K and ∆G, derived in equations 6.16:

σmin Kmin
R= ∼
σmax Kmax
Kmin = R · Kmax
∆K = Kmax · (1 − R)
1 2 2
∆G = Gmax − Gmin = · (Kmax − Kmin )
E0
1+R
∆G = · ∆K 2 (6.16)
(1 − R) · E 0

Since the element lengths are known, the minimum number of cycles needed for the release of at
least one node is calculated. If GGTTC > r2 , the node will be released in the next cycle already. This
procedure supports linear elements only. More information about the Abaqus low cycle fatigue
criterion can be found in the Abaqus Analysis User’s Guide chapter 11.4.3 - Low-cycle fatigue
criterion [18].

6.2.2 Identification of the material parameters needed


For the procedure described above, several material parameters are needed. Therefore their identi-
fication via specific material tests is described in the following.
The lamina stiffness parameters such as Young’s moduli in each direction, shear moduli and Poisson’s
ratio are obtained by several standard tests which can be found in literature.
The parameters for the low cycle fatigue fracture criterion are obtained by monotonic and cyclic
specimen tests and curve fitting. In the present case, the BK criterion is used for the static part.
Therefore static fracture toughness experiments have to be performed. As mentioned in section 2.4
already, these include tests with double cantilever beam (DCB) specimens for G1C , end notch
flexure (ENF) specimens for G2C and mixed mode bending (MMB) specimens with several mixed
mode ratios (MMR, see section 2.4) for mixed mode fracture toughnesses. A common technique
for measuring the fracture toughness there is the compliance calibration method which is described
in [10, p.5-7]. The parameter η is obtained by a curve fit when plotting the mixed mode fracture
toughnesses over the mixed mode ratio.
6 Theories behind selected fatigue models in FEA-packages 36

For the constants c1 to c4 which are needed for crack initiation and propagation, fracture mechanics
fatigue tests have to be performed for the corresponding mode and stress ratio. Thereby c1 and
c2 are obtained by fitting equation 6.14 to ∆G = Gmax − Gmin over N , whereas N is indicating
the cycle number when the initial compliance rises by 1% which indicates the start of stable crack
propagation [30, p.7]. The constants c3 and c4 which are responsible for stable crack propagation
da
according to Paris law are achieved by fitting equation 6.15 to dN over ∆G.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 37

7 Simulations and validations of Abaqus VCCT low cycle fatigue


method using 3D elements
In this section, the capabilities of Abaqus version 6.14 regarding delamination are tested. The work
is focused on the fracture-mechanics based VCCT method which is described in section 3.3.2 in
combination with the Abaqus Direct Cyclic solve and the low cycle fatigue approach.

7.1 Development of a suitable reference case


In the following section, reference case simulations for the assessment of fatigue delamination were
developed.

General assumptions and workflow:

ˆ For the reason of applicability to realistic 3D structures and the option of modeling multiple
interfaces via stacked solid laminates, 3D element discretization was chosen. Therefore models
with solid and continuum shell elements were built up.

ˆ Due to the need of trustworthy experimental and material data for validation and comparison,
the basis simulations were anchored at simulations from R. Krueger. He performed mode 1
and mode 2 VCCT fatigue simulations which were validated by experiments, [48] and [49].
Thereby the focus lay on mode 2 for the reason of better data.

ˆ Because of stable crack propagation over a longer period of cycles, a displacement controlled
fatigue loading was chosen. Therefore, Krueger’s synthetic crack propagation curves were used
for comparison [48, 49]. These were derived semi-analytically (compare [50]) for the displace-
ment controlled fatigue crack propagation simulations and are based on 2D assumptions and
stress controlled experiments.

ˆ By testing and comparing simulations with different parameters and solver options, suitable
reference cases for further investigation were chosen.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 38

7.1.1 End notch flexure (ENF) FE model


The end notch flexure specimen shown in figure 7.1 is common for interlaminar crack propagation
in pure mode 2. Therefore this experiment was re-modeled in Abaqus. Test data and specimen
dimensions were taken from [30] and [49]. In this case, the specimen consisted of 24 plies 0° IM7/8552
UD carbon-epoxy.

~u = f (t)
2h

e
a0
L 2L z
B
y
x
Ltot

Figure 7.1: ENF test configuration with a pre-cracked specimen

parameter value/inch value/mm


Ltot 7 177.8
L 2 50.8
B 1 25.4
e ∼1 ∼25.4
a0 1 25.4
h 0.089 2.25

Table 7.1: dimensions of the ENF specimen

In table 7.1 the dimensions of the specimen are given. The crack is produced by inserting a 76.2 mm
(3 inch) long teflon film of 0.01 mm thickness, followed by pre-cracking in a static test. Therefore
the dimension e varies slightly. For the simulations Krueger set it to 25.4 mm (1 inch) in [49]. The
vectors ~u(t) indicate the applied cyclic displacement loading.

Model description

As already mentioned in section 7.1, the base model is very similar to mode 2 simulations which
were performed by R. Krueger in 2011. However, some simplifications regarding the geometry were
added for the reason of reduced calculation time. Figure 7.2 depicts the model used for the finite
element discretization. It can be clearly seen that the sections after the supports are cut off since
they are completely unloaded and hence do not affect the crack growth.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 39

bonded face
support B
~u(t)
2h

support A z
y
a0 x
B L 2L

Figure 7.2: simplified ENF model

Material properties

The material properties of the lamina were taken from [49] and can be seen in table 7.2. Additionally
the interface fracture parameters were taken from [30]. These are shown in table 7.3 and divided
into static fracture properties using a BK law on the left, delamination growth onset coefficients in
the middle and Paris law coefficients for delamination propagation on the right.

E1 E2 = E3 ν12 = ν13 ν23 G12 = G13 G23


161GPa 11.38GPa 0.32 0.45 5.2GPa 3.9GPa

Table 7.2: material properties for Hexcel ®IM7/8552 graphite epoxy


G1C G2C = G3C η c1 c2 Gth c n
0.21J/mm² 0.78J/mm² 2.5713 0.213 -6.329 0.08J/mm² 0.33 5.55

Table 7.3: fracture parameters for Hexcel ®IM7/8552 graphite epoxy 0°/0° interface
Simulation steps

The simulations were performed with Abaqus/Standard version 6.14-1 which is an implicit solver
algorithm. At the beginning of each simulation, a static verification step was performed with the
step type "Static, General". After that, the fatigue simulation was started using a "Direct Cyclic"
step. Table 7.4 shows the parameters in the Direct Cyclic step which were used in general. Due to
the chosen increment size of 1ms at a loading frequency of 5Hz a single loading cycle was divided
into 200 increments each. The maximum number of Fourier terms of 50 was used based on the
recommendation Krueger made in [49] after variation of this parameter.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 40

variable value
Maximum number of increments 1 · 106
Increment size 1 · 10−3 s
Maximum number of iterations 1 · 103
Number of Fourier terms 50
Maximum number of cycles 1 · 108

Table 7.4: parameters used in the Direct Cyclic step

For achieving convergence in the Direct Cyclic step (see paragraph 7.1.4), the control parameters of
the Direct Cyclic step had to be adjusted as suggested in [49]. Thereby, the ratio of the maximum
residual coefficients to the time averaged force CRnα was set to 100, which means that the criterion
was more or less deactivated. The remaining criterion, the ratio of the maximum correction to
the displacement coefficient to the largest displacement coefficient CUnα , was refined to 1 · 10−3 .
These coefficients were taken for the plastic ratcheting detection criteria CR0α and CU0α , too. In
Appendix D, modified control parameters and variable Fourier terms were tried. There, the adaptive
Fourier term algorithm of the Direct Cyclic approach was tested. The same results were obtained,
but since the calculation time did not decrease, it was not used in further simulations.

VCCT interaction

The interface was modeled using a node-to-surface contact with VCCT interaction property using
BK law for mode mixture, a SN-curve approach for crack onset and Paris law for delamination
propagation with the parameters depicted in table 7.3. Thereby the initially bonded nodes were
defined by a node set, which contained the corresponding nodes from both contact surfaces, as
depicted in figure 7.2 and a VCCT crack was defined on the contact. Furthermore, node adjustment
was used for precise initial node match.

Boundary conditions and displacement loadings

At support B (blue dotted line), translation was locked in all direction over the whole edge. Fur-
thermore at support A (blue continuous line), translation was locked in y- and z-direction over the
whole edge, too.

The vector ~u(t) indicates the periodic time-dependent displacement loading. In the static step, a
displacement of 0.001mm was applied in negative z-direction. In the fatigue step, periodic sine
displacement loadings with a frequency of 5Hz were applied in negative z-direction. Equation 7.1
shows the Abaqus definition of an amplitude a:

a(t) = A0 + A1 cos(ω(t − t0 )) + B1 sin(ω(t − t0 )) (7.1)


In the case of a stress ratio of R = 0.1 and considering no phase-shift t0 , the coefficients, which are
shown in table 7.5, are obtained. This leads to the amplitude depicted in figure 7.3.

ω/s−1 t0 /s A0 A1 B1
31.42 0 0.55 0 0.45

Table 7.5: loading amplitude coefficients for R=0.1 without phase-shift t0


7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 41

amplitude a(t)
0.5

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
time t / s

Figure 7.3: periodic sine loading for R=0.1 without phase-shift t0

Eventually the displacement vector ~u(t) is attained by multiplying the amplitude with a displace-
ment scale vector d~ (1 mm in negative z-direction in the present ENF model), depicted in equa-
tion 7.2.

 
0
~u(t) = d~ · a(t) ~
d =  0  mm (7.2)
 

−1

Element types

Generally, linear elements had to be chosen, because the Abaqus low cycle fatigue analysis, described
in section 6.2.1, does not support higher order elements, as mentioned in the Abaqus Analysis User’s
Guide chapter 11.4.3 [18].
In the solid models, linear hexahedral elements with incompatible mode formulation (in Abaqus:
C3D8I) were used due to their better bending behavior. These are described in [18] in the Abaqus
Analysis User’s Guide chapter 28.1.1 and in the Abaqus Theory Guide chapter 3.2.5 in detail.
In the continuum shell models, linear hexahedral continuum shell elements (in Abaqus: SC8R) were
used. As a result of their shell-like behavior, the elements require an orientation assignment.

Node sets for evaluation

For easier evaluation of the crack length, specific node sets were created. These cover a node set at
each edge and one in the center of the master surface, which is located at the lower stack. Figure 7.4
shows the locations of these node sets with the names edge 1, center and edge 2 schematically. As
it can be seen, they start at the initial crack and proceed until the end of the specimen. For
the evaluation of the current crack length, the bond state, which is 1 when bonded and 0 when
released, is summed up, subtracted from the initial number of bonded nodes and multiplied by
the element length. This procedure does not cover phenomena in which some bonds remain in
between the cracked area. However, these effects would indicate calculation problems anyway since
the specimen does not have any interlaminar crack arrestors and hence the crack propagation is
straight forward.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 42

edge 1

center

edge 2

initial crack

Figure 7.4: node sets for evaluation

7.1.2 Description of the relevant Abaqus keywords


In the following section, relevant Abaqus keywords used for interlaminar fatigue simulations using
VCCT on 3D elements are described. Thereby, letters in < > are variables, which have to be defined
in the input file or have to be placed directly on the corresponding position and letters in > < are
names of node or surface sets. The full description can be found in [18] Abaqus Keywords Reference
Guide under the corresponding keywords.

Amplitude definition

For the sine displacement loading used in the fatigue simulations, the corresponding amplitude
curve, in the following example with the name "amp1", has to be defined:

*Amplitude, name=amp1, definition=periodic


<NFourier terms >, <ω>, <t0 >, <A0 >
<A1 >, <B1 >

As it can be seen, the keyword "definition" is set to "periodic" for the definition of sine and cosine
loadings. In the second line, the number of Fourier terms, the circular frequency, the phase shift and
the constant term have to be specified. Finally additional lines consist of the Fourier coefficients
An and Bn for cosine and sine each in the same style of the second line.

Contact definition

For implementing a contact with VCCT behavior, an interaction property with an arbitrary name,
here "IntProp", must be defined:

*Surface Interaction, name=IntProp


<OUT-OF-PLANE-THICKNESS>,

Since frictional behavior is neglected, no optional parameters are needed. The out-of-plane thickness
is needed for 2D models and hence can be omitted when using 3D models. The contact is built up
in the following way:
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 43

*Contact Pair, interaction=IntProp, adjust=>NAME-BN<


>NAME-MS<, >NAME-SS<

Thereby the master- and slave contact surfaces (NAME-MS, NAME-SS) have to be given. By
default, a node-to-surface contact using finite sliding is defined. In addition, the parameter "adjust"
is optional and adjusts the initial positions of the opposing nodes in the defined node set of the
bonded nodes (NAME-BN) for perfect match at the beginning of the analysis. Since VCCT requires
an initial crack which is represented in the model by an unbounded contact area, the initial condition
of the contact has to be defined:

*Initial Conditions, type=contact


>NAME-MS<, >NAME-SS<, >NAME-BN<

Direct cyclic step

The Direct Cyclic approach for low cycle fatigue analysis is implemented with the following keyword
and the variables described in table 7.6:

*Direct Cyclic, fatigue


<tinit >, <T>, <tinit >, <tmax >, <NF,init >, <NF,max >, <NF,inc >, <Nit, max >
<∆Nmin >, <∆Nmax >, <Ntot >, <∆ex >

tinit initial time increment


T time of a single loading cycle
tinit minimum allowed time increment
tmax maximum allowed time increment
NF,init initial number of Fourier terms
NF,max maximum number of Fourier terms
NF,inc inrement number of Fourier terms
Nit, max maximum number of iterations allowed
∆Nmin minimum number of extrapolated cycles
∆Nmax maximum number of extrapolated cycles
Ntot total number of cycles
∆ex damage extrapolation tolerance (default: 1)

Table 7.6: variables in the Direct Cyclic step

Crack location and fracture criterion

The location of the crack and the fracture criterion are defined in the individual steps. Therefore
the keyword *Debond defines, that crack propagation may occur between two partially bonded
surfaces, starting at the unconnected regions of the contact:

*Debond, slave=>NAME_SS<, master=>NAME_MS<

When adding the optional keyword DEBONDING FORCE=RAMP in the next line, the bond force
is not released immediately but gradually over the next iterations instead, when debonding of a
node occurs.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 44

Right after the debond keyword, the corresponding fracture criterion has to be defined. In the
Direct Cyclic step, the low cycle fatigue criterion is implemented in the following way:

*Fracture Criterion, type=fatigue, mixed mode behavior=BK, tolerance=<reltol>, viscosity=<visc>


<c1 >, <c2 >, <c3 >, <c4 >, <r1 >, <r2 >, <G1C >, <G2C >,
<G3C >, <η>

As it can be seen, BK mixed mode behavior is chosen (compare section 2.4). The optional keywords
"tolerance" (default: 0.2) and "viscosity" (default: 0) can be implemented for defining another release
tolerance or viscosity than the default values (see section 7.1.4). In the second and third line, the
Gpl
constants for crack initiation, propagation, r1 = GthreshGC , r2 = GC and the BK parameters are
defined. For the static step, a simple VCCT criterion with BK mode mixture is used. Therefore,
the same initial keyword with type=vcct and without the fatigue parameters has to be written.

Direct cyclic controls

To change the default values of the criteria for stabilized state detection, the following keyword can
be used:

*Controls, type=Direct Cyclic


<IP I >, <CRnα >, <CUnα >, <CR0α >, <CU0α >

Thereby IP I indicates the iteration number at which the periodicity condition is first imposed (de-
fault: 1). The following parameter <CRnα > defines the maximum allowable ratio of the largest
residual coefficient on any terms in the Fourier series to the corresponding average flux norm,
whereas <CUnα > describes the maximum allowable ratio of the largest correction to the displace-
ment coefficient on any terms in the Fourier series to the largest displacement coefficient (also see
section 6.2.1). The last two entrances describe in principal the same ratios but for the constant
terms of the Fourier series. Hence they become important when plastic ratcheting occurs. Per
default, these ratios are set to 5 · 10−3 .
Additional information can be found in the Abaqus Analysis User’s Guide chapter 7.2.2 - "Con-
trolling the solution accuracy in Direct Cyclic analysis" [18].

Contact stabilization

Contact stabilization is defined in each step separately. For including damping of one specific contact
pair, the following implementation has to be used:

*Contact Controls, stabilize, master=>NAME-MS<, slave=>NAME-SS<

Stabilization is applied to any contact pairs if the definition of master and slave surface is omitted.
The damping coefficient can be set manually or automatically calculated by Abaqus, depending on
the implementation:

ˆ If no dataline is included, automatic stabilization is used. The damping can be scaled by


adding a scaling factor after the keyword stabilize, e.g. "stabilize=2".
ˆ For setting of the damping coefficient manually, a data line with the damping coefficient, the
fraction of damping that remains at the end of the step (default: 0) and the clearance at
which damping becomes zero has to be added.
ˆ By using "stabilize=user adaptive", damping factors can be set for each iteration of an incre-
ment.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 45

7.1.3 Example input file


An example input file for an interlaminar fatigue simulation using 3D elements, VCCT and the
Abaqus Direct Cyclic approach can be found in Appendix F.

7.1.4 Parameters and options used for increasing accuracy and convergence
In this section, specific parameters which were used to improve convergence and accuracy are de-
scribed.

Direct cyclic control parameters

Since the Abaqus Direct Cyclic Approach is an iterative solution technique, infinite loops can
occur due to unreachable convergence criteria. Therefore these have to be adjusted in some cases.
However, one should take care when changing these parameters since they have a strong influence on
the solution, since they control the size of the allowed residuals for convergence. The corresponding
implementation is described in section 7.1.2.

Release tolerance

The release tolerance induces a cut-back operation of the time increments if the fracture criterion
f is over-fullfilled by the release tolerance in the case of the VCCT or low cycle fatigue criterion.
It can be defined in the corresponding fracture criterion (see section 7.1.2). When using fixed time
increments, it does not affect the simulation.

Viscous regularization

Viscous regularization applies to nodes which have just debonded in a VCCT contact. It causes the
tangent stiffness matrix of the softening material to be positive for sufficiently small time increments.
Since too high values of viscous regularization can distort the solution, the damping energy, which
can be found in the history output request under ALLVD, should be compared to the elastic strain
energy ALLSE. As well as the release tolerance, the viscous regularization is defined in the fracture
criterion (see section 7.1.2).

Contact stabilization

Contact stabilization applies tangential and normal damping to contact pairs which are within a
specific opening distance. As well as viscous regularization, it influences the solution significantly
if set too high. Therefore the corresponding dissipation energy ALLSD, which can be found in the
history output, should be compared to the elastic strain energy ALLSE. A more detailed description
can be found in the Abaqus Analysis User’s Guide section 36.3.6 - Adjusting contact controls in
Abaqus/Standard [18].
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 46

7.1.5 Simulations and determination of the reference case


The following section gives a summarized overview of the development of appropriate reference cases
for further investigation of the numerical fatigue model in Abaqus.

Selected Simulations and results

For the first simulations, simple meshes were used, where the element dimensions are homogeneously
divided into a certain number of elements. These meshes are named mesh A in the following. The
individual number of elements and the resulting element sizes can be found in table 7.7. Figure 7.5
depicts mesh A for solid elements (the corresponding figure for continuum shell elements can be
found in Appendix E). Thereby the red section indicates the bonded area of the contact while the
blue section shows the released node area, which follows a contact definition.

number of elements resulting element size


length 2L 102 0.996 mm
width B 25 1.016 mm
height H (solid elements) 3 0.75 mm
height H (continuum shell elements) 1 2.25 mm

Table 7.7: element data in the mesh A models

Figure 7.5: mesh A for solid elements

The first simulations were done with continuum shell elements for the reason of their computational
efficiency compared to the solid element model, since former already needed about 30 hours of
calculation time for 1 · 108 cycles on the PC (3rd generation intel core i5 processor with 8GB RAM
and a 256GB solid state drive).
In the very first test case using mesh A for continuum shell elements, a release tolerance of 1 · 10−3
and no viscosity or contact stabilization were used. However, since an increment size of 0.001 s is
used in the Direct Cyclic step, which results in 200 increments per cycle, the release tolerance does
not affect the results.

Figure 7.6 shows the increase in crack length a over the cycles N at the edges and in the center of
the specimen. As it can be seen, the crack propagation is very inhomogeneous over the crack front,
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 47

especially at the edges.

50

40
crack length a / mm

30

edge 1
20 center
edge 2
benchmark curve
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.6: crack progress curve of mesh A and continuum shell elements

Figure 7.7 shows the final bond state of mesh A using continuum shell elements. It can be seen, that
some nodes remain bonded (red) in between the area of released nodes (blue). Further investigation
showed calculation errors of the strain energy release rates in mode 2 and mode 3 ("dead" nodes),
which are needed for debonding in the ENF model. These are caused by the crack onset criterion
under certain loadings due to a questionable implementation (see section 7.2.4).

Figure 7.7: final bond state of mesh A and continuum shell elements

Other simulations of the continuum shell model using viscous regularization in the range from 1·10−4
to 1 · 10−6 or deactivating the initial node adjustment (see section 7.1.1) did not affect the results
significantly.

Using mesh A and solid elements resulted in much smoother crack propagation as it can be seen
in figure 7.8, where neither contact stabilization nor viscous regularization were used. On edge 1,
crack propagation was still not perfectly smooth. Viscous regularization in the range from 1 · 10−4
to 1 · 10−5 did not impact the results.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 48

50

40
crack length a /

30

edge 1
20 center
edge 2
benchmark curve
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.8: crack progress curve for mesh A and solid elements

Since the area in front of the crack and at the end of the specimen are less important in the calcula-
tion, these sections were roughened in an enhanced mesh B. Figure 7.9 depicts the partitioning of
a stack. It can be seen that partition A with the length a = 25.4 mm reaches to the crack tip a0 . In
partition B, the mesh size was kept small and over the remaining length c = 11.2 mm of partition
C, the mesh was coarsened again. The element sizes can be found in table 7.8.
z
a = a0 c

A B C h x
2L y

Figure 7.9: partitioning of mesh B

number of elements resulting element size


partition A 17 ∼1.5 mm
partition B 65 ∼1 mm
partition C 7 1.6 mm
width B 24 ∼1.1 mm
height h in the solid model 3 0.75 mm
height h in the continuum shell model 1 2.25 mm

Table 7.8: element data in the mesh B models

Figure 7.10 shows mesh B for solid elements. Thereby the red area indicates the bonded nodes of
the VCCT contact and the blue area the area of released nodes.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 49

Figure 7.10: mesh B for solid elements

Simulations using mesh B showed very smooth crack propagation on all edges, as depicted in
figure 7.11:

70

60
crack length a / mm

50

40

30 edge 1
center
20 edge 2
benchmark curve
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.11: crack progress curve for mesh B and solid elements

However, the crack progress in the simulation deviated from the synthetic benchmark curve with
increasing crack length.
Using automatic contact stabilization lead to almost perfect overlap with the synthetic benchmark
curve, which is shown in figure 7.12:
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 50

50

40
crack length a / mm

30

edge 1
20 center
edge 2
benchmark curve
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.12: crack progress curve for mesh B and solid elements with automatic contact stabilization

Nevertheless, the solution is delusive, since the automatic contact stabilization damps the main cause
of the derivation: unexpected high strain energy release rates in mode 1, which occur after passing
the line of load introduction at a crack length of a=25.4 mm. This can be observed in figure 7.13,
where the relative mode 1 strain energy release rates per cycle are plotted over the center crack
length a for the mode 2 simulation using mesh B. The blue dots indicate the simulation with
automatic stabilization whereas the red dots show the simulation without any stabilization. In
mode 2 and 3, the relative strain energy release rates per cycle behave as expected, as it can be
seen in figures 7.14. There, the rather high values in mode 3 result from the 3D crack front and
edge effects.

0.1
∆G1,max / 103 mJ2

0.08

automatic stabilization
0.06 unstabilized

0.04

0.02
0 5 10 15 20 25 30 35 40 45 50
crack length a / mm

Figure 7.13: comparison of ∆G1,max over the center crack length


7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 51

0.2
0.6
∆G2,max / 103 mJ2

∆G3,max / 103 mJ2


0.15
0.4
automatic
0.1 stabilization
0.2
unstabilized

0 0.05
0 20 40 0 20 40
crack length a / mm crack length a / mm

Figure 7.14: comparison of strain energy release rates over the center crack length,
left: ∆G2,max ; right: ∆G3,max

To quantify the amount of damping applied, the increase of static dissipation energy, Abaqus History
Output request ALLSD, was calculated for every loading cycle and compared with the maximum
internal energy, ALLIE, in the corresponding loading cycle, as suggested in the Abaqus manual
[18]. Since the present case is linear elastic and without electrostatic, plastic or viscous dissipation,
the maximum internal energy corresponds to the maximum strain energy, ALLSE. In the case of
automatic stabilization, the values of ALLSD were in the same order of magnitude as ALLSE after
the crack passing the line of load introduction, which is too high since it should be under about 5%
compared to ALLIE/ALLSE. The impact of automatic contact stabilization was also investigated
on a double cantilever beam (DCB) model Appendix B.

Interestingly, when performing a simulation with the crack already beyond the line of load introduc-
tion, the phenomenon does not occur, which means that the values of ∆G1,max /G1C remain under
10−2 . This was tested by using mesh B with an initial crack front shifted to a=26 mm, when the
line of load introduction is at a=25.4 mm. In addition, the crack onset criterion was deactivated,
since in general, it does not influence the crack propagation curve with these input parameters, but
occasionally causes significant bugs in simulations with 3D crack fronts (see section 7.2.4).
For better comparison with the 2D benchmark case, the starting point was shifted to the benchmark
curve by adding 1.1 · 104 cycles. Figure 7.15 depicts the comparison between the crack progress
curves at the individual node sets. The little deviation in the curve and final crack length results
from ∆G3 , which is caused by the 3D crack front and neglected in Krueger’s 2D benchmark curve.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 52

50

40
crack length a / mm

30

edge 1
20 center
edge 2
benchmark curve
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.15: crack progress curves for mesh B and shifted initial crack position

In mesh B, the upper stack is slightly distorted due to an imperial sized line of load introduction
at the upper surface and a metric mesh size of 1 mm in the significant region of the VCCT contact
at the lower surface. Therefore, attempts with an undistorted model and imperial mesh sizing were
made, which can be found in Appendix C. Nevertheless, these showed exactly the same behavior.

Due to very high calculation times of more than a week per simulation, roughened derivatives of
mesh B were formed, which will be called mesh B 1.5 mm and mesh B 2 mm in the following.
Tables 7.9 and 7.10 show the number of elements over the section lengths and the resulting element
sizes:
number of elements resulting element size
partition A 13 ∼2 mm
partition B 43 ∼1.5 mm
partition C 6 ∼1.9 mm
width B 18 ∼1.4 mm
height H 2 ∼1.1 mm

Table 7.9: element data in the mesh B 1.5 mm

number of elements resulting element size


partition A 8 ∼3.2 mm
partition B 33 ∼2 mm
partition C 4 2.8 mm
width B 12 ∼2.1 mm
height H 1 2.25 mm

Table 7.10: element data in the mesh B 2 mm

Visualizations of the coarser meshes can be found in Appendix E.


7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 53

In these simulations, automatic contact stabilization was applied again since the actual driving
variables - ∆G2 and partly ∆G3 - remain unaffected and the calculation time slightly decreases.
Figure 7.16 shows crack progress curves of the center nodes in comparison. It can be seen, that
a mesh size of 2 mm leads to overestimation of the crack progress. In the case of a mesh size of
1.5 mm, the curve still overlaps with both the finer mesh and the benchmark curve at significantly
decreased calculation times by one order of magnitude - from about 200 hours to 20 hours.

50

40
crack length a / mm

30

mesh B 1 mm
20 mesh B 1.5 mm
mesh B 2 mm
benchmark curve
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.16: comparison of the center crack progress using different mesh sizes

Choosing the reference cases

For the reason of calculation efficiency, the assessments were performed with mesh B 1.5 mm,
unless otherwise stated. In assessments where only deviations between simulations were important,
automatic stabilization was used for the reason of saving calculation time, since it only affects the
region after a crack length of 25 mm in a specific, reproducible manner.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 54

7.2 Assessment of fatigue delamination under cyclic loading


In the following section, the fatigue solver was analyzed regarding its treatment of the stress ratio
R, mixed-mode loading states, damage accumulation and the impact of the crack onset criterion.
Furthermore, the solver was tested regarding its robustness to other parameters which may influence
the results. These included the effect of CPU-parallelization, phase-shifted and non-sinusoidal
loading amplitudes.

7.2.1 Treatment of the stress ratio


Since both the crack onset and propagation are driven by a single load-dependent parameter only
when considering the load at the crack tip in the Abaqus low cycle fatigue analysis, which is ∆G
(compare section 6.2), a certain stress ratio dependency is implied.
Starting with the quadratic relationship between stresses at the crack tip and strain energy release
rate, which was mentioned in section 2.3.2, the equation of ∆G can be expanded and set in relation
to the stresses multiplied by a scalar constant c, as depicted in equation 7.3.

2 2
∆G = Gmax − Gmin = c · (σmax − σmin ) (7.3)
Bringing in the stress ratio R (see equation 4.2), the stress ratio dependency is described as in
equation 7.4:

σmin
R=
σmax
2
∆G ∼ σmax · (1 − R2 ) (7.4)

To give a better visualization of this behavior including the different regions of crack propagation,
a synthetic diagram, similar to a Haigh diagram, which is used for crack initiation in the general
case, is derived in the following.

The minimum and maximum stresses can be replaced through the mean stress and stress amplitude,
which is shown in equation 7.5. Then, the constant terms for obtaining a curve that describes all
conditions of the same crack progress are placed on one side, resulting in equation 7.6.

∆G = c · ((σm + σa )2 − (σm − σa )2 ) =
2
= c · (σm + 2σm σa + σa2 − (σm
2
− 2σm σa + σa2 ))
= 4c · σm σa (7.5)

∆G
= σm · σa (7.6)
4c
Since the stresses at the crack tip σ and the applied crack opening loading P are linearly dependent
at every length of the crack a when considering small deformations, the equation can be converted
to a load based form, which is done in equation7.7:

∆G
= (Pm · f (a)) · (Pa · f (a)) = Pm · Pa · f (a)2 (7.7)
4cP
Thereby the function f (a) describes the conversion factor of the applied load to the stresses at the
crack tip, dependent on the crack length a.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 55

For the example of a DCB specimen with mode 1 fatigue crack propagation, the strain energy release
rate is calculated as shown in equation 7.8 (see [10] for the derivation via compliance method). There,
b indicates the width of the specimen, h is the height of the individual ply stacks equal to the mode
2 ENF specimen (see figure 7.2) and E the elastic modulus for the corresponding bending direction.
The conversion factor f (a), as well as the constant cP can be obtained as depicted in equations 7.9.

12P 2 a2
G= = (P · a)2 · cP (7.8)
b2 h3 E

s
b2 h3 E
cP = f (a) = a (7.9)
12

The plot searched for is obtained by unifying the load to the dimensionless form by dividing the
mean and amplitude loading through the critical load.
Since crack onset or propagation does not occur when Gmax is smaller than a certain Gthresh , the
rate of crack propagation increases dramatically over a certain Gpl and the static boarder of GC
(compare section 6.2.1), the diagram can be divided into multiple regions. Considering the fatigue
constants from [48] for the mode 1 DCB model, which are listed in table B.3 and normalizing to
the static fracture load PC , the following figure 7.17 shows the individual regions of crack growth
and curves for selected values of Gthresh
GC , depicting the same speed of crack progress:

R=-1
1

R=0
0.8
Pa / PC (a)

N
0.6
∆G da
GC
↑ onset ↓, dN

∆G
GC = 0.25
∆G Gthresh
0.4 GC = GC = 0.353
∆G
GC = 0.6
∆G
GC = 0.85
0.2

R=1
0
0.2 0.4 0.6 0.8 1
Pm / PC (a)

Figure 7.17: regions of crack propagation with curves of the same speed of crack progress

There the green area indicates the region under Gthresh , where no crack onset or propagation occurs.
The yellow region indicates the Paris regime, where stable crack propagation occurs after a certain
number of cycles for crack onset. In the orange region, crack propagation is dramatically increased,
hence Paris law is not valid anymore in this region. In Abaqus, it means that the bond at the crack
tip will be released and the cycle count is increased by 1. At higher values of Gmax , completely
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 56

unstable crack propagation occurs for the reason of exceeding the static GC . The hyperbolic lines
indicate the conditions of same crack progress. The higher the values of ∆G
GC , the lower is the num-
da
ber of cycles needed for crack onset Nonset and the higher is the speed of crack propagation dN .
However, these are only valid in the yellow region. Since in mode 1, compressive loadings do not
affect the crack, the left side of the diagram is not drawn. In mode 2 and mode 3, the graph is
mirrored along the vertical axis for negative values of the mean stress σm , since interlaminar shear
stresses affect the crack in both directions.

To validate this theory, synthetic benchmark cases with different stress ratios and maximum stresses,
but equal ratios of the applied relative energy release rates with respect to the critical energy release
rate at a certain crack length ∆G
GC were created. These cases should show the same evolution of the
crack length a over the cycles N at a given loading. Equation 7.10 describes the connection of ∆G GC
to the minimum and maximum loading applied. Using this relationship, a new pair of minimum
and maximum load level can be determined, shown in equation 7.11.

∆G Gmax − Gmin (P 2 − Pmin2 ) · f (a)


= ∼ max (7.10)
GC GC PC (a) · f (a)

2
Pmax 2
− Pmin P 2 − Pmin2
= max
PC (a) PC (a)
2 2 2 2
Pmax − Pmin = Pmax,new − Pmin,new
q
Pmin,new = 2
Pmax,new 2
− (Pmax 2 )
− Pmin (7.11)

Since the applied loading and the resulting displacement is linearly dependent when considering
small deformations and linear elastic materials, the equation can be also used for the displacements
applied in a displacement-controlled fatigue analysis. The reference case of the ENF simulation
uses a maximum displacement dmax = 1 mm and a stress ratio of R = 0.1. For validation of the
derivations made above, two additional cases were constructed with the dependency shown in equa-
tion 7.11. Therefore, maximum displacements of dmax = 1.1 mm and dmax = 1.2 mm were chosen.
According to equation 7.11 and inserting the displacements instead of the loadings, this results in
stress ratios of R = 0.426 (dmin = 0.469 mm) and R = 0.559 (dmin = 0.671 mm), respectively. These
load cases should result in a similar behavior as the reference case.
Figure 7.18 shows a comparison of the reference case defined in section 7.1.5 with the synthetic
loading cases using mesh B 1.5 mm. It can be seen that in the case of dmax = 1.1 mm, the curve
completely overlaps with the reference case. In the case of dmax = 1.2 mm, the region of accelerated
crack growth is reached at about 2000 cycles. After getting in the Paris regime again, the curve
converges to the reference case. Using mesh B with an element length of 1 mm only, which is
depicted in figure 7.19, the same results, except a higher resolution at the region of accelerated
crack growth, are obtained.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 57

50

40
crack length a / mm

30

R=0.426, dmax =1.1 mm


20 R=0.559, dmax =1.2 mm
reference case R=0.1
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.18: comparison of synthetic equal ∆G cases with the reference R=0.1: mesh B 1.5 mm

50

40
crack length a / mm

30

R=0.426, dmax =1.1 mm


20 R=0.559, dmax =1.2 mm
reference case R=0.1
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.19: comparison of synthetic equal ∆G cases with the reference R=0.1: mesh B

Many researchers already tried to find a proper treatment of the stress ratio in crack propagation.
Thereby most approaches base on Gmax or ∆G only. However, recent papers such as [51] show
dependency of both parameters on the speed of crack growth in the Paris regime.

7.2.2 Treatment of mixed-mode loading conditions


Since Abaqus uses the BK-criterion for the critical strain energy release rate, a certain treatment of
mode mixity is included. However, it only covers the static mode mixity for forced fracture. For the
reason that the threshold and upper Paris limit of the strain energy release rate are scaled by GC ,
the distribution of the regions as shown in figure 7.17 remains the same. Nevertheless, in Abaqus,
the Paris coefficients remain constant over all mixed mode ratios, which is not the case in reality. In
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 58

[52], the developers of the BK-criterion suggest the following modification of the Paris law criterion,
depicted in equations 7.12 to 7.14, which varies the constants depending on the mixed mode ratio
(MMR, see equations 2.6) in a similar way as in the BK-criterion:

da n(MMR)
= c(MMR) · ∆GT (7.12)
dN
mc
c(MMR) = e(ln c2 −(ln c1 −ln c2 )·(1−MMR) ) (7.13)
mn
n(MMR) = d1 + (d2 − d1 ) · MMR (7.14)

Thereby the variables c1 and c2 are the constants at pure mode 1 and pure mode 2, respectively.
The constant mc controls the behavior of c and has to be determined from mixed mode fatigue
experiments. The same applies to the exponent n.

It can be concluded, the Abaqus low cycle fatigue criterion does not take mixed-mode loading
condition into account. Therefore, fatigue simulations in which the mixed mode ratio changes, are
not covered.

7.2.3 Damage accumulation at the crack front in 3D simulations


In a 3D model, multiple nodes are subjected to a certain strain energy release rate at the crack front.
Therefore, some sort of damage accumulation (virtual crack growth) is needed for the reason that
the bonds which remained were also subjected to fatigue damage already. Since the Abaqus manual
[18] does not mention any damage accumulation in this procedure, an attempt for re-engineering
was made to check the behavior of the solver regarding damage accumulation. In the simulation, a
stress ratio of R=0 and a maximum displacement of dmax =1 mm were used since the output request
was assumed to be the maximum strain energy release rate Gmax in a cycle at that stage of the
work. However, this would not have been necessary since Abaqus returns the relative strain energy
release rates when using the Direct Cyclic approach, as it was found out in Appendix A.

For re-engineering, the strain energy release rate outputs G1 to G3 (ENRRT11 to ENRRT13) were
analyzed at the nodes which debonded in the first three consecutive cycles and summed up to GT ,
since a BK law is used, as mentioned in section 6.2.1. After that, the cycles needed for onset and
propagation, ∆Nonset and ∆Nprop , were calculated using equations 7.15 with the constants given
in table 7.3 and a mesh size a=1 mm. These were derived from the corresponding criteria (see
section 6.2.1).

a
Nonset = c1 · ∆Gc2 Nprop = (7.15)
c · ∆Gn
In addition, the incremental linear damage from the last cycle to the beginning of the corresponding
cycle, ∆Dinc , was calculated as depicted in equation 7.16 and summed up with the accumulated
damage from previous cycles, Dprev , to a total damage at the beginning of the cycle, Dtot . As it
can be seen in the tables below, the crack onset criterion does not affect the crack growth at any
node and cycle in the present example. This behavior will be reviewed in detail in section 7.2.4.

Nk − Nk−1
∆Dinc,k = (7.16)
∆Nprop,k−1
Figure 7.20 shows a fragment of the bond state plot at edge 1 after 527 cycles, which corresponds
to the 4th stabilized cycle. Thereby red colored nodes indicate a bonded node pair (bond state = 1)
while blue colored nodes signal a bond which is released already (bond state = 0). Starting with a
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 59

straight crack front at N=1, the bond at node 1776 releases in the second cycle at N=391, the bond
at node 1778 releases in the third cycle at N=507 and the bond at node 1780 releases in the fourth
cycle at N=527. At edge 2 of the specimen, the same behavior was obtained, since the initial crack
growth was symmetrical as expected.

Figure 7.20: bond state at N=527 (4th stabilized cycle) at edge 1

Tables 7.11 to 7.14 show the extracted data and results from the calculations done for every stabilized
cycle:
J
node Dprev ∆Dinc Dtot bond state ∆GT / mm 2 ∆Nonset ∆Nprop
1776 0 0 0 1 0.417 54 387
1778 0 0 0 1 0.395 76 523
1780 0 0 0 1 0.392 79 546

Table 7.11: strain energy release rates and damages at N=1

J
node Dprev ∆Dinc Dtot bond state ∆GT / mm 2 ∆Nonset ∆Nprop
1776 1.008 0 1.008 0 - - -
1778 0 0.746 0.746 1 0.395 76 523
1780 0 0.715 0.715 1 0.399 71 494

Table 7.12: total strain energy release rates and damages at N=391

J
node Dprev ∆Dinc Dtot bond state ∆GT / mm 2 ∆Nonset ∆Nprop
1776 0 0 1.008 0 - - -
1778 0.746 0.283 1.029 0 - - -
1780 0.715 0.235 0.95 1 0.421 51 368

Table 7.13: total strain energy release rates and damages at N=507
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 60

J
node Dprev ∆Dinc Dtot bond state ∆GT / mm2
∆Nonset ∆Nprop
1776 1.008 0 1.008 0 - - -
1778 1.029 0 1.029 0 - - -
1780 0.95 0.054 1.004 0 - - -

Table 7.14: total strain energy release rates and damages at N=527

As it can be seen, the bond always releases (=bond state switches to 0), when the linear accumulated
damage reaches a value slightly above 1. Therefore it can be said, that the assumed linear damage
accumulation seems to be considered in this procedure.

7.2.4 Impact of the crack onset criterion


Since both criteria for onset and propagation are depending on ∆G (compare section 7.2.1, in
particular figure 7.17), two cases are possible at a certain ∆G:

ˆ The propagation-only case, in which the number of cycles needed for onset is smaller than
the number of cycles needed for crack propagation to the next node at the beginning of the
analysis already.

ˆ The onset-propagation case, in which the number of cycles needed for onset is larger than
the number of cycles needed for crack propagation to the next node.

In the case of the example above in section 7.2.3, where the ENF specimen with the fracture
parameters from table 7.3 are used, the behavior of these criteria can be compared by plotting the
cycles needed for onset and propagation over ∆G by using equations 7.15:

107

106

105
cycles N

104
Nonset
103
Nprop
102

101

100
10−1 100
J
∆G / mm2

Figure 7.21: crack onset and propagation criterion for the ENF mode 2 model over ∆G in the Paris regime
for the nodes at the initial crack front

It can be clearly seen, that in the case of the mode 2 ENF model with the present material parameters
and mesh size, the onset criterion is smaller than the propagation criterion over the whole Paris
regime. Therefore, as already shown in section 7.2.3, the crack onset criterion does not affect the
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 61

crack propagation at any point and hence can be neglected. To analyze the behavior in the case,
where the crack onset criterion actually impacts the results, a simulation with a modified crack
onset constant was created based on the simulation from the section above. Therefore, the constant
c1 for crack initiation was multiplied by 10 to c1 =2.13, which shifts the crack onset curve above the
crack propagation curve in figure 7.21. Tables 7.15, 7.16 and 7.17 show the extracted ∆GT , bond
state, the re-calculated criteria and damages with equations 7.15 and 7.16 for selected nodes. As
in section 7.2.3, Dprev indicates the accumulated damage from previous cycles, ∆Dinc the increase
in damage from the last stabilized cycle to the current one and Dtot the damage at the start of the
current cycle. Thereby, ∆Dinc is set to 0, if the onset criterion is not fulfilled in the current cycle.
J
cycle Dprev ∆Dinc Dtot bond state ∆GT / mm 2 ∆Nonset ∆Nprop onset?
0 0 0 0 1 0.358 1418 906 0
1 0 0 0 1 0.383 932 627 0
156 0 0 0 1 0.382 935 629 0
249 0 0 0 1 0.382 935 629 0
250 0 0 0 1 0.382 936 629 0
252 0 0 0 1 0.382 936 630 0
257 0 0 0 1 0.382 936 630 0
261 0 0 0 1 0.382 937 630 0
266 0 0 0 1 0.382 937 630 0
275 0 0 0 1 0.382 938 631 0
284 0 0 0 1 0.382 938 631 0
285 0 0 0 1 0.382 938 631 0
293 0 0 0 1 0.382 939 631 0
294 0 0 0 1 0.382 939 631 0
301 0 0 0 1 0.382 939 632 0
307 0 0 0 1 0.382 940 632 0
310 0 0 0 1 0.382 940 632 0
312 0 0 0 1 0.382 940 632 0
706 0 0 0 1 0.383 929 625 0
1133 0 0.683 0.683 1 0.386 886 600 1
1188 0.683 0.092 0.774 1 0.391 811 556 1
1230 0.774 0.076 0.850 1 0.401 693 484 1
1263 0.850 0.068 0.918 1 0.422 504 366 1
1281 0.918 0.049 0.967 1 0.424 489 356 1
1285 0.967 0.011 0.979 1 0.416 549 395 1
1291 0.979 0 0.979 1 0.314 3252 1877 0
1332 0.979 0 0.979 1 0.319 2946 1721 0
1343 0.979 0 0.979 1 0.326 2553 1518 0
1359 0.979 0.011 0.989 1 0.362 1311 846 1
1373 0.989 0.017 1.006 0 - - - -

Table 7.15: total strain energy release rates and damages at node 1789 (first row, near the center)
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 62

J
cycle Dprev ∆Dinc Dtot bond state ∆GT / mm 2 ∆Nonset ∆Nprop onset?
706 0 0 0 1 - - - -
1133 0 0 0 1 0.463 277 216 1
1188 0 0.254 0.254 1 0.503 165 138 1
1230 0.254 0.305 0.559 1 0.518 137 117 1
1263 0.559 0.282 0.842 1 0.526 125 108 1
1281 0.842 0.167 1.009 0 - - - -

Table 7.16: total strain energy release rates and damages at node 1878 (at a=1 mm, near edge 1)

J
cycle Dprev ∆Dinc Dtot bond state ∆GT / mm 2 ∆Nonset ∆Nprop onset?
3693 0 0 0 1 - - - -
3696 0 0 0 1 0.381 962 645 1
3698 0 0.003 0.003 1 0.456 305 235 1
3712 0.003 0.059 0.063 1 0.460 291 226 1
3727 0.063 0.066 0.129 1 0.466 269 211 1
3742 0.129 0.071 0.200 1 0.469 256 202 1
3747 0.200 0.025 0.225 1 0.472 248 197 1
3793 0.225 0.234 0.459 1 0.473 244 194 1
3885 0.459 0.475 0.934 1 0.477 231 185 1
3893 0.934 0.043 0.977 1 0.502 166 138 1
3895 0.977 0.014 0.992 1 0.470 253 200 1
3899 0.992 0.020 1.012 0 - - - -

Table 7.17: total strain energy release rates and damages at node 3387 (at a=16 mm, near the center)

In the case of node 1789, which is located at the initial crack front, the crack onset criterion delays
the crack propagation. As it can be seen, the re-calculated damage is in agreement with the response
of the bond state.
At node 1878, which is the first debonding node pair, that is not located at the initial crack front,
the onset criterion is satisfied already. This is also the case at node 3387, which is placed at a crack
length of 16 mm. In both cases, the re-calculated damage confirms the behavior of the bond state
again.
Figure 7.22 shows the crack progress of the simulation with modified onset criterion and the refer-
ence simulation. As expected, the simulation with modified onset criterion delays the initial crack
progress, but the curve converges to the reference simulation later on.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 63

70

60
crack length a / mm

50

40

30 reference
modified onset
20

10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.22: influence of the onset criterion onto the crack progress

Errors induced by the crack onset criterion

In certain simulations, the criterion for crack onset, which is included in the Abaqus low-cycle
fatigue fracture criterion, leads to calculation errors of ∆G2 and ∆G3 . This strongly impacts mode
2 or mode 3 driven crack propagation.
In the following, the behavior is shown in a representative example, which is a modified version
of the setup used in section 7.2.4. Thereby the constant c1 was multiplied by 100 to c1 =21.3.
Furthermore the synthetic load case R=0.426 from section 7.2.1 was used.
Figure 7.23 shows the bond state at 9880 cycles, where red indicates bonded nodes while blue signals
a broken bond. It can be seen, that two bonds remain in the debonded area, namely node 1988
(left) and node 2088 (right).

Figure 7.23: bond state at N=9880

In the damage accumulation table 7.18 for node 1988, which is similar to the tables shown in
section 7.2.4, it can be seen that with decreasing load due to 3D effects on the crack front, the
onset criterion swaps to 0 again, hence blocking further damage accumulation. As a consequence,
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 64

node 1988 and the node behind it, node 2088, form a peninsula until they are undermined and
separated. Since these two nodes are not subjected to the crack front anymore, the strain energy
release rates in mode 2 and 3 cannot be calculated. However, Abaqus is still able to calculate the
mode 1 strain energy release rate, which is not sufficient to force debonding in the present mode 2
ENF simulation, though.

bond ∆G1 ∆G2 ∆G3


∆GT
cycle Dprev ∆Dinc Dtot J J J J ∆Nonset ∆Nprop onset?
state / mm 2 / mm 2 / mm
/ mm2 2

7737 0 0 0 1 0 0 0 0 - - 0
7738 0 0 0 1 0 0.415 0
0.415 5567 399 1
7782 0 0.110 0.110 1 0 0.404 0
0.404 6612 464 1
7856 0.110 0.159 0.270 1 0 0.406 0
0.406 6413 452 1
7857 0.270 0.002 0.272 1 0 0.408 0
0.408 6225 440 1
7894 0.272 0.084 0.356 1 0 0.416 0
0.416 5460 393 1
7895 0.356 0.003 0.358 1 0 0.480 0
0.480 2215 178 1
7938 0.358 0.242 0.600 1 0 0.490 0
0.490 1956 160 1
7966 0.600 0.175 0.776 1 0 0.491 0
0.491 1926 157 1
7967 0.776 0.006 0.782 1 0 0.499 0
0.499 1741 144 1
7968 0.782 0.007 0.789 1 0 0.506 0
0.506 1581 132 1
7969 0.789 0.008 0.796 1 0 0.514 0
0.514 1439 122 1
7970 0.796 0.008 0.805 1 0 0.521 0
0.521 1320 113 1
7978 0.805 0 0.805 1 0 0.341 0
0.341 19160 1180 0
3.2 · 3.2 · 1.23 · 9.51 ·
7979 0.805 0.001 0.806 1 QNAN QNAN 0
10−6 10−6 1036 1030

Table 7.18: total strain energy release rates and damages at node 1988 (third row, near the center, see
figure 7.23)

The behavior of the onset criterion was re-calculated successfully. However, in the Abaqus im-
plementation, the criterion uses ∆GT instead of G1,max . This means, the Abaqus implementation
generalizes the initial model (see section 4.4.2) to mixed mode conditions, which is almost certainly
insufficient apart from using the relative strain energy release rate instead of the maximum value.
In addition, the behavior at node 1789 in the cycles 1291, 1332 and 1343 is physically questionable
since once the onset criterion is fulfilled, it should not affect the propagation anymore.
As a conclusion, the current implementation of the criterion for crack onset in the Abaqus low
cycle fatigue criterion remains highly questionable and should be used with care or just excluded
by setting the constant c1 to 0.

7.2.5 Behavior at phase-shifted and non-sinusoidal cyclic loadings


The Direct Cyclic algorithm delivers the relative strain energy release rate needed for further cal-
culations of the crack growth. To test its robustness and applicability to general load cases, the
initially used sinusoidal loadings were modified in the reference case simulation.

Phase-shifted loadings

At the beginning, different phase shifts t0 were applied to the sine loading of the reference case (see
section 7.1.1), as depicted in figure 7.24:
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 65

0.8
amplitude a(t)

0.6
reference case, t0 =0s
0.4 t0 =-0.05s=+0.15s
t0 =+0.05s
0.2 t0 =+0.1s
t0 =+0.15s
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
time t / s

Figure 7.24: periodic sine loading for R=0.1 with different phase shifts t0

Figure 7.25 shows the resulting output curves of the mode 2 strain energy release rate (ENRRT12)
at the first fatigue cycle:

0.5

0.4

0.3
mm2
J
G2 /

reference case, t0 =0s


0.2 t0 =-0.05s
t0 =+0.05s
t0 =+0.1s
0.1 t0 =+0.15s

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
time t / s

Figure 7.25: evolution of the output variable ENRRT12 over a single loading cycle with different phase
shifts t0

The following statements are obtained:

ˆ The maximum relative strain energy release rate ∆Gmax is held till the end of a stabilized
cycle. It occurs when reaching the maximum load.
ˆ The level of the maximum energy release rate is dependent on the position in a cycle when
facing the same load.
ˆ The relative strain energy release rate output is not increased until the slope of the load is
not negative anymore and the initial value is reached (compare t0 =+0.1s).
ˆ The following behavior remains unclear:
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 66

– In the case of t0 =+0.05s, an unexpected rise followed by a plateau is obtained in the first
quarter cycle.
– the same loading (t0 =-0.05s and t0 =+0.15s) does not give the same response.

For further investigation, some more cases were simulated and the maximum values of ENRRT12 at
the end of the first stabilized cycle were plotted over the phase shift t0 , which is shown in Figure 7.26:

0.5
mm2
J

0.4
maximum ENRRT12 /

0.3

0.2

0.1

0
−0.1 −0.05 0 0.05 0.1 0.15 0.2
phase shift t0 / s

Figure 7.26: maximum output value of ENRRT12 of simulations with different phase shifts t0

It can be seen that the maximum value is obtained at a phase shift t0 of 0s. Negative phase shifts
result in marginally lower values, higher phase shifts than t0 =+0.075 lead to significantly lower
values. It is questionable that simulations with the sine shifted by a whole cycle, e.g. t0 =-0.05 and
t0 =+0.15, do not lead to the same solution. Increasing the phase shift to the cycle time of 0.2s and
beyond causes much smaller strain energy release rates, hence these results are clearly wrong.
Since ∆G controls the whole crack progress, the shifted levels of the relative strain release rate
outputs influence the whole crack propagation, hence shifting the whole curve while keeping the
same shape, as shown for selected cases in figure 7.27:
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 67

50

40
crack length a / mm

30

reference case, t0 =0s


20 t0 =-0.05s
t0 =+0.05s
t0 =+0.1s
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.27: comparison of crack propagation curves resulting from different phase shifts

Non-sinusoidal cyclic loadings

In the next step, loading cases in which multiple nested cycles in a given cycle duration or points
of discontinuity occur, were evaluated. Latter refers to loadings with swapping load vector in
mode 2 (0 < R < -∞), since a single loading cycle can be interpreted as two individual half
cycles due to the fact that the acting shear stresses affect the crack propagation in positive and
negative direction (linear-elastic material). This also shows multiple peaks of the maximum strain
energy release rates. Therefore, a simulation using R=-1 with scaled maximum displacement for
the reason of comparability to the reference case concerning ∆Gmax was created with the findings
from section 7.2.1. Note that these simulations must be considered synthetic since in the current
test setup, no tension forces can be applied. In addition, a second simulation using an amplitude
curve which represents R=-1 but with a mirrored second half-cycle was created to test the treatment
of a point of discontinuity in the displacement function. Eventually a simulation using R=0 and
two cycles was built up, since this simulation shows peaks with the same maximum relative strain
energy release rate. The corresponding amplitude curves are shown in figure 7.28:
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 68

0.5
amplitude a(t)

0
reference case
R=-1
−0.5 R=-1 absolute
R=0, two cycles
−1
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
time t / s

Figure 7.28: periodic loadings

For the reason of convergence problems in the simulation with R=-1 and absolute amplitude, the Dir-
ect Cyclic control parameters for displacement corrections (3rd and 5th entrance, see section 7.1.4),
which were set to 1 · 10−3 in the reference case, were increased to 5 · 10−3 (compare Appendix D).
This is caused by the point of discontinuity in the function, which complicates the search for a
proper Fourier representation. In Figure 7.29 shows the evolution of G2 over the first stabilized
cycle. As it can be seen, the response of G2 differs at the beginning when comparing R=0 with
R=-1, but reaches about the same level at 1/4 of the cycle, as expected in the case without phase
shift. In the case of R=-1 with the mirrored second half cycle, the maximum strain energy release
rate reaches a lower level.

0.5

0.4

0.3
mm2
J
G2 /

reference case
0.2 R=-1
R=-1, absolute
R=0, two cycles
0.1

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
time t / s

Figure 7.29: ∆G2,max over a single loading cycle

Figure 7.30 depicts the resulting crack propagation curves. Thereby the curve from R=0 overlaps
with the reference case. As a result of lower ∆GT,max , the curve from R=-1 and absolute amplitude
is shifted to the right by about factor 2, which means that the crack propagates at half velocity.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 69

50

40
crack length a / mm

30

reference case
20 R=-1
R=-1, absolute
R=0, two cycles
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure 7.30: comparison of the crack propagation curves

In summary, the levels of the maximum relative strain energy release rate outputs are influenced
by the position of the maximum loading in the cycle. When it occurs at 1/4 of the cycle duration,
which corresponds to a simple sine loading, the relative strain energy release rate outputs coincide
with the level from static calculations, while later occurrence causes lower outputs. Multiple load
peaks in a cycle seem not to be detected. However, the case of R=-1 returns a slightly higher output
than R=0 despite having the same position of the (first) maximum load.

7.2.6 CPU-parallelization, cycle limit


Due to the high calculation times needed for simple models already, the effects and efficiency of
CPU-parallelization were tested. Therefore, a simulation with mesh B 2 mm, automatic contact
stabilization and the reference loading case of R=0.1 was started with 1 and 3 CPU cores on the same
computer and the results were compared to each other. Figures 7.31 depict the crack propagation
curves for considered node sets. It can be seen, that CPU parallelization impacted the results on
the edges in the region after passing the point of load entry in the middle of the specimen. Due to
CPU-parallelization, the calculation time decreased by 40%.
crack length a / mm

45 45 45

30 30 30
1 CPU
15 15 15 3 CPUs

0 0 0
100 105 100 105 100 105
cycles N cycles N cycles N

Figure 7.31: comparison of the crack propagation curves regarding CPU-parallelization: edge 1 (left), center
(mid) and edge 2(right)
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 70

In another attempt, a simulation with smaller mesh size was taken (mesh A, R=0, dmax =1 mm,
viscosity=1 · 10−6 ). Thereby, simulations with 1, 3 and 4 CPUs were compared to each other. After
passing the line of load introduction, small deviations arise, which decay with rising crack length,
as it can be seen in figures 7.32. Compared to the simulation with only 1 CPU, the calculation time
decreased by about 30% in the case of 3 CPUs and by 31% in the case of 4 CPUs.
crack length a / mm

50 50 50

25 25 25 1 CPU
3 CPUs
4 CPUs
0 0 0
100 105 100 105 100 105
cycles N cycles N cycles N

Figure 7.32: comparison of the crack propagation curves regarding CPU-parallelization: edge 1 (left), center
(mid) and edge 2(right)

Tests at the "Mach" mainframe of the University have shown, that excessive parallelization with 8
cores and more does not decrease the calculation time significantly. Therefore, it is not recommended
to use more than 3 CPUs, since the parallelization does not work that well using the Direct Cyclic
approach. However, high clock rates and fast storage devices significantly accelerate the calculation.

In addition, tests with a very high number of cycles have shown, that the calculation crashes above
about 2 · 109 cycles, since the step time jumps to negative values there. Therefore it seems that the
cycle number is a 32bit integer. A workaround for simulating even a higher number of cycles, an
additional equal Direct Cyclic step can be created.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 71

7.3 Guidelines for implementation and input parameters


In the course of the work, several parameters were found for successful, efficient fatigue crack growth
simulations using the Abaqus VCCT low cycle fatigue analysis. These are listed in the following:

ˆ For accurate results, especially when using 3D models, solid elements are recommended over
continuum shell elements, due to overestimation of the strain energy release rates and pos-
sible convergence problems [39] when using latter. For bending problems, linear elements
with incompatible node formulation are preferred due to their significantly better bending
representation.

ˆ Element sizes of 0.1 ≤ ∆a/h ≤ 1 are recommended for composites according to [19]. Fur-
thermore until the treatment of different element sizes over the estimated crack surface is
validated, it should be kept constant.

ˆ Sinusoidal fatigue loadings without phase shift are recommended for proper calculation of the
strain energy release rates via the Abaqus Direct Cyclic Approach (compare section 7.2.5).

ˆ According to current knowledge, crack propagation over the line of load introduction should
be avoided and if not, evaluated with care (compare section 7.1.5 and Appendix C).

ˆ The crack onset criterion should be excluded in 3D simulations since it can lead to calcula-
tion errors (see section 7.2.4). Furthermore its implementation is questionable anyway since
∆GT is used instead of G1,max , which is taken in in the original proposal of the theory (see
section 6.2.1).

ˆ According to Krueger [49], the maximum number of Fourier terms (50) are recommended
for accurate results. However, tests have shown, that the automatic incremental definition is
capable of returning highly accurate results at slightly decreased calculation times, too (see
Appendix D).

ˆ The Direct Cyclic control parameters are essential for convergence, accuracy and calculation
time and hence have to be set with care. In cases where convergence is not given when
using the default parameters, these should be increased step-by-step. In the present case,
the parameter which is responsible for the residual displacement criterion was significantly
increased to deactivate the corresponding criterion. Furthermore in Appendix D, modifying
the other parameters was tested.

ˆ Stabilization or viscous regularization is usually not required. When using fixed time steps,
which is required in non-dissipative simulations, the release tolerance has no effect.

ˆ In a single Direct Cyclic step, only 2 · 109 cycles are possible. For larger numbers, multiple,
equal Direct Cyclic steps have to be created.
7 Simulations and validations of Abaqus VCCT low cycle fatigue method using 3D elements 72

7.4 Summary of the assessment regarding fatigue delamination under cyclic


loading
The Abaqus VCCT low cycle fatigue method offers the following capabilities and advantages:

ˆ the approach is based on physical assumptions

ˆ the parameters needed can be obtained from classical fracture mechanics tests

ˆ the effect of stress ratio is covered in the way that equal values of the relative total strain
energy release rate return equal rates of crack growth (see section 7.2.1)

ˆ crack onset, which is needed for e.g. ply drop-offs, is included

ˆ in 3D simulations, the virtual crack growth of the non-debonding nodes at the crack front is
taken into account (see section 7.2.3)

However, the method still has clear limits and drawbacks, which are:

ˆ Simulations with changing mixed-mode ratio (MMR) are not covered in the low cycle fatigue
criterion due to the 1D implementation of Paris law. Therefore, it is only applicable to very
simple applications such as typical fracture mechanics specimens.

ˆ The crack onset criterion’s implementation is questionable, since the equivalent relative strain
energy release rate ∆G is taken instead of the maximum mode 1 strain energy release rate
G1,max . Furthermore it can lead to unreasonable calculation errors in 3D crack fronts.

ˆ The method does not work properly when large stiffness differences between the layers occur
[53].

ˆ Certain simulations in which the crack passes the line of load introduction show unreasonable
strain energy release rates in certain modes after passing it. Further research is recommended
regarding this behavior.

ˆ The Direct Cyclic step is very sensitive to the applied load amplitude curve. A simple phase
shift can significantly change the results. For the reason of the Fourier series representation
of the displacement and residual vectors, points of discontinuity lead to convergence problems
or unreasonable results.

ˆ Due to the VCCT implementation, an initial crack and a predefined crack surface is needed.
Furthermore a more or less serrated crack propagation is obtained, depending on the mesh
size. However, to damp this behavior, which is especially observable in 2D simulations, the
debonding of node pairs can be modeled gradually, see section 7.1.2 and [54].

ˆ Calculation times are very high for simple models already, hence complex 3D geometries would
need an unbearable amount of computational resources.

For an application to the composite rim, which was used as the motivation for this work, mixed
mode treatment would have been needed. Furthermore, crack initiation is not covered by VCCT,
which is another major drawback. In addition, the erratic behavior of the Direct Cyclic algorithm
for phase shifted and more general loading functions is not clear and hence a possible source of
errors. Eventually, the calculation times would have become unbearable. Therefore, no simulations
of the composite rim were performed.
8 Conclusions and future work 73

8 Conclusions and future work


8.1 Research goals and performed work
The aim of this work was to give an overview of the available modeling techniques for fatigue damage
of laminated composite structures with a focus on interlaminar damage. Therefore, an extensive
review of methods to describe the complex fatigue behavior of laminated composite materials was
made. In a second step, promising approaches of certain modeling techniques in finite element
packages - namely from Siemens Samtech Samcef and 3DS Abaqus - were further investigated and
assessed regarding its capabilities and limitations in their implementation, if already included in the
corresponding finite element code. In addition, guidelines for proper implementation were derived
based on experience.

8.2 Conclusions and recommendations for future work


Fatigue of laminated composite materials is still at the very beginning of its development since
even the static behavior is not fully clarified at the moment. This results from the complex damage
mechanisms which occur due to multiple constituents of this material. In early attempts, S/N-curve
approaches from metals were used to describe damage in laminates. However, since laminated com-
posites can be varied in their ply stacking sequence, as well as their fiber and matrix types, an
immense amount of testing is needed. Moreover, these models do not consider the measurable stiff-
ness degradation during fatigue life, which causes stress redistribution in the structural component
during fatigue life.
Other fatigue life approaches focus on the residual mechanical properties such as strength or stiffness.
Especially the residual stiffness approaches seem to be very promising since stiffness can be measured
by non-destructive testing methods besides its characteristic behavior of exhibiting three different
stages (compare figure 5.2).
Compared to laminate based modeling, ply level modeling represents an alternative technique.
This method applies a fatigue law directly to the individual plies to model intralaminar damage. To
include interlaminar fatigue damage, these methods are usually combined with fracture mechanics
and continuum damage mechanics methods - namely cohesive zone methods. However, most of these
ply based models do not consider the interactions between intralaminar and interlaminar matrix
failure. Furthermore, as determined in section 7.2 for VCCT for example, these methods are still
very limited and inefficient in calculation.
The most promising intralaminar composite fatigue models are progressive damage models, which
use a continuously rising damage variable to represent the damage state of the laminas or the
laminate. Stiffness reduction is calculated according to the arising damage. To include stress
redistributions, the FE-model is recalculated after a certain increase in damage. However, these
models need many input parameters and are still in an early stage of development.
In the last years, new micromechanical approaches emerged, primarily for static strength. These
models focus on a representative volume element, describing the smallest possible repeating unit of
a certain lamina or laminate, in which the physical behavior is described. The resulting response
can be homogenized in a CDM approach and used for macroscopic application further on. This
procedure is called multi-scale modeling. According to the author, these models should be partic-
ularly suitable for the derivation of fatigue properties. The following multi-scale fatigue procedure
would be conceivable for an applicable approach:

ˆ For a fundamental understanding of the occurring damage mechanisms considering interac-


tions between intra- and interlaminar damage, a micromechanics model representing the lam-
inate is suggested. Furthermore its in-plane dimensions should be sufficiently large to cover
all damage mechanisms, including micro-buckling in compression. Thus, the accurate fatigue
8 Conclusions and future work 74

behavior of the matrix, the fiber and the interface should be represented by input parameters,
as they are intrinsic and the cause of all damage appearance.

ˆ The statistical nature of the main material parameters and the production tolerances should
be evaluated carefully and considered in the micromechanics model to include the evolution of
damages in a statistical manner. Thus, an accurate classification of manufacturing tolerances is
recommended since even small change in the fiber angle or fiber waviness cause high deviations
in the material response and damage mechanisms.

ˆ From the statistical micromechanics model, continuous residual stiffness and residual strength
evolution curves and their probability can be derivated for certain loading conditions and
states of damage. These can be calculated beforehand and stored in databases.

ˆ For finite element modeling of structural elements, shell and continuum shell elements should
be preferred over solid elements and further adapted to the composite’s needs due to their
computational efficiency and accurate representation of thin structures. As in other progressive
damage techniques, the structural response and damage state should be derived incrementally
at certain time steps until failure.

ˆ The whole procedure should be developed for flat structures first and extended to curved
structures in a next step.

ˆ In service, the remaining stiffness of a structure could be monitored within a structural health
monitoring (SHM) system, since the remaining fatigue life of a structure can be accurately
estimated from that remaining stiffness.

For this presented approach, extensive work in the field of micromechanics and deep understanding
of the elemental behavior, especially of the polymer matrix and the fiber-matrix interface, is needed.
Therefore, existing testing methods for each individual materials have to be evaluated regarding
their applicability for determining the data required in a micromechanical model and, if necessary,
modified. In addition, the statistical behavior of each constituent and input parameter has to be
determined - starting from the intrinsic material behavior up to the manufacturing of the structural
component - since all these influences sum up. This may better explain the very high scatter of
material tests of laminated composites.

As it can be seen in the present work, the complex fatigue behavior of composite materials is still
an unsolved question. This often leads to significant oversizing and improper design despite high
testing costs, which impairs the immense potential for lightweight structures of composite materials.
Therefore, despite the big challenge due to the high complexity, further fundamental research for
more optimized laminated composite structures is desirable to eventually exploit the performance
of laminated composite materials.
Literature 75

References
[1] Wiedemann, J.: Leichtbau: Elemente und Konstruktion. Springer-Verlag, 2007.
[2] Schuermann, H.: Konstruieren mit Faser-Kunststoff-Verbunden. Springer, 2007.
[3] Galucio, A.; Jetteur, P; Trallero, D and Charles, J.: ‘Toward numerical fatigue pre-
diction of composite structures: application to helicopter rotor blades’. In: 3rd ECCOMAS
Thematic Conference on Mechanical Response of Composites. Vol. 94. 2011.
[4] Putnam, T. W.: X-29 flight-research program. 1984.
[5] Shirk, M. H.; Hertz, T. J. and Weisshaar, T: ‘Aeroelastic tailoring-theory, practice, and
promise.’ In: J. AIRCRAFT. 23.1 (1986), pp. 6–18.
[6] Griffith, A. A.: ‘The Phenomena of Rupture and Flow in Solids’. In: Philosophical Trans-
actions of the Royal Society of London. Series A, Containing Papers of a Mathematical or
Physical Character. 1920.
[7] Orowan, E.: ‘Fracture and Strength of Solids’. In: Rep. Progr. Phys. 185 (1948).
[8] Gross, D. and Seelig, T.: Bruchmechanik. 6. Auflage. 2016.
[9] Sharpe, W. N.: Handbook of Experimental Solid Mechanics. 2008. Chap. 5.
[10] Roylance, D.: ‘Introduction to fracture mechanics’. In: Massachusetts Institute of Techno-
logy, Cambridge (2001).
[11] Benzeggagh, M. L. and Kenane, M.: Measurement of mixed-mode delamination fracture
toughnesses of unidirectional glass/epoxy composites with mixed-mode bending apparatus. re-
port. Université de Technologie de Compiègne, 1996.
[12] Barbero, E. J.: Introduction to Composite Materials Design. Ed. by Edition, S. 2011.
[13] Denkendorf, I.: Hochleistungsfasern für Verbundwerkstoffe. 2015.
[14] Campbell, F.: Introduction to Composite Materials. ASM International, 2010.
[15] Puck, A.: Festigkeitsanalyse von Faser-Matrix-Laminaten. 1996.
[16] Khashaba, U.: Delamination in drilling GFR-thermoset composites. 2003.
[17] Hognestad, G.: ‘The use of C-shaped specimens to investigate the interlaminar fracture of
woven composite laminates’. In: Journal of Strain Analysis Vol 30 No 2 1995 (1995).
[18] Simulia, D.: Abaqus 6.14 User Manual. 2015.
[19] Krueger, R.: The Virtual Crack Closure Technique: History, Approach and Applications.
ICASE, Hampton, Virginia. 2002.
[20] Dugdale, D. S.: ‘Yielding of steel sheets containing slits’. In: Journal of the Mechanics and
Physics of Solids 8.2 (1960), pp. 100–104.
[21] Barenblatt, G. I.: ‘The mathematical theory of equilibrium cracks in brittle fracture’. In:
Advances in applied mechanics 7 (1962), pp. 55–129.
[22] LS-Dyna: LS-Dyna Keyword User’s Manual Volume II Material Models. Version R7.0. 2013.
[23] Turon, A.; Dávila, C. G.; Camanho, P. P. and Costa, J.: ‘An engineering solution for
using coarse meshes in the simulation of delamination with cohesive zone models’. In: (2005).
[24] Radaj, D. and Vormwald, M.: Ermüdungsfestigkeit. 2007.
[25] Rösler, J.; Harders, H. and Bäker, M.: Mechanisches Verhalten der Werkstoffe. 2012.
[26] Vassilopoulos, A. and Keller, T.: Fatigue of Fiber-reinforced Composites. 2011.
Literature 76

[27] Hammer, F. and Leopold, J.: Großschaden an einem 300MW Dampfturbosatz. 1988.
[28] Wiedemann, J: Leichtbau Elemente und Konstruktion. 2007.
[29] Murri, G. B.; Salpekar, S. A. and O’Brien, T. K.: ‘Fatigue delamination onset prediction
in unidirectional tapered laminates’. In: Composite Materials: Fatigue and Fracture (Third
Volume). ASTM International, 1991.
[30] O’Brien, T. K.; Johnston, W. M. and Toland, G. J.: ‘Mode II interlaminar fracture
toughness and fatigue characterization of a graphite epoxy composite material’. In: (2010).
[31] Paris, P. C. and Erdogan, F.: ‘A critical analysis of crack propagation laws’. In: ASME.
1963.
[32] Walker, K: ‘The effect of stress ratio during crack propagation and fatigue for 2024-T3 and
7075-T6 aluminum’. In: Effects of environment and complex load history on fatigue life. ASTM
International, 1970.
[33] Vassilopoulos, A. P.; Manshadi, B. D. and Keller, T.: ‘Piecewise non-linear constant life
diagram formulation for FRP composite materials’. In: International journal of fatigue 32.10
(2010), pp. 1731–1738.
[34] Philippidis, T. and Vassilopoulos, A.: ‘Fatigue design allowables for GRP laminates based
on stiffness degradation measurements’. In: Composites science and technology 60.15 (2000),
pp. 2819–2828.
[35] Hashin, Z. and Rotem, A.: ‘A fatigue failure criterion for fiber reinforced materials’. In:
Journal of composite materials 7.4 (1973), pp. 448–464.
[36] Tsai, S. W. and Wu, E. M.: ‘A general theory of strength for anisotropic materials’. In:
Journal of composite materials 5.1 (1971), pp. 58–80.
[37] Plumtree, A and Cheng, G.: ‘A fatigue damage parameter for off-axis unidirectional fibre-
reinforced composites’. In: International Journal of fatigue 21.8 (1999), pp. 849–856.
[38] Degrieck, J. and Van Paepegem, W.: ‘Fatigue damage modeling of fibre-reinforced com-
posite materials: Review’. In: Applied Mechanics Reviews 54.4 (2001), pp. 279–300.
[39] Camanho, P. P. and Hallett, S. R.: Numerical modelling of failure in advanced composite
materials. Woodhead Publishing, 2015.
[40] Bak, B. L.; Sarrado, C.; Turon, A. and Costa, J.: ‘Delamination under fatigue loads in
composite laminates: a review on the observed phenomenology and computational methods’.
In: Applied Mechanics Reviews 66.6 (2014), p. 060803.
[41] Carrella-Payan, D et al.: ‘Implementation of fatigue model for unidirectional laminate
based on finite element analysis: theory and practice’. In: Frattura ed Integrità Strutturale 38
(2016), p. 184.
[42] Carrella-Payan, D et al.: Implementation of Fatigue Model for UD Laminate based on
FEA: Theory and Practice. 2016.
[43] Van Paepegem, W: ‘Development and finite element implementation of a damage model for
fatigue of fibre-reinforced polymers’. PhD thesis. University of Ghent, 2002.
[44] Brokate, M.; Dreßler, K. and Krejci, P.: Rainflow Counting and Energy Dissipation for
Hysteresis Models in Elastoplasticity. report. University of Kaiserslautern, 1996.
[45] Van Paepegem, W. and Degrieck, J.: ‘Calculation of damage-dependent directional failure
indices from the Tsai–Wu static failure criterion’. In: Composites science and technology 63.2
(2003), pp. 305–310.
Literature 77

[46] Van Paepegem, W. and Degrieck, J.: ‘Simulating damage and permanent strain in com-
posites under in-plane fatigue loading’. In: Computers and structures 83.23 (2005), pp. 1930–
1942.
[47] Bruyneel, M. et al.: Calculating Fatigue and Fatigue Failure of Structures. US Patent App.
14/273,781. 2014.
[48] Krueger, R.: Development of a benchmark example for delamination fatigue growth predic-
tion. 2010.
[49] Krueger, R.: ‘Development and application of benchmark examples for mode II static
delamination propagation and fatigue growth predictions’. In: (2011).
[50] Krueger, R.: ‘An approach to assess delamination propagation simulation capabilities in
commercial finite element codes’. In: (2008).
[51] Khan, R.; Alderliesten, R.; Badshah, S. and Benedictus, R.: ‘Effect of stress ratio or
mean stress on fatigue delamination growth in composites: critical review’. In: Composite
Structures 124 (2015), pp. 214–227.
[52] Kenane, M; Azari, Z; Benmedakhene, S and Benzeggagh, M.: ‘Experimental develop-
ment of fatigue delamination threshold criterion’. In: Composites Part B: Engineering 42.3
(2011), pp. 367–375.
[53] Krueger, R.; Shivakumar, K. N. and Raju, I. S.: ‘Fracture Mechanics Analyses for Interface
Crack Problems-A Review’. In: 54th AIAA/ASME/ASCE/AHS/ASC Structures, Structural
Dynamics, and Materials Conference. 2013, p. 1476.
[54] Bisagni, C.; Brambilla, P. and Bavila, C. G.: Modeling Delamination in Postbuckled Com-
posite Structures Under Static and Fatigue Loads. 2013.
Appendix 78

Appendix A Additional information to the strain energy release rate outputs


In the course of the work, it was found out that the Abaqus manual [18] gives wrong information
about the energy release rate outputs in the direct cyclic analysis in combination with the low
cycle fatigue criterion. According to the definition in "Output variables", chapter 11.4.3 - Crack
propagation analysis in the Abaqus Analysis User’s Guide, the (current) strain energy release rate is
displayed using the VCCT, enhanced VCCT or low-cycle fatigue criterion. Furthermore, in chapter
6.2.7 - Low-cycle fatigue analysis using the direct cycle approach, no additional information is
provided regarding the outputs. However, using the direct cyclic approach in combination with the
low cycle criterion and VCCT crack growth, the strain energy release rate outputs ENRRT give the
maximum ∆G = Gmax − Gmin to the corresponding increment in a cycle. This statement is proved
in the following lines.

With the findings from section 7.2.1, load cases with different stress ratios, minimum and maximum
displacements but same relative strain energy release rate ∆G could be created, as depicted in
figure A.1:

1.2

1
amplitude a(t)

0.8

0.6
reference case
0.4 R=0, dmax =0.995 mm
R=0.426, dmax =1.1 mm
0.2

0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
time t / s

Figure A.1: periodic sine loading with equal ∆G

In figure A.2, the resulting mode 2 strain energy release rate outputs (ENRRT12) are plotted over
the first stabilized cycle. In addition, levels with the maximum energy release rates for all three
cases and the relative strain energy release rate over the cycle were added.
Appendix 79

0.6

0.5

0.4
mm2
J

0.3
G2 /

Gmax at R=0.1
Gmax at R=0.426
0.2 Gmax =∆G at R=0
reference case
0.1 R=0, dmax =0.995 mm
R=0.426, dmax =1.1 mm
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2
time t / s

Figure A.2: evolution of G2 over a single loading cycle with different loadings of equal ∆G

All of the output curves converge to Gmax of R=0, which corresponds to ∆G in all loadings. Since
the overlap occurs at 1/4 of the cycle, which coincides with the point of maximum displacement,
it may be assumed that these outputs provide the maximum ∆G from the start of a cycle until
the current increment. However, this is also influenced by the point of maximum displacement in a
cycle, as described in section 7.2.5.
Appendix 80

Appendix B Further study of the automatic contact stabilization at a mode 1


DCB model
For further study of the high normal damping of the automatic contact stabilization, a mode 1 DCB
model was created.

Appendix B.1 Double cantilever beam (DCB) FE model


For investigating interlaminar crack propagation in mode 1, the double cantilever beam (DCB)
specimen depicted in figure B.3 is commonly used. To test the appropriate solver parameters found
for mode 2 crack propagation in mode 1, a DCB model was created. Because of the need for
appropriate material and test data, the case study was leaned on simulations from R. Krueger [48].
In this case, a specimen consisting of 24 0° UD carbon epoxy plies was used. Table B.1 shows
the dimensions of the specimen. It is positioned in the middle of the layup, between the 12th and
13th ply. On the hinges, which are glued onto the specimen on both sides, the cyclic displacement
loadings, ~uu (t) and ~ul (t), are introduced.

B
~uu (t)
2h

a0 L z
y
~ul (t) x

Figure B.3: DCB test configuration with pre-cracked specimen

parameter value / mm
L 150
B 25
a0 30.5
h 1.5

Table B.1: dimensions of the DCB specimen


Appendix 81

Model description

Figure B.4 shows the simplified model used for the simulations. Since the sections behind the hinges
do not exhibit transverse forces, they are unloaded and hence can be neglected.
bonded face

~uu (t) 2h

δ z

~ul (t) y
x
a0 L
B

Figure B.4: simplified DCB model

Material properties

The material properties of the lamina and the 0°/0° interface properties were taken from [48].
Table B.2 depicts the lamina properties whereas the interface properties are depicted in table B.3.
The latter are divided into static fracture properties using a BK law on the left, delamination growth
onset coefficients in the middle and Paris law coefficients for delamination propagation on the right.

E1 E2 = E3 ν12 = ν13 ν23 G12 = G13 G23


139.4GPa 10.16GPa 0.3 0.436 4.6GPa 3.54GPa

Table B.2: material properties for the graphite epoxy lamina used

G1C G2C = G3C η c1 c2 Gth c n


0.17J/mm² 0.49J/mm² 1.62 2.8 10−9 -12.415 0.06J/mm² 2.44 106 10.61

° °
Table B.3: fracture parameters for graphite epoxy 0 /0 interface used

Simulation steps

Equally to the ENF model in section 7.1.1, the same solver, steps and parameters, as depicted in
table 7.4, were used. As a result of a loading frequency of 10Hz, a single cycle was divided into
100 increments each. Regarding the direct cyclic control parameters, these were set according to
Krueger’s simulation in [48]. Thereby, the ratio of the maximum residual coefficients to the time
averaged force CRnα , was set to 100, which means that the criterion was turned off. The other
control parameters were kept at default values.
Appendix 82

VCCT interaction

The interface was modeled in the same way as in the ENF specimen, which is described in sec-
tion 7.1.1. Thereby the material parameters were taken from table B.3.

Boundary conditions and displacement loadings

At the idealized lines of the hinges (green lines in figure B.4), translation was locked in x- and
y-direction. In the static step, a displacement of δmax /2 = 0.00067 mm was applied on each hinge in
crack opening z-direction, respectively. In the direct cyclic fatigue step, periodic sine loadings with
a frequency of 10Hz were applied in crack opening z-direction each. As a result, the same approach
was used as in the ENF model shown in section 7.1.1, except for a doubled angular frequency of
62.832Hz. Equations B.1 and B.2 depict the corresponding displacement vectors.
 
0
~uu (t) = d~u · a(t) d~u =  0  mm (B.1)
 

δmax /2
 
0
~ul (t) = d~l · a(t) d~l =  0  mm (B.2)
 

−δmax /2

Element type, node sets for evaluation

As in the ENF model (section 7.1.1), linear hexahedral elements with incompatible mode formulation
were used since less elements are needed through the thickness of each stack then. The same applies
to the node sets for evaluation (section 7.1.1).

Appendix B.2 Setting of the simulations


The mesh for the DCB model was modeled in a similar way as mesh B, which can be found in
section 7.1.5. Figure B.5 shows the partitioning of a stack. Thereby the less important partitions
A with a length of a = 20.5 mm and C with a length of c = 69.5 mm were assigned a coarse mesh in
length direction whereas in partition B with a remaining length of 60 mm, a finer mesh was used.
Table B.4 depicts the number of elements over the partitions and the resulting element sizes.
z
a c

A B C h x
L y

Figure B.5: partitioning of the DCB model


Appendix 83

number of elements resulting element size


partition A 7 ∼3 mm
partition B 48 1.25 mm
partition C 23 ∼3 mm
width B 20 1.25 mm
height H 2 0.75 mm

Table B.4: element data in the mesh B models

In figure B.6 the resulting mesh is shown. As in the visualizations of the ENF meshes, the red area
indicates the bonded nodes of the VCCT contact. It can be seen that the mesh refinement is drawn
beyond the initial crack front for the reason of better calculation of the bending deformation near
the crack tip.

Figure B.6: mesh of the DCB model

The same solver parameters as in the ENF model were used. Two cases were created: one with
automatic contact stabilization and one with very little manual contact stabilization, as suggested
in [49].

Appendix B.3 Results and discussion


Figure B.7 shows the center crack propagation curve of these two cases. It can be seen that the
simulation using constant stabilization follows Krueger’s semi-analytical benchmark curve from [48].
The overestimation is induced by the very low number of elements over the thickness. In the case of
automatic stabilization, the crack propagates significantly too slow due to notably reduced relative
mode 1 strain energy release rates ∆G1 . This is induced by the high normal damping of the
automatic contact stabilization.
Appendix 84

12

10
crack length a / mm

6
automatic stabilization
constant stabilization 10−6
4
benchmark curve

0
100 101 102 103 104 105 106 107 108
cycles N

Figure B.7: comparison of the crack propagation curves

0.25
∆G1 / 103 mJ2

0.2

automatic stabilization
0.15 constant stabilization 10−6

0.1

0.05
0 1 2 3 4 5 6 7 8 9 10
crack length a / mm

Figure B.8: comparison of ∆G1 over the crack length

Comparison of the increase of static dissipation energy to the maximum strain energy in a cycle
show that these have the same order of magnitude in the case of automatic stabilization, whereas
using constant stabilization, ALLSD/ALLSE is under 1ppm. The excessive normal damping when
using automatic stabilization can be even seen in the deformation, as depicted in figure B.9. There,
the end of the red line indicates the crack front.
Appendix 85

Figure B.9: upper figure: deformation with low constant stabilization; lower figure: deformation using
automatic stabilization
Appendix 86

Appendix C ENF Simulations using an undistorted mesh and non simplified


geometry
To inspect the effect of mesh distortion in Mesh B in the ENF model and a possible influence of
the unloaded regions, two similar models, one full size model and one with the unnecessary regions
cropped, were created with imperial mesh size dimensioning.
In mesh B, the upper stack is slightly distorted because of the imperial sized point of load entrance
at the upper surface combined with a mesh size of 1 mm in the significant region of the VCCT
contact at the lower surface.

Appendix C.1 Setting of the simulations


Figure C.10 depicts the full size model while in the simplified model, the mesh was cropped at the
supports (red lines). The corresponding dimensions can be found in table 7.1. As well as in mesh
B, the length of partition A to the support, a, was kept at 25.4 mm (1 inch) whereas c was shifted
to 12.7 mm (1/2 inch). In the full size model, the unloaded part behind the crack e was set to 1
inch whereas the total length Ltot was set to 7 inch, resulting in an unloaded length of 2 inch in the
pre-cracked area. These dimensions were taken from [30].
z
a = a0 c e

AL A B C CL h x
2L y
Ltot

Figure C.10: partitioning of mesh B

Table C.5 depicts the mesh sizes in each direction and partition used:

number of elements resulting element size


partition AL 28 ∼1.8 mm
partition A 14 ∼1.8 mm
partition B 50 1.27 mm
partition C 7 ∼1.8 mm
partition CL 14 ∼1.8 mm
width B 30 ∼0.85 mm
height H 2 1.125 mm

Table C.5: element data in the the undistorted mesh models

Material properties, simulation steps, boundary conditions and displacement loading were kept
unchanged compared to the other ENF models described in section 7.1.1 and manual stabilization
of 1 · 10−6 up to a clearance of 0.1 mm according to [49] was used which is very small. The release
tolerance, which should not affect the fatigue analysis anyway, was set to 1 · 10−3 .
In addition, the simulation with a propagated crack was repeated. Therefore, the initial crack was
shifted to a=30.48 mm.

Appendix C.2 Results and discussion


The resulting crack propagation curves are compared in figure C.11. It can be seen that the curves
are overlapping until a crack length of 40 mm. After that, the simplified, undistorted model delivers
Appendix 87

a marginally stiffer response, increasing with crack length. The more clunky crack propagation of
mesh B at the beginning results from the higher mesh size of 1.5 mm instead of 1.27 mm.

70

60
crack length a / mm

50

40

30 simplified model,
undistorted mesh
20 full size model,
undistorted mesh
10 mesh B 1.5 mm

0
100 101 102 103 104 105 106 107 108
cycles N

Figure C.11: comparison of the crack progress curves from the undistorted and simplified models

Figure C.12 shows the maximum relative mode 1 strain energy release rates ∆G1 per cycle over the
crack length in comparison with mesh B and mesh B 1.5 mm. As it can be seen, the phenomenon
observed in the mesh B simulations also occurs when using an undistorted model and including
the unloaded regions.

0.1
∆G1 / 103 mJ2

0.08

mesh B 1.5 mm,


0.06 constant stabilization 10−6
long model,
constant stabilization 10−6
0.04
mesh B,
no stabilization
0.02
0 5 10 15 20 25 30 35 40 45 50
crack length a / mm

Figure C.12: comparison of ∆G1,max over the center crack length

The marginal amount of stabilization does not affect the results at all since ALLSD/ALLIE is far
below 1ppm.
Figure C.13 depicts the comparison between the crack progress curves at the individual node sets.
For better comparison with the 2D benchmark case, the starting point was shifted to the benchmark
Appendix 88

curve by adding 4 · 104 cycles. It can be seen that the same result as in figure 7.15 is obtained.

50

40
crack length a / mm

30

edge 1
20 center
edge 2
benchmark curve
10

0
100 101 102 103 104 105 106 107 108
cycles N

Figure C.13: crack progress curves for the undistorted, long mesh and shifted initial crack position
Appendix 89

Appendix D Influence of direct cyclic control parameters and variable Fourier


terms
Appendix D.1 Setting of the simulations
To test the influence of increased direct cyclic control parameters and variable Fourier terms instead
of fixed 50, two simulations with these changes were performed. Therefore, the long, undistorted
model from Appendix C was chosen as reference. In the case of direct cyclic control parameters,
the variables CUnα and CU0α were increased from 1 · 10−3 to 1 · 10−2 , whereas in the case of variable
Fourier terms, the maximum was kept at 50 Fourier terms but the initial number of Fourier terms
was set to 1 and the increment was set to 7 in addition to setting CUnα and CU0α to 5 · 10−3 , which
is the suggested default value in the Abaqus manual [18].

Appendix D.2 Results and discussion


Figure D.14 shows the crack progress curves of the reference and the modified cases. The simulation
with a large increase of the direct cyclic control parameters shows more clunky crack propagation
in the region of fast crack growth. It also deviates from the reference case there.
Using variable Fourier terms with the default proposed direct cyclic control parameters for CUnα
and CU0α did not impact the results at all.
The calculation times were on the same level and hence, no further investigation was done.

70

60
crack length a / mm

50

40

30 reference
(CUnα = CU0α = 1 · 10−3 )
20 CUnα = CU0α = 1 · 10−2
modified Fourier terms,
10 CUnα = CU0α = 5 · 10−3

0
100 101 102 103 104 105 106 107 108
cycles N

Figure D.14: comparison of the crack progress curves with increased DC control parameters and variable
Fourier terms
Appendix 90

Appendix E Finite element meshes

Figure E.15: mesh A for continuum shell elements

Figure E.16: mesh B 1.5 mm for solid elements


Appendix 91

Figure E.17: mesh B 2 mm for solid elements


Appendix 92

Appendix F Example input file for an interlaminar fatigue simulation using 3D


elements, VCCT and the Abaqus Direct Cyclic approach
In the following an input file of a simulation using mesh B of the ENF model is inserted. Thereby
one asterisk marks an Abaqus keyword while two asterisks start a commentary line. The text in
red was inserted by the author for additional description.

** PARTS
**
*Part, name=lower_stack
... nodes, elements, node and element sets and orientations: left out for the reason
of clarity
**
*Part, name=upper_stack
... nodes, elements, node and element sets and orientations: left out for the reason
of clarity
**
** ASSEMBLY
... part instances, node and element sets and contact surface definitions: left out
for the reason of clarity
**
** MATERIALS
**
*Material, name=cfrp
*Elastic, type=ENGINEERING CONSTANTS
161000.,11380.,11380., 0.32, 0.32, 0.45, 5200., 5200.
3900.,
**
** INTERACTION PROPERTIES
**
*Surface Interaction, name=vcct-contact
1.,
**
** BOUNDARY CONDITIONS
**
** Name: support_1 Type: Displacement/Rotation
*Boundary
line_support_1, 2, 2
line_support_1, 3, 3
** Name: support_2 Type: Displacement/Rotation
*Boundary
line_support_2, 1, 1
line_support_2, 2, 2
line_support_2, 3, 3
**
** INTERACTIONS
**
** Interaction: vcct
*Contact Pair, interaction=vcct-contact, adjust=bonded_nodes
surf_lowerstack, surf_upperstack
Appendix 93

*Initial Conditions, type=CONTACT


surf_lowerstack, surf_upperstack, bonded_nodes
** ––––––––––––––––––––––––––––––––
**
** STEP: static preload
**
*Step, name="static preload", nlgeom=YES, inc=10000
*Static
0.001, 0.2, 1e-25, 0.02
**
*controls, type=vcct linear scaling
0.9
*controls, parameter=time incrementation
„„„,50
**
** BOUNDARY CONDITIONS
**
** Name: displacement Type: Displacement/Rotation
*Boundary
line_displacement, 3, 3, -0.001
**
** INTERACTIONS
**
** Contact Controls for Interaction: vcct
*Contact Controls, master=surf_upperstack, slave=surf_lowerstack, reset
*Contact Controls, master=surf_upperstack, slave=surf_lowerstack, stabilize=1.
**
** OUTPUT REQUESTS
**
*Restart, write, frequency=0
**
** FIELD OUTPUT: F-Output-1
**
*Output, field
*Node Output
U,
*Contact Output
BDSTAT, CSTRESS, DBT, ENRRT, OPENBC
*Output, history, frequency=0
*Debond, slave=surf_lowerstack, master=surf_upperstack, debonding force=STEP, frequency=1
*Fracture Criterion, type=VCCT, mixed mode behavior=BK, tolerance=0.001
0.208, 0.78, 0.78, 2.5713
*End Step
** ––––––––––––––––––––––––––––––––
**
** STEP: fatigue
**
*Step, name=fatigue, inc=1000000
*Direct Cyclic, fatigue
Appendix 94

0.001, 0.2, , , 50, 50, 50, 1000


3, 6, 10000,
**
** BOUNDARY CONDITIONS
**
** Name: displacement Type: Displacement/Rotation
*Boundary, op=NEW, amplitude="R=0,1"
line_displacement, 3, 3, -1.
** Name: support_1 Type: Displacement/Rotation
*Boundary, op=NEW
line_support_1, 2, 2
line_support_1, 3, 3
** Name: support_2 Type: Displacement/Rotation
*Boundary, op=NEW
line_support_2, 1, 1
line_support_2, 2, 2
line_support_2, 3, 3
**
** CONTROLS
**
*Controls, reset
*Controls, type=direct cyclic
, 100., 0.001, 100., 0.001
**
** INTERACTIONS
**
** Contact Controls for Interaction: vcct
*Contact Controls, master=surf_upperstack, slave=surf_lowerstack, reset
*Contact Controls, master=surf_upperstack, slave=surf_lowerstack, stabilize=1.
**
** OUTPUT REQUESTS
**
*Restart, write, frequency=0
**
** FIELD OUTPUT: F-Output-1
*Output, field, frequency=1
*Node Output
U,
*Contact Output
BDSTAT, CSTRESS, DBT, ENRRT, OPENBC
**
** HISTORY OUTPUT: H-Output-1
*Output, history, variable=PRESELECT
** *Debond, slave=surf_lowerstack, master=surf_upperstack, debonding force=STEP, frequency=1
*FRACTURE CRITERION, TYPE=fatigue, MIXED MODE BEHAVIOR=BK, tolerance=0.001, viscosity=0
0.213,-6.329,0.33185,5.5519,0.102,0.9,0.21,0.78,
0.78,2.5713
*End Step

You might also like