You are on page 1of 203

MODELLING OF COMPOSITES

WITH DIRECT FE2 – A MULTI-LEVEL FEM

KARTHIKAYEN RAJU
(B.Tech, NIT-Trichy)

A THESIS SUBMITTED
FOR THE DEGREE OF DOCTOR OF PHILOSOPHY
DEPARTMENT OF MECHANICAL ENGINEERING
NATIONAL UNIVERSITY OF SINGAPORE

2019

Supervisors:
Associate Professor Lee Heow Pueh
Associate Professor Tan Beng Chye, Vincent

Examiners:
Dr Ong Eng Teo
Associate Professor Lim Kian Meng
Professor Nagashima Toshio, Sophia University
DECLARATION

I hereby declare that the thesis is my original work and it has been written

by me in its entirety. I have duly acknowledged all the sources of

information which have been used in the thesis.

This thesis has also not been submitted for any degree in any university
previously.

______________________________

Karthikayen Raju

i
ACKNOWLEDGEMENTS

I firstly thank my parents and relatives for all the sacrifices they made for me

and placing me ahead of everything else in their lives. I thank my supervisors

Associate Professors Vincent Tan and Lee Heow Pueh for the freedom,

encouragement, motivation and advice, which helped me traverse this journey.

I am grateful to Long Bin and Tse Kwong Ming for guiding me into this

frightening yet curiosity quenching high pressure research life. I thank Low

Chee Wah, Cheng Kok Seng, Amy Chee Sui Cheng, Teo Lay Tin Sharen and

Priscilla Lee Siow Har for their administrative and technical assistance in the

laboratory. I would like to thank Antoni, Nigel and Yingxi for introducing and

teaching me about material characterization tests.

I thank Balaji, Malar, Shahrokh, Jan, Ralf, Abhishek, Bob, Deepan, Venky,

Kyrin, Prashanth, Prachee and Sanjay for their company and heart-warming

conversations in the last four years. Special thanks to Vignesh Kannan for

persisting me to pursue graduate education without whom this document would

have not existed. I finally thank Vasantha without whom this thesis would not

have been completed or presented. I would like to thank all those who have

contributed to this work, but whom I have forgotten to mention in this

acknowledgement.

Lastly, I am extremely grateful to my supervisor Associate Professor Vincent

Tan again for putting up with my idiosyncrasies, trusting that I would work

better and guiding me through with this research work.

ii
TABLE OF CONTENTS

Declaration i
Acknowledgements ii
Table of Contents iii
Abstract vii
List of symbols ix
List of Figures xi
List of Tables xviii
Nomenclature xx
Introduction 1

Composite materials 1

Modelling of composites 2

Microscale models 3

Mesoscale (Ply-based) models 4

Macroscale models 6

Multiscale methods 7

Review of FE2 methods 10

Motivation and objectives 15

Thesis organization 17
Direct FE2 – Theory, implementation and 2D validation 18

Introduction 18

Direct FE2 theory 19

Getting the macroscale quantities from scaled microscale


computations 24

Implementation techniques 29

Boundary conditions at the microscale 31

Direct FE2 implementation in AbaqusTM 35

Setting up the Direct FE2 model domain 36

Microscale boundary conditions: applying the MPCs 37


iii
Solution procedure in AbaqusTM 38

Single Direct FE2 macroscale element models 38

Comparing with single FE element for a homogenous


material 39

Heterogeneous single macroscale element models 42

Conclusions 51
Application of Direct FE2 to 2D beams and 3D composite
laminates 53

Effect of macroscale mesh 53

FE2 macroscale mesh convergence study with a small


cantilever beam with geometric non-linearity 55

Effect of beam size on macroscale mesh requirements 57

Demonstration with different matrix constitutive models 59

Linear elastic matrix model with geometric non-linearity 60

Higher order interpolation function 63

Increasing the number macroscale integration points 67

Non-linear plastic matrix model with geometric non-linearity


68

Viscoelastic matrix model 71

Effects of microscale RVE: fibre arrangement and microscale


mesh size 73

Modelling the fibre matrix interface using cohesive elements 75

Examples with irregular elements: benchmarking irregular


elements with full FE composite material model 79

Modelling laminates in 3D 80

Modelling methodology 80

Benchmarking a quasi-isotropic FE2 laminate model with


CLT 82

Comparing the homogenized elastic properties of a single


3D macroscale element with a single RVE 86

iv
Modelling a tube structure from constituent bottom-up with
Direct FE2 87

Summary 91
Experiments on composite structures: Glass Reinforced
Epoxy (GRE) pipes under tension, torsion and bending loads 93

Pipe specifications and material details 93

A brief review of work carried out in the past 94

Experiments 98

Hexapod test set-up 98

Specimen preparation and mounting 99

Loading details 101

Observations 103

Tensile loading of the pipe specimens 104

Torsion 108

Bending 110

Summary 114
Direct FE2 analysis of composite structures – simulation of
GRE composite pipe experiments 116

Modelling of the composite pipes 116

Constitutive modelling of fibre and matrix 118

Tension 120

Torsion 122

Modelling torsion with microbuckling of fibres 126

Bending 128

Using cohesive elements along with 3D Direct FE2 model 132

Discussion 133

Summary 137

v
Conclusions and Future Work 139

Conclusions 139

Future directions 141

Future possibilities for Direct FE2 142

Future work for modelling composite structures with Direct


FE2 142
Appendix A: Applying microscale boundary conditions in Direct FE2 144
Appendix B: Straight edge boundary constraints 147
Appendix C: Periodic Boundary Conditions (PBC) 148
Appendix D: Setting up a Direct FE2 Model in ABAQUS 150
Appendix E: Defining the scale transition relationship (L matrix) through
MPCs 162
Appendix F: Coupling Direct FE2 with Full FE regions 166
Bibliography 168
List of Publications 181

vi
ABSTRACT

Many engineering structures are made of heterogeneous materials such as concrete,

fibre reinforced composites, polymer nano-composites, etc. Modelling of such

heterogeneous structures from the bottom-up constituents is the driving motivation for

developing homogenization and multi-scale methods to predict its behaviour. These

methods may be classified as concurrent and sequential. Sequential or hierarchical

methods loses information regarding the interaction between various constituent phases

when moving from constituent level to the homogenized macro level. A

computationally attractive method would be needed to model structures with their

heterogeneous micro-structures concurrently. This work describes a few steps taken

towards this direction by employing a FE2 based method.

FE2 is a concurrent multi-scale method, which performs Finite Element (FE)

calculations at two scales – the macroscale (homogenized material structural model)

and microscale (that specifies the heterogeneous micro-structure details). A FE analysis

of a Representative Volume Element (RVE) microstructure is carried out for every

macro finite element integration point calculation. FE2 is traditionally implemented in

such a way that at least two finite-element simulations are run concurrently - one for

the macroscale and one for the microscale.

This work describes the development of a new FE2 method called Direct FE2. The

Direct FE2 method requires only a single finite element simulation for both micro and

macroscales, which makes it computationally attractive. A Direct FE2 model requires

only a small fraction of the computational time and memory required by a brute force

fully meshed FE model for the same mesh size. This is achieved by implementing the

micro-macro-transition with interpolation (shape) functions using Multi–point

constraints (MPCs). Its ease of implementation along with material libraries and other

vii
modelling features available in any commercial FE software that supports MPCs is an

added advantage.

The Direct FE2 model is validated with a 2D composite cantilever beam and

demonstrated to be equivalent to brute force full FE models. The Direct FE2

implementation in 3D, allows the modelling of composite laminates, by taking into

account the behaviour of the constituents (comprising the fibre, matrix, interphase of

the lamina) along with the inter-laminar regions in the same model concurrently. The

methodology of 3D laminate modelling is validated by showing that results obtained

from the model without debonding and delamination reduces to the Classical Laminate

Theory (CLT) analysis, when fibre and matrix are assigned homogenized ply properties

(the same inputs as CLT).

The response of continuous glass fibre reinforced polymer composite pipes to various

loads--tension, bending and torsion--are performed on a Stewart platform-based test

ring to gain exposure to test(experiment)design and hands-on experience in specialized

testing of composite structures. 3D Direct FE2 models of the composite pipe mentioned

above are developed. The developed 3D composite model is employed for the Direct

FE2 simulation of the composite pipe experiments. The predictions are contrasted with

experimental measurements. The experimental data was a useful reference to determine

the capabilities and gaps in applying Direct FE2 in 3D for the analyses of actual

structures.

viii
LIST OF SYMBOLS

𝑆𝑢𝑏𝑠𝑐𝑟𝑖𝑝𝑡 𝑖, 𝑗 Co-ordinate directions

𝑆𝑢𝑏𝑠𝑐𝑟𝑖𝑝𝑡 𝐼 Macroscale finite element node number

𝑆𝑢𝑏𝑠𝑐𝑟𝑖𝑝𝑡 𝛼 Integration/Gauss point

𝑏 Body force (N/m3)

𝛿𝑤 Weight used in weighted residual methods

𝑢 Displacement (macroscale)

𝑢̃ Displacement (microscale)

e Macroscale element

𝐟 Vector of nodal forces

̃
𝒅 Vector of microscale nodal displacements

̃∗
𝑲 Overall stiffness matrix

𝑳 Scale transition relationship

𝑊𝑖𝑛𝑡 Internal work at macroscale

̃𝑖𝑛𝑡
𝑊 Internal work at microscale

𝑊𝑒𝑥𝑡 External work

𝜎 Stress (macroscale)

𝜎̃ Stress (microscale)

N Shape function

V Volume of computational domain

S Surface (or boundary of V)

𝑆𝑡 Traction boundary (boundary where traction is


specified)

t Traction (N/m2)

w Weight for Gaussian quadrature integration/weighted


residual method

ix
J 𝑥,𝜉 𝑦,𝜉
Determinant of Jacobian matrix,[𝑥
,𝜂 𝑦,𝜂 ]

V𝑒 Volume of macroscale element

𝑉𝛼 Volume of the RVE associated with the macro scale


element Gauss(integration) point α

𝑤
̅𝛼 Scaling factor

𝐸0 Instantaneous modulus of the matrix

𝜏𝑖 Relaxation time

ν Poisson’s ratio

τ Traction vector

Δ Separation vector

K Stiffness coefficient

c Cohesion of the material

𝜀𝑖𝑗 Strain in i,j direction

M Bending moment

P Tensile force

𝐹𝑥 Shear force

γ Shear strain

x
LIST OF FIGURES

Figure 1.1 Virtual testing of composites: (a) Microscale: individual


undamaged ply level made of individual material constituents;
(b) Mesoscale: laminate level made of individual homogenized
plies; (c) Macroscale: component level made of homogenized
laminates ........................................................................................ 2

Figure1.2 Parallelization of the macro and micro calculations in


ABAQUS/Standard for non-linear case (reproduced with
permission from [68]) .................................................................. 11

Figure 1.3 A typical FE2 implementation reconstructed from Smit et al.[65]


where Fmacro is the macroscopic deformation gradient, ui is the
nodal displacement of RVE (microscopic) nodes, y0i are the
positions of RVE nodes in the initial undeformed reference
configuration ................................................................................ 14

Figure 2.1 Microscale gauss point RVEs in (a) regular and (b) irregularly
shaped Direct FE2 macroscale elements ...................................... 24

Figure 2.2 Solution algorithms for (a) usual FE simulation, (b) classical FE2
and (c) Direct FE2. f ̃ is the vector of microscale nodal forces .... 28

Figure 2.3 RVE at macroscale element integration point. ............................... 33

Figure 2.4 A FE2 macro scale-element: the dotted lines marks the boundary of
the macroscale element domain defined by the 4 nodes of the
macro-element defined as Reference point); The macroscale nodal
displacements are ui, vi; i=1,2,3,4; and the micro structure RVEs
at the gauss-points; MPCs at the edge nodes of the RVEs. ......... 36

Figure 2.5 Flowchart depicting the Direct FE2 method solved using AbaqusTM
...................................................................................................... 38

Figure 2.6 Deformed homogeneous single element models for a small


deformation. (Strain E11 refers to stress in the horizontal
direction). The deformed dimensions of the element are given
along the edges. The vertical reaction force in the fixed end for
full FE model: 154807.6 µN and FE2 model: 154957.4 μN. All the
dimensions are in µm. .................................................................. 40

Figure 2.7 Deformed homogeneous single element models for a large


deformation. The deformed dimensions of the element are given
along the edges. The vertical reaction force in the fixed end for
full FE model: 2.140774N and FE2 model: 2.143697N. All the
dimensions are in µm. .................................................................. 41
xi
Figure 2.8 (a) 10x10 Full FE model; (b) Equivalent FE2 model with 4 RVEs at
the gauss points and the macro scale nodes (modelled as reference
points in Abaqus). ........................................................................ 42

Figure 2.9 (a) Boundary conditions of the full FE model; (b) boundary
conditions in the FE2 model. ........................................................ 44

Figure 2.10 FE2 microscale boundary conditions: displacement boundary


conditions (equation ((2.20)) vs periodic boundary Conditions
(equations ((2.23) and (2.24)) for (a) a simple shear deformation
and (b) tensile deformation. ......................................................... 45

Figure 2.11 Deformed RVE shapes in: (a) the bulk of full FE model; (b) FE2
macroscale element with displacement boundary condition
(equation (2.20) and (c) FE2 macroscale element with periodic
boundary conditions (equations (2.23) and (2.24)). ..................... 46

Figure 2.12 (a)The force vs normalized displacment behaviiour of differently


sized macroscale element models (b) The difference between the
forces in full FE model and FE2 model with macroscale element
size. .............................................................................................. 48

Figure 2.13 (a) Full FE model edge effect due to straight exterior edge
constraint (b) deformed RVE in FE2 macroscale element with
periodic boundary conditions. ...................................................... 49

Figure 2.14 (a) Full FE (10x10 RVEs)Model with PBCs in which no edge
effect is visible; (b) deformed FE2 RVE used in the presented
analysis (c) a deformed RVE from a 10x10 FE2 model with fibre
centered RVEs to show the effect of the RVE mesh on the stress
contours. ....................................................................................... 50

Figure 2.15 Comparing errors between full FE with PBCs and FE2 and full FE
with straight edge constraints and FE2. ........................................ 51

Figure 3.1 Cantilever beam (a) FE2 model with 2x20 macroscale elements with
the macro-element boundaries (b) full FE model. ....................... 54

Figure 3.2 RVE used in the simulations. ......................................................... 54

Figure 3.3 Stress contours of region closest to the fixed end of the beam (a)
full FE, (b) 1x10 FE2 linear macroscale elements, (c) 2x20 FE2
linear macroscale elements, (d) 4x40 FE2 linear macroscale
elements, (e) 8x80 FE2 linear macroscale elements; (f) Force-
displacement behaviour for linear elastic epoxy and fibre with
geometric non-linearity cantilever beam of various linear
interpolated FE2 models and full FE model. ................................ 56
xii
Figure 3.4 Relative size of the cantilever beam models used in this work (a)
Large beam on top with size of -2000RVEs(in Figure 3.2) x 200
RVEs with a 8x80 macroscale element mesh and (b) Short beam
on the bottom with size of 200RVEs x 20 RVEs with a 2x20
macroscale element mesh............................................................. 57

Figure 3.5 Normalized force vs normalized displacement: comparison of large


and short beam behaviour. ........................................................... 59

Figure 3.6 (a) Cantilever beam with a 8x80 macroscale element mesh with an
integration point region of a highlighted macroscale element
represented in Direct FE2 (thickness scaled RVE) and full FE
(156.25 RVEs); (b) A RVE at FE2 macroscale element integration
point; (c) A full FE region equivalent to single integration point
RVE in 8x80 Direct FE2 model ( 𝑤𝛼 = 156.25) as per
equations (2.13, 2.19) for this macroscale discretization). .......... 60

Figure 3.7 Homogeneous beam: Validating the Direct FE2 method. ............... 61

Figure 3.8 Periodic vs Linear Displacement Boundary Conditions (BCs):


Homogeneous and composite materials (FE2 models contain 8x80
macroscale elements). .................................................................. 63

Figure 3.9 Effect of macroscale element interpolation on Force-Displacement:


quadratic interpolated SE converges faster than linear SE (FE2
models contain 8x80 macroscale elements). ................................ 64

Figure 3.10 Stress contours in (a) linear FE2; (b) quadratic FE2; (c) Full FE
reference at free end deflection of 0.18*beam length. ................. 65

Figure 3.11 Equivalent points near the coordinate system for displacement
comparison in (a) Full FE reference; (b) Direct FE2 model (c)
vertical displacement comparison of Direct FE2 model with full
FE model- linear interpolation vs quadratic interpolation. .......... 66

Figure 3.12 Composite cantilever beam force displacement showing


consistency and convergence as more Gauss points are used with
macroscale quadratic elements..................................................... 67

Figure 3.13 Load-displacement plots for large deformation of elastic-plastic


composite beam............................................................................ 69

Figure 3.14 Equivalent plastic strain contours of RVEs closest to top left of
beam for (a) linear FE2, (b) quadratic FE2 and (c) full FE; (d)
Excessive deformation of elements in full FE simulation in grey at
top left edge of beam. ................................................................... 70

xiii
Figure 3.15 (a) Load-displacement plots of viscoelastic composite beam.
Residual stresses (MPa) at the end of loading cycle within the
RVEs of Direct FE2 models with (b) linear and (c) quadratic
macroscale elements and (d) full FE model. ................................ 72

Figure 3.16 (a) Effect of RVE architecture: FE2 models contain 8x80
macroscale elements on force-displacement behaviour; Axial
stress contours at deflection of one-third beam length in models
with: (b) non-symmetric RVE with fine mesh; (c) Symmetric
RVE with coarse mesh; (d) symmetric RVE with intermediate
mesh; (e) symmetric RVE with fine mesh. .................................. 74

Figure 3.17 Comparison of Direct FE2 (left) and full FE (right) cantilever
beam axial stress contours near the fixed end top left corner
with(deformation magnified 100 times) (a) 2c debond strength at
18 μm free end deflection; (b) c debond strength at 6 µm free end
deflection before debonding; (c) c debond strength after
debonding at 9µm free end deflection and (d) Force deflection of
a beam (Figure 3.4(b)) with fibre-matrix debonding. .................. 77

Figure 3.18 (a) Full FE model reference (b) FE2 model with quadratic
interpolated irregular 2D macroscale elements with PBCs on the
RVE and tensile displacement applied on the macroscale nodes. 80

Figure 3.19 RVEs placed as per ply orientation-RVE white region represent
the fibres and green region represent the matrix. ......................... 81

Figure 3.20 Conventional modelling of laminates from constituents involves


laminate specific RVEs for multi directional laminates for (a)
0°/90° laminate and (b) [+55°/-55°] laminate. ............................. 81

Figure 3.21 RVE used for modelling the quasi-isotropic laminate. RVE white
region represent the fibres and green region represent the matrix.
A coarse mesh is used to reduce computation time. .................... 83

Figure 3.22 The local stresses in each ply when the laminate is loaded in the x
–direction, εxx=0.02 . Each RVE’s fibre and matrix regions are
given ply properties in Table 3 7. The laminate global coordinate
system is shown in bottom left hand corner. ................................ 84

Figure 3.23 (a) Shear stresses in the FE2 macroscale element when sheared in
the YZ plane (plane of this page and the fibre direction is x
pointing out of the plane of the page); (b) Shear stresses in the
single RVE when sheared in the YZ plane; (c) RVE mesh used for
the calculations. ............................................................................ 87

xiv
Figure 3.24 Scaling in 3D Direct FE2 pipe model with (a) cubic RVE in Figure
3.21 with fibre diameter of 15 µm and (b) cubic RVE volume
scaled so that volume of the RVE is same as macroscale
integration point volume with scaled fibre diameter of 1260 µm to
satisfy equation (2.19). ................................................................. 88

Figure 3.25 RVE used in the elastic pipe simulations with (a) coarse mesh; (b)
fine mesh. ..................................................................................... 89

Figure 3.26 (a) Boundary conditions for the pipe section twisting; (b) Elastic
behaviour of pipe structure: Comparing full FE model and FE2
model. ........................................................................................... 90

Figure 4.1 Composite pipe coupon section: (a) inner diametrical surface of the
pipe- +55°/-55° filament windings are clearly visible; (b) through
thickness section view which shows the ply interfaces. .............. 94

Figure 4.2 Hexapod test rig at Hamburg University of Technology, Germany


with fixtures and test specimen. ................................................... 98

Figure 4.3 (a) Additional pins used for transmission of torque and (b)
Neoprene rubber sheet cross-section configuration used for
clamping (without the pins) (c) Pin hole failure. ....................... 100

Figure 4.4 Strain gauge configuration A used for tension, bending and strain
gauge configuration B used for torsion. ..................................... 100

Figure 4.5 The different loading conditions applied on the pipe: Py is the axial
tensile force, Ty is the torsional moment, Mz is the bending
moment and Fx is the transverse shear force in the pipe. ........... 102

Figure 4.6 In-house tensile tests setup with Shimadzu machine. (A) Pipe
specimen inserted into the mandrel assembly; (B) Collars clamped
the pipe specimen with rubber sheet; (C) Entire test setup onto the
Shimazhu machine; (D) Protective shield to prevent any flying
chips and digital video recorder to record the entire test. .......... 105

Figure 4.7 DIC image showing the uniformity of the tensile load throughout
the gauge length of the pipe (a) before peak failure load; (b) after
peak failure load with the formation of the strain localization
bands along the filament winding direction. .............................. 106

Figure 4.8 (a) Stress strain behaviour of composite pipe in tensile load
(b) short pipe section before failure; (c) and (d) short pipe failed
specimens; (e) and (f) failed region in long pipe very similar to
the damage pattern observed in the short pipe test at a lower
loading rate. ................................................................................ 107
xv
Figure 4.9 Strain rosette with strain gauge P, Q and R; θ=45°; ϕ=45°. ......... 108

Figure 4.10 Torsional shear stress vs shear strain of the pipes. ..................... 109

Figure 4.11 Torsion pipe failure modes: external visible failure dependent on
the relative direction between the torque and state of stress in the
outermost ply; pipes in (a) test 6 were twisted in the –z direction
in which the outermost ply was in tension; (b) failed specimen in
test 7 twisted in +z direction in which outermost ply was in
compression (c) local ply buckling failure of the inner plies
observed inside the torqued (-z) pipe. ........................................ 110

Figure 4.12 Axial strains recorded along the gauge length of the tensile side of
pipe during the bending test 4 (in Tables 4.4 and 4.5). .............. 111

Figure 4.13 Bending stress vs strain (longitudinal direction). ....................... 112

Figure 4.14 Failed regions upon bending by failure in test 4 on (a) tensile side,
(b) and (c) compressive side; failed regions in test 5 (d) matrix
compression on the compressive side and (e) delamination upon
ultimate failure. .......................................................................... 113

Figure 5.1 FE models of pipe at: (a) macroscale level, (b) ply-level, (c)
constituent level, (d) Direct FE2 pipe model, (e) coarse mesh RVE
and (f) fine mesh RVE. .............................................................. 117

Figure 5.2 Boundary conditions for tension pipe test on the FE2 macroscale
elements. .................................................................................... 120

Figure 5.3 Tensile stress vs tensile strain: Comparing FE2 model response with
pipe tensile experiments. ............................................................ 121

Figure 5.4 Axial Stress in the fibres and matrix of the +55° and -55° plies at
(a) 0.0025 axial strain(25 µm axial displacement) (b) 0.01 axial
strain(100 µm axial displacement) and (b) equivalent plastic strain
contours in the fine meshed RVEs in tension at (c) 0.0025 axial
strain(25 µm axial displacement) and (d) 0.01 axial strain(100µm
axial displacement). ................................................................... 122

Figure 5.5 Boundary conditions for torsion pipe test on the FE2 macroscale
elements. .................................................................................... 123

Figure 5.6 Stress- strain behavior of pipe in torsion: Comparing FE2 model
simulation results with experiments. .......................................... 124

xvi
Figure 5.7 All contours are at macroscale shear strain of 0.01346. Plastic stress
contours when torqued in (a) -z direction and (b) +z direction. The
corresponding axial stresses in the (c) outermost ply fibres are in
tension when torqued in -z direction and the (d) outermost ply
(bottom left) fibres are in compression when torqued in the +z
direction. (e) Tensile stresses in the matrix and the plies with
fibres in compression, which causes the plastic strain observed in
Figure 5.7(a). .............................................................................. 125

Figure 5.8 Sinusoidal function used to model initial misalignment and the FE
model of the RVE used for the simulation to determine the
microbuckling strain. ................................................................. 126

Figure 5.9 Torsion with fibres modelled to fail at a preassigned microbuckling


strain rather than the compressive strain. ................................... 127

Figure 5.10 Boundary Conditions for bending. ............................................. 128

Figure 5.11 Bending longitudinal stress- strain behaviour: Comparing FE2


models with experiments. .......................................................... 129

Figure 5.12 Stress contours of the tensile side RVEs in bending at a bending
axial stress of 39.8MPa (point of divergence of FE2 model
behaviour). The top right RVE in each image is the outermost ply.
The left column of contours-(a), (c), (e) and (g)-has zero
transverse shear stress and the right column-(b), (d), (f) and (h)-
has 0.1 τyz : σzz as reported in bending Test no. 5 in Chapter
4.Plastic strain contours in (a) and (b);Shear stress in xy in (c) and
(d); Shear stress in yz in (e) and (f); normal(longitudinal) stress in
axial direction in (g) and (h). ..................................................... 130

Figure 5.13 Compression: FE2 vs experiments: UC1,UC2,UC3 and UC4 refer


to compression coupon experiments [151]. ............................... 131

Figure 5.14 Cohesive elements (COH3D8) used with the Direct FE2 model:
Cohesive element Damage variable SDEG (Overall scalar
stiffness degradation) contours are shown. ................................ 132

Figure 5.15 Effect of model length for (a) pipe tension; (b) pipe torsion. ..... 134

Figure 5.16 Effect of model length for (a) pipe bending without shear force;
(b) pipe bending with shear force. ............................................. 136

xvii
LIST OF TABLES

Table 2.1 Material properties ........................................................................... 39

Table 2.2 Boundary displacements given to various model sizes; x and y are
the displacements shown in Figure 2.9 ........................................ 43

Table 3.1 Material properties. .......................................................................... 54

Table 3.2 Size of the beam model. ................................................................... 58

Table 3.3 Material elastic properties ................................................................ 68

Table 3.4 Matrix plastic properties [115] ........................................................ 68

Table 3.5 Material properties ........................................................................... 71

Table 3.6 Traction separation law parameters ................................................. 76

Table 3.7 Elastic properties of the carbon fibre ply – T300-BLS914C ........... 83

Table 3.8 Comparison of FE2 model and CLT solutions for global laminate
normal strain εxx=0.02 .................................................................. 84

Table 3.9 Comparing FE2 calculated stresses with CLT predictions for global
laminate normal strain εyy=0.02.................................................... 85

Table 3.10 Comparing FE2 calculated stresses with CLT predictions for global
laminate normal strain εxy=0.02.................................................... 85

Table 3.11 Material properties of the fibre and matrix used for 3D single
element and pipe structure (section 3.6.4) ................................... 86

Table 3.12 Homogenized elastic properties of the RVEs and FE2 macroscale
elements ....................................................................................... 86

Table 4.1 GRE pipe specifications................................................................... 94

Table 4.2 Past research in GRE pipe failure analysis ...................................... 95

Table 4.3 Loading specifications for Hexapod test facility ............................. 99

Table 4.4 Experimental test details ................................................................ 103


xviii
Table 4.5 Failure strengths of the pipe under various loading conditions ..... 103

Table 5.1 Material elastic properties [145] .................................................... 118

xix
NOMENCLATURE

FEM Finite Element Method

PBC Periodic Boundary Conditions

GRE Glass Reinforced Epoxy

MPC Multi-Point Constraint

CLT Classical Laminate Theory

DOF Degree of Freedom

DIC Digital Image Correlation

FPZ Fracture Process Zone

VARTM Vacuum Assisted Resin Transfer Moulding

CZM Cohesive Zone Models

RVE Representative Volume Element

xx
Introduction

This chapter briefly reviews composite materials, techniques to model

composites and multi-scale methods, particularly focusing on the FE2 method

to set up the motivation and objectives for this thesis. The chapter concludes

with the outline of the thesis.

Composite materials

Many engineering structures are made of heterogeneous materials such as

concrete, fibre reinforced composites, polymer nano-composites. The goal of

making composites is to get the best properties of each constituent and suppress

their inferior qualities in the resulting heterogeneous material. Composites

encompass a broad range of materials varying from particle reinforced

polymers, concrete, short fibre reinforced polymer, and continuous/long fibre

reinforced polymers. In this work we have focused more on fibre reinforced

composites because of their application as structural components, such as wind

turbine blades [1–3], automotive body parts [4] and aeroplane fuselage [5,6],

which demand lightweight materials with high rigidity and strength. Fibre

reinforced polymer composites are prepared by various methods such as

filament winding for tubes, hand layup with prepreg , vacuum assisted resin

transfer moulding (VARTM) [7] and other methods.

A fibre reinforced composite typically comprises of thin fibres cured in a

polymer resin or matrix. Using thin fibres, the strength of the materials

reinforced with these fibres is increased due to the reduced possibility of

microstructural defects and influence of cracks in these thin fibres. Unlike alloys

1
which are strengthened with improvements in technology of grain size

reduction, precipitation hardening, the strength of fibre reinforced composite is

due to the natural potential property of the material [8], namely less defects and

high fibre strength. Therefore, glass fibres, carbon fibres and aramid fibres are

very widely used in composite materials for various engineering applications.

It is essential to understand the behaviour of these composites to design them

for a defined application. Modelling composites enable us to understand the

mechanical behaviour and also how it fails (the failure mechanism). The next

section discusses various composite modelling techniques.

Modelling of composites

Figure 1.1 Virtual testing of composites: (a) Microscale: individual


undamaged ply level made of individual material constituents;
(b) Mesoscale: laminate level made of individual homogenized plies;
(c) Macroscale: component level made of homogenized laminates

The most favoured way of modelling long fibre reinforced composites is

modelling laminates and the structures with individual plies (layers/laminae) or

the laminate as a whole depending on the resolution and availability of

computational resources. Around 10000 tests [9,10] are needed before putting

2
a composite structural component to service. This staggering experimental

requirement for component development has inspired a roadmap for the virtual

testing of these materials as shown in the Figure 1.1 (adapted from [11,12]).

It is important to know the nature of damage in composites and also how it

propagates to use them in engineering applications. The damage modes in

laminated composites can be primarily classified as intra-laminar damage

modes such as matrix cracking, fibre failure, fibre pull out and interlaminar

damage modes such as delamination. Modelling of composite non-linear

behaviour involves capturing these damage mechanisms with an appropriate

model. The laminate and component levels referred here correspond to Figure

1.1 (b) and (c), respectively. In this section some of the past research in

modelling composites at each of these levels are reviewed.

Microscale models

Microscale modelling (Figure 1.1(a)) refers to the modelling of the individual

constituents of the composite in an appropriate representative volume element

(RVE) in an individual ply. Generally, it is the first step of a hierarchical multi-

scale modelling methodology or a scale at which concurrent models obtain their

homogenized constitutive behaviour for their macroscale model. Microscale

modelling of composites have been carried out by numerous researchers [13–

16] with an objective to obtain a homogenized ply constitutive model.

Microscale is the scale at which individual material constituent damage such as

fibre fracture, matrix cracking or fibre debond initiates. At this level, since only

individual ply constituents are being modelled, interplay between damage

modes in various plies and interply damage, such as delamination, cannot be

3
captured from a laminate and in-situ behaviour of a ply [17] is also not captured

since only a single ply’s constituents are being modelled.

Microscale models are unable to predict damage evolution [18] even though

they have been very successful at predicting the elastic properties of undamaged

materials and also the onset of damage. This is because the unit cell for the

virgin material cannot be the same as the one assumed for predicting the

behaviour of the extensively damaged material. Smearing and homogenization

cannot be performed at the same scale for both the damaged and undamaged

material. In this work, only the onset of structural damage is shown to be

possible to predict and not the evolution because of this limitation, i.e. unit cells

are not used to model material behaviour but used to model and predict

structural behaviour up to the point of localization.

Mesoscale (Ply-based) models

At the next scale, mesoscale models, shown in Figure 1.1(b), are used to predict

the composite laminate behaviour at the individual ply level [19–26]. At this

level, all damage modes can be accounted for- both interlaminar and

intralaminar failure mechanisms with smeared crack material

degradation/damage models and cohesive zone models (CZM) respectively.

Modelling delamination is a vast field in itself. Yuan et al. [27] developed a dual

purpose damage model that can simultaneously model intraply and interply

damage as an alternative to the conventional practice of modelling ply failure

with continuum damage mechanics and delamination with cohesive elements.

Individual ply properties, such as ply elastic moduli, (E11, E22, E33, ν12, ν13, ν23,

G12, G13 and G23) and ply strengths (Xt, Xc, Yt, Yc, SL, St) are the input required

4
for ply-based models. These ply properties are obtained numerically or

experimentally. These homogenized plies are further used to model laminates

with inter-laminar regions. Numerically, constituent RVE models are

homogenized with constituents such as fibre, matrix and interphase into plies.

Ply properties may also be obtained from an underlying micromechanical

theory. Experimentally, these homogenized ply properties are obtained from

characterization tests on unidirectional laminates. It still remains to be

established whether experimental data obtained under uniaxial loading, can be

used to predict in-situ strength under multi-axial load states [28–35].

Since ply models do not have any geometric or constitutive information of the

individual constituents, damage modes in the ply model are activated or

accounted for implicitly based on failure criteria (like Hashins [29] or Tsai Wu

[36], which requires individual ply properties as input, with many issues as cited

previously). For example, if the stress in the 11 direction of the ply coordinate

system exceeds, its corresponding strength value Xt in tension, the stiffness in

11 direction of the ply is made zero or reduced to a negligible fraction of the

initial value- this damage mode is called fibre tensile failure mode in the ply.

These ply-based models assume that the ply behaviour is independent of

neighboring plies. This implies that the inelastic behaviour of the laminate is

not influenced by the interaction of damage mechanisms between neighbouring

plies. In the inelastic regime where damage initiation and evolution are strongly

affected by the degree of confinement provided by the nearby plies in the

laminate, it is known and accepted that the orientation of neighbouring plies

play a major role in the response of a lamina particularly. When placed in a

5
multidirectional laminate, generally plies exhibit different effective behaviour

[17]. The other limitation with ply-based models to model structures is the

computational cost involved because each and every layer is simulated, which

makes it impractical for simulation of real size structures.

In spite of high computational cost to model structures and assumption of

independent behaviour of the plies, this level is the most widely used level since

it is convenient to model both interply and intraply damage.

Macroscale models

Macroscale models shown in Figure 1.1(c) are the only practical method to

model large scale structures. These models are designed to simulate overall

response of the structure in terms of the laminate as a whole. The macroscale

models are typically not capable of predicting the details of the damage events

in the plies. Numerical efficiency and accurate structure response prediction are

the main objectives of using macroscale models. Since the response of a single

lamina in a stack is governed by its in-situ characteristics, only a model

constructed at the scale of block of layers has the potential to incorporate the

interaction of damage mechanisms between layers caused by the stacking

sequence.

The sublaminate based approach introduced by Williams et al. [37], uses the

sublaminate as the building block of the laminated composite structure. The

goal of the sub laminate approach is to predict in a smeared manner the essence

of the overall non-linear response of the laminated composite structure (eg,

stiffness, load carrying capacity, stability and post-peak response) due to

progressive damage caused by a given loading condition rather than capturing

6
the damage of each and every layer individually. Since the sublaminate is the

basic building block, each variant of the stacking sequence must be regarded as

a separate material for which the material properties need to be recalibrated.

In this section the problems involved in modelling of composites have been

identified. In the following section multi-scale methods for composites

modelling are reviewed.

Multiscale methods

With the increased use of heterogeneous materials like composites and alloys,

predicting the macro-structural behaviour from material microstructure using

multiscale methods would help to tune macroscopic structural design and

material microstructure design. A multiscale method consists of modelling the

behaviour of material/structure at different scales.

These simulations are required not just to understand the response but also for

studying the nature of failure or damage of the composites under various loads.

The formation and propagation of damage in heterogeneous materials like

composites occurs at multiple scales like fibre debond failure at the ply-

constituent level and delamination at the laminate level in laminated fibre

reinforced composites. The multiscale nature of damage and failure in these

materials has been a major driving force for developing multiscale models [18].

Multiscale model development is also motivated by virtual testing paradigms to

cut down on the composite component development cost and time.

Multiscale methods seek more physically derived bottom-up explanations for

inelastic mechanical behaviour [38–41] and their interactions with non-

7
mechanical phenomena [42] observed in heterogeneous material structures,

which can be provided only at the scale where dissipative (inelastic)

mechanisms operate in the material. This scale is generally a much lower length

scale than the structural length scale, referred to as the microscale (which is the

length scale of the material micro-structure).

Multiscale methods can be classified as sequential/hierarchical and concurrent

methods. Sequential or hierarchical methods loses information regarding the

interaction between various constituent phases when moving from constituent

level to the homogenized macro level. Concurrent methods contain information

regarding both the macro scale and micro scale in the same model and both

scales pass information to one another. The computational intensiveness of

concurrent models (expenses) arises as it is required to perform analysis of

models at two scales through a control code to handle the information transfer

between the two scales, which is a bottleneck to scale up modelling of large-

scale structures. The main advantage of such an approach is that it is possible to

take into account precisely the interaction between the different phases [43]

which are present in heterogeneous materials like composites.

Hierarchical models [44–49] are the first generation of multiscale models.

Homogenization allows the transfer of the predicted behaviour from a lower

scale (e.g. micromechanical RVE) to a higher scale (e.g. ply). While

hierarchical models are capable of predicting the elastic response of materials

with complex architectures at microscale, for damage or failure predictions,

they are faced with the issue of homogenization being invalidated once strain

localization sets in. As a result, hierarchical models cannot predict fracture

8
energy, size of fracture process zone (FPZ) and thus size effect objectively. The

characteristic length linked to the constitutive model at the macro scale cannot

be built up from lower scales in a hierarchical manner. Concurrent models avoid

this issue of invalidation of homogenization upon strain localization by having

both scales in a single model. In undamaged locations, macroscale models with

effective homogenized behaviour are employed while in damaged areas, the

model resolution is increased, and meso/micro features are explicitly taken into

account.

As one of the earlier concurrent models, Ghosh et al. proposed an adaptive

multi-level computational model [50] for prediction of damage due to matrix

cracking in composite lamina. It is based on their concurrent [51] modelling

strategy in which three levels of computation are carried out simultaneously -

(i) macro, (ii) macro/micro (RVE) and (iii) purely microscopic domain when

RVEs cease to exist. They have also adopted the method to model

microstructurally debonding materials and applied it to a double lap bonded

composite for demonstrating the model’s capability in handling large structural

problems [52]. Waas and co-workers have extended the concurrent modelling

framework to simulation of damage growth in 3D braided composites[53].

The shortcomings of the concurrent models are their immense computational

costs and explicit characterization of small-scale models. Concurrent models

due to their computational intensity and implementation difficulties are not

widely adopted in the industry.

Multiscale methods can also be classified as (i) mean field and (ii) full field

approaches. Mean field approaches (describing the macro-behaviour using

9
microstructure with the Mori-Tanaka method [54,55] or self-consistent scheme

[56–59]) are adequate for linear cases, but are inaccurate for modelling

scenarios involving non-linear constitutive behaviour. Moreover, mean field

approaches do not describe the local response and its interaction within the

material, but describes the overall macroscopic mechanical behaviour. This has

necessitated the development of full field multiscale approaches to model

heterogeneous material structures with non-linear two-scale analysis [60–63].

FE2 is one of the non-linear two scale methods that allows us to model the

macro-structural behaviour derived from the material micro-structure.

Review of FE2 methods

FE2 is a direct computational homogenization approach, which is material

independent and hence versatile [64]. Though Smit et al. presented the idea of

multi-level Finite element method [65], the term FE2 was first used by Feyel

et al. [63].

FE2 is a multi-level finite element method, which performs finite element (FE)

calculations at two scales-namely the macroscale (homogenized material

structural model) and microscale (that specifies the heterogeneous micro-

structure details). An FE analysis of a representative volume elements (RVEs)

microstructure is carried out for every macro finite element integration point

calculation. FE2 is traditionally implemented in such a way that two finite-

element simulations are run concurrently - one for the macroscale and one (or

many in the case parallelized computing as shown in Figure1.2) for the

microscale [43,63,65–68]. As the macroscale FE model contains multiple

integration points, there are multiple FE computations of microscale RVEs

10
corresponding to each macroscale integration point for every incremental. This

requires large computing memory and processing resources and a need to

parallelize these microscale finite element computation as shown in Figure1.2

[68] to ensure that the FE2 computation are performed over a reasonable

duration.

Figure1.2 Parallelization of the macro and micro calculations in


ABAQUS/Standard for non-linear case (reproduced with permission from
[68])

User defined subroutines to facilitate data interchange between the two scales

have also been employed in commercial software like Abaqus [67,68]. The FE

simulation of the microscale RVE may also be carried out beforehand offline

11
and the micro-stresses are looked up from a scalable calculated database [69]

based on the macroscale strain increment (or deformation gradient), although

this approach is highly dependent on the comprehensiveness of the database.

A comprehensive literature overview of the FE2 method has been given by

Schröder [70]. A brief description of the FE2 literature is given here.

The research work carried out in FE2 method can be broadly classified as work

done in development and implementation of the FE2 method [43,63,65–69],

application of the FE2 method for studying material behaviour/phenomena [38–

42,71,72], and for different macroscopic structural elements/components like

plates, shells and beams [73,74]. The FE2 method has been used to study

composite behaviour [38–41,75,76], thermo-mechanical behaviour [71,72],

magneto electro-mechanical behaviour [42], and interaction of microscopic and

macroscopic instabilities [60,77].

FE2 method development work itself can be further differentiated into (a) RVE

design research, which focusses on the effects of the RVE model parameters

like size, volume fraction, and inclusion/reinforcement architecture on the

macroscopic measures [78–82] and (b) scale transition research, which is the

exploration of how the measures in each scale interact and affect one and

another. Even though generally discussed together, work in scale transition can

be further distinguished between upscale transition literature [83–86][87]

(regarding the homogenizing of the microscale measures through various

homogenization techniques) and work in downscale transition (determining the

microscale boundary problem for the given macroscale deformation).

12
The upscale transition performed using the concept of homogenization theory,

involves volume averaging of the microscale quantities (stresses, strains) of the

unit cell to obtain the macroscopic quantities. A constitutive model has to be

defined (programmed) for the macro and microscale. This implies that it would

be required to develop a subroutine for the information interchange between

these scales [see 40–43] seen in Figure 1.3 [65], i.e., macroscopic constitutive

calculations through a suitable homogenization technique [83–87] of the local

microscale unit cells.

The downscaling problem has been widely discussed in literature, which

comprises of determining the boundary conditions to be applied at the

microscale RVEs associated with each integration point. Miehe [92]

investigated the three classical boundary conditions - (a) linear displacements,

(ii) constant tractions and (iii) periodic displacements and anti-periodic tractions

to be applied on deformation-driven microstructures. Nezamabadi [39] used

classical periodic conditions citing that they have the least drawbacks.

Kaczmarczyk et al. [93] have studied scale transition and enforcement of

generalized RVE boundary conditions for second order homogenization. The

boundary conditions are further discussed in detail by van der Sluis et al. [94],

Terada et al. [95], Kouznetsova et al. [96], Miehe and Bayreuther [97] and Perić

et al. [98].

An implementation of typical FE2 or any multiscale FE implementation

mentioned previously involves extensive programming and parallelization

techniques to run both the macro and micro FE calculations simultaneously.

Apart from the implementation difficulties, the computation time and resources

13
required to scale up with the problem domain size and model complexity for

real world application is high because of the multiple instances of FE

simulations needed to compute different scales and parallelizing several FE

simulations (equal to macro-scale integration points) of micro-scale RVEs

required for one macro-scale strain increment.

Figure 1.3 A typical FE2 implementation reconstructed from Smit et al.[65]


where Fmacro is the macroscopic deformation gradient, ui is the nodal
displacement of RVE (microscopic) nodes, y0i are the positions of RVE nodes
in the initial undeformed reference configuration

Since each material constitutive model requires different internal variables like

strain rate for rate-dependent constitutive model in addition to stress and strain

– a separate algorithm will be needed for different combinations of macro and

micro level constitutive models to handle the data transfer between the scales.

This requires considerable programming and constitutive modelling expertise.

Therefore, this leads to intensive computations compared to the sequential

method, but the cost is far less expensive than the full FE brute force approach.

It would be advantageous if we can reduce the number of FE simulations or

better combine the microscale and macroscale FE simulations so that FE2

14
simulation of structures become computationally attractive and can be made

more easily accessible to the computer aided engineering (CAE) community.

Motivation and objectives

A computationally attractive method would be needed to model structures with

their heterogeneous microstructures concurrently. Some researchers have tried

to accomplish modelling heterogeneous structures from bottom up [50–52] and

have successfully demonstrated it, but these approaches can be difficult to

implement and require expert level programming skills. Concurrent methods

having to deal with two scales (which means two separate models which

communicate with each other) is computationally intensive for large structures.

If we can remove one of the scales or deal with both scales by cutting down on

the communication between these scales, that would reduce the computational

burden. This work describes a few steps taken in this direction by employing a

FE2 based method.

The ply-based models and the macroscale models need properties that are

obtained from experiments on unidirectional laminate or specific laminate or

from sub-scale models like a micromechanical RVE for ply properties. As

mentioned earlier, it still remains to be established whether experimental data

obtained under uniaxial loading alone, can be used to predict in situ strength

under multi-axial load states [28–35]. Ply properties from angled ply laminates

are preferred because the results are less sensitive to specimen preparation and

testing skills and eventually lower data scatter as has been reported in strength

measurements [99–101].

15
More experiments to obtain these ply property inputs are required - whether

unidirectional laminate properties are sufficient, or they have to be back

calculated from angle plied laminates. This situation can be avoided if we can

model laminates from the constituents directly and ply properties emerge

naturally from the model. This is the motivation behind this work. This work

demonstrates a new method called Direct FE2, which provides a means to short

circuit from ply microstructure in Figure 1.1(a) to component in Figure 1.1(c)

bypassing the homogenized plies stacked as the laminate level in Figure 1.1(b),

while not compromising on the simulated behaviour of the composite.

The Direct FE2 modelling method presented in this dissertation can be used to

model any heterogeneous material once a unit cell has been established for the

material of interest.

The following objectives are laid out for this dissertation:

1. To implement a computationally attractive concurrent method – Direct FE2

2. To benchmark the Direct FE2 method with brute force full FE models

3. To develop and validate a 3D Direct FE2 Laminate modelling methodology


that takes constituent level inputs and predicts macroscale behaviour.

4. Demonstrate the method’s capability and shortcomings in modelling


structures bottom-up bypassing the ply level (not giving any ply level input)
until strain/damage localization and get useful and worthy insights from these
models. This is carried out by simulating the behavior of the composite pipe
subjected to tension, torsion and bending (and compare with experiments
carried out on a Stewart platform-based setup) with the above method.

16
Thesis organization

A review of the literature in FE2 has been presented in this chapter. The theory,

implementation and validation of the Direct FE2 method is presented in

Chapter 2. It discusses how the macroscale computations can be directly

obtained from the microscale computations.

In Chapter 3, the Direct FE2 is validated for a 2D composite cantilever problem.

The method is then extended to 3D in which 3D solid elements are used for

modelling of both the macro and micro scales. A Direct FE2 laminate modelling

methodology is presented that has the potential to capture both intra and inter

ply failure modes explicitly in a single model. The inputs will comprise of the

constituent constitutive and failure properties. The methodology is validated by

showing that results obtained from the model reduces to the classical laminate

theory (CLT) analysis, when fibre and matrix are modelled with homogenized

ply properties (the same inputs as CLT).

In Chapter 4, tension, bending and torsion pipe experiments performed on

Stewart platform-based test setup are described in detail. Chapter 5 employs the

3D modelling strategy explained in chapter 3 to model composite pipe loading

experiments and the model predictions are compared with the experimental

measurements. The macroscopic and microscopic predictions are contrasted

with the experimental observations.

Chapter 6 concludes the work highlighting the major outcomes of this work

with an outlook for the future scope of research.

17
Direct FE2 – Theory, implementation and 2D

validation

Introduction

Chapter 1 gives a brief review of multiscale methods and FE2. In this chapter

the Direct FE2 method is introduced. A novel aspect of the Direct FE2 method

proposed here is that, it allows us to combine two concurrent Finite Element

(FE) calculations for the micro and macro scales into one FE calculation. In FE2

modelling, the finite element calculations at the macroscale can be eliminated

if it is not required (as will be explained in Section 2.1), which implies there is

no requirement for homogenizing the constitutive material model for the

macroscale. The model comprises of only volume-scaled Representative

Volume Elements (RVEs) placed at the macroscale element integration points,

which is connected to the macroscale element nodes with appropriate

interpolation functions using Multi-Point Constraints (MPCs).

FE2 is often employed for problems where there is a clear separation of length

scales between the size of the RVEs and the macroscale finite elements, i.e., the

RVEs appear as points inside the macroscale finite elements. The principle of

separation of scales states that “The microscopic length scale is assumed to be

much smaller than the characteristic length over which the macroscopic

loading varies in space”[102]. This implies that the macroscopic deformation

gradient remains constant over the spatial length scale associated with the RVE

size. There are no numerical limits to define the separation of the two scales.

18
In the problems discussed in this chapter, the micro scale is not infinitely smaller

but finitely smaller than the macro scale element similar to [90,103–105]. The

problems discussed in this and the next chapter do not have a clear separation

of length scales to ensure that our brute force full FE reference models (with

explicitly meshed heterogeneities) are computationally tractable for comparing

with the Direct FE2 models.

Section 2.2 discusses the theoretical background of the FE2- how the macroscale

computations are eliminated, and the macroscopic-microscopic scale transition

relationship is brought about through MPCs. Section 2.3 briefly describes the

implementation of the Direct FE2 method in AbaqusTM. Section 2.4 discusses

the appropriate boundary conditions of a full FE reference model to benchmark

and validate a single Direct FE2 macroscale element. The importance of

separation of length scales is also discussed. Section 2.5 concludes the chapter.

Appendices A, B and C sections elaborate on the application of micro scale

boundary conditions with MPCs and straight edge and periodic boundary

conditions employed on the full FE reference model respectively.

Direct FE2 theory

FE2 methods are employed for stress analysis of heterogeneous materials

[40,65,68,106]. The method comprises of two nested levels of FE simulations

namely the – macroscale and microscale. At the macroscale, the entire

heterogeneous material or structure is discretized into homogenized continuum

finite elements. The microscale FE simulations are performed on multiple

identical computational models of a RVE, where the different phases of the

heterogeneous material are explicitly modeled. Each Gauss quadrature point of

19
the macroscale finite elements is associated with its own microscale RVE. The

FE2 method consists of three main components according to Feyel [43,106],

which are :

1. A geometrical description and a FE model of the unit cell, which is the

RVE.

2. The local constitutive laws describing the response of each component

of the composite within the unit cell.

3. Scale transition relationships that define the connection between the

microscopic and the macroscopic fields (stress and strain). These are: (i)

a downscaling rule that determines the microscale boundary problem for

the RVE given the macroscopic deformation measures; (ii) an upscaling

rule for the macroscopic stress given the micromechanical stress state.

Since the macroscopic behaviour is determined by the homogenization of

microstructures in FE2, the effective macroscopic constitutive properties are not

explicitly given as inputs. The macroscopic stress-strain relation (constitutive

definition), is determined from the homogenization of the RVE microstructures,

which also serve as the up-scaling transition relation. This macroscopic

constitutive definition at a point can be determined by (i) applying the local

macroscopic strain to a microstructural RVE belonging to the point in such a

way that the RVE averaged strains equal the local macroscopic strains, and by

(ii) averaging the non-uniform RVE stress field.

In this method of Direct FE2, the scale transition is provided by interpolation

functions given as multi-point constraints (MPCs) linking the boundary

20
microscale nodes with the macroscale nodes. One of the aspects that

differentiates this method from traditional FE2 is that no separate microscale

boundary conditions are required. The microscale boundary conditions arise

naturally because of this strong coupling between the macroscale and the

microscale through the MPCs.

The remaining section describes how the macroscale computations can be

directly obtained from the micro FE analysis, while eliminating the macroscale

finite element computations.

Here 𝒖, 𝝈, 𝒃 and 𝒕 represents the displacement, stress tensor, body forces and

tractions respectively, and 𝑉 and 𝑆 denote the computational domain and its

boundary respectively. Equation (2.1) is the static equilibrium equation:

𝜎𝑖𝑗,𝑗 + 𝑏𝑖 = 0

(2.1)

The weak form of equation (2.1) is given as

∫ 𝛿𝑤𝑖 (𝜎𝑖𝑗,𝑗 + 𝑏𝑖 ) 𝑑𝑉 = 0
𝑉

(2.2)

where, δwi represents the weight functions. A weak form is a weighted-integral

statement of a differential equation in which the differentiation is transferred

from the dependent variable to the weight function[107] and includes the natural

(Neumann) boundary conditions of the problem.

21
Integrating equation (2.2) by parts:

⇒ ∫ ((𝛿𝑤𝑖 𝜎𝑖𝑗 ),𝑗 − 𝛿𝑤𝑖,𝑗 𝜎𝑖𝑗 + 𝛿𝑤𝑖 𝑏𝑖 ) 𝑑𝑉 = 0


𝑉

⇒ ∫ δwi,j σij dV = ∫ ((δwi σij ) + δwi bi ) dV


,j
V V

(2.3)

Applying the Divergence theorem: ∫𝑉(𝛿𝑤𝑖 𝜎𝑖𝑗 ),𝑗 𝑑𝑉 = ∫𝑆 𝛿𝑤𝑖 𝜎𝑖𝑗 𝑛𝑗 𝑑𝑆 to


equation (2.3) we obtain:

⇒ ∫ 𝛿𝑤𝑖,𝑗 𝜎𝑖𝑗 𝑑𝑉 = ∫ (𝛿𝑤𝑖 𝑏𝑖 )𝑑𝑉 + ∫ 𝛿𝑤𝑖 𝜎𝑖𝑗 𝑛𝑗 𝑑𝑆


𝑉 𝑉
𝑆

Substituting traction 𝜎𝑖𝑗 𝑛𝑗 = 𝑡𝑖 we get:

∫ 𝛿𝑤𝑖,𝑗 𝜎𝑖𝑗 𝑑𝑉 = ∫ (𝛿𝑤𝑖 𝑏𝑖 )𝑑𝑉 + ∫ 𝛿𝑤𝑖 𝑡𝑖 𝑑𝑆


𝑉 𝑉
𝑆

(2.4)

Using Galerkin formulation in which the weights are the displacement

approximation functions (𝛿𝑢𝑖 = 𝑁𝐼 𝛿𝑢𝐼𝑖 ) of the degrees of freedom, i.e.

substituting 𝛿𝑤𝑖 = 𝛿𝑢𝑖 in equation (2.4):

∫ 𝛿𝑢𝑖,𝑗 𝜎𝑖𝑗 𝑑𝑉 = ∫ (𝛿𝑢𝑖 𝑏𝑖 )𝑑𝑉 + ∫ 𝛿𝑢𝑖 𝑡𝑖 𝑑𝑆


𝑉 𝑉
𝑆

(2.5)

In FE, the domain is discretized into elements and substituting the

displacements as the interpolation of the nodal displacements (𝛿𝑢𝑖 = 𝑁𝐼 𝛿𝑢𝐼𝑖 )

in equation (2.5), we get:

22
𝛿𝑢𝐼𝑖 ∫ 𝑁𝐼,𝑗 𝜎𝑖𝑗 𝑑𝑉 = 𝛿𝑢𝐼𝑖 ∫ (𝑁𝐼 𝑏𝑖 )𝑑𝑉 + 𝛿𝑢𝐼𝑖 ∫ 𝑁𝐼 𝑡𝑖 𝑑𝑆
𝑉 𝑉
𝑆

Eliminating 𝛿𝑢𝐼𝑖 and substituting ∫𝑆 𝑁𝐼 𝑡𝑖 𝑑𝑆 = 𝑓𝐼𝑖 , as nodal force on node I in i


𝑡

direction:

⇒ ∫ 𝑁𝐼,𝑗 𝜎𝑖𝑗 𝑑𝑉 = ∫ (𝑁𝐼 𝑏𝑖 )𝑑𝑉 + ∫ 𝑁𝐼 𝑡𝑖 𝑑𝑆


𝑉 𝑉
𝑆𝑡

⇒ ∫ 𝑁𝐼,𝑗 𝜎𝑖𝑗 𝑑𝑉 = ∫ (𝑁𝐼 𝑏𝑖 )𝑑𝑉 + 𝑓𝐼𝑖


𝑉 𝑉

(2.6)

Using Gaussian quadrature to evaluate the integrals in equation (2.6):

∑(𝑤𝛼 (𝑁𝐼,𝑗 𝜎𝑖𝑗 )𝐽𝛼 ) = ∑(𝑤𝛼 (𝑁𝐼 𝑏𝑖 )𝐽𝛼 ) + 𝑓𝐼𝑖


𝛼 𝛼

(2.7)

The equations ((2.1) to (2.7)) describes how the equilibrium equation can be

numerically integrated with Gaussian quadrature in FE analysis. In FE2, the

objective is to solve equation (2.5) at the macroscale FE level using values of

𝜎𝑖𝑗 that are concurrently determined at the microscale FE level. The following

paragraphs explain how the microscale FE analysis can be directly inserted into

equation (2.5) through geometric scaling and applying kinematic constraints

resulting in a Direct FE2 approach that combines the two levels of nested FE

computations into a single FE analysis. It is also shown that this Direct FE2 can

be carried out on a commercial FE platform that supports the application of

multi-point constraint equations (MPCs).

23
Getting the macroscale quantities from scaled microscale computations

(a) (b)

Figure 2.1 Microscale gauss point RVEs in (a) regular and (b) irregularly
shaped Direct FE2 macroscale elements

Figure 2.1 (a) shows a regular and Figure 2.1(b) shows an irregularly shaped

two-dimensional macroscale element with four RVEs centered at the Gauss

quadrature points of the macroscale element. Even though here it is

demonstrated for a two-dimensional case, it is to be noted that the idea of Direct

FE2 is not limited to two dimensional analyses, which will be illustrated in the

next chapter.

Equation (2.5) is the statement of the principle of virtual work, which states that

the internal virtual work, 𝛿𝑊𝑖𝑛𝑡 , is equal to the external virtual work, 𝛿𝑊𝑒𝑥𝑡 .

Gaussian quadrature (equation (2.7)) is invariably employed to perform

numerical integration in FE analyses. Hence, the computation of the Left Hand

Side (LHS) of equation (2.5) takes the form of:

𝛿𝑊𝑖𝑛𝑡 = ∑ ∑ (𝑤𝛼 𝐽𝛼 𝛿𝑢𝑖,𝑗 (𝒙𝛼 )𝜎𝑖𝑗 (𝒙𝛼 ))


𝑒
𝑒 𝛼

(2.8)

24
where, 𝛼 denotes a Gauss quadrature point in element 𝑒, 𝐽𝛼 is the Jacobian and

𝑤𝛼 is the weight for the Gauss point.

Since the stresses for each Gauss point is the volume averaged stresses

calculated from the corresponding RVE, equation (2.8) becomes:

𝛿𝑊𝑖𝑛𝑡 = ∑ ∑(𝑤𝛼 𝐽𝛼 〈𝛿𝑢̃𝑖,𝑗 〉𝛼 〈𝜎̃𝑖𝑗 〉𝛼 )𝑒


𝑒 𝛼

(2.9)

where 〈∙〉𝛼 denote volume averaged quantities over the RVE associated with

Gauss point 𝛼 within element 𝑒. The accent ‘~’ is used to denote quantities from

the microscale calculations.

The Hill-Mandel homogenization condition requires that:

〈𝛿𝑢̃𝑖,𝑗 〉〈𝜎̃𝑖𝑗 〉 = 〈𝛿𝑢̃𝑖,𝑗 𝜎̃𝑖𝑗 〉

(2.10)

Hence, equation (2.9) is equivalent to:

𝑤𝛼 𝐽𝛼
𝛿𝑊𝑖𝑛𝑡 = ∑ ∑ ( ∫ 𝛿𝑢̃ 𝜎̃ 𝑑𝑉)
𝑉𝛼 𝑉𝛼 𝑖,𝑗 𝑖𝑗
𝑒 𝛼 𝑒

(2.11)

where, 𝑉𝛼 is the volume of the RVE associated with the macro scale element

Gauss point and 𝛿𝑊𝑖𝑛𝑡 is the internal work at the macroscale.

Note that the expression for the total sum of internal virtual work calculated

from the microscale FE analysis of all RVEs is given by:

25
̃𝑖𝑛𝑡 = ∑ ∑ (∫ 𝛿𝑢̃𝑖,𝑗 𝜎̃𝑖𝑗 𝑑𝑉)
𝛿𝑊
𝑒 𝛼 𝑉𝛼 𝑒

(2.12)

Comparing equations (2.11) and (2.12), it is seen that 𝛿𝑊𝑖𝑛𝑡 is equivalent to a

̃𝑖𝑛𝑡 , i.e., the internal virtual work for every RVE, each scaled
scaled sum of 𝛿𝑊

by a factor equal to:

𝑤𝛼 𝐽𝛼
𝑤
̅𝛼 =
𝑉𝛼

(2.13)

The scaling factors are straightforward to determine. For example, for two

dimensional FE2 analyses where rectangular elements with 2 × 2 Gauss

quadrature points are employed for the macroscale FE analysis:

1 𝑉𝑒
𝑤
̅𝛼 =
4 𝑉𝛼

(2.14)

where, 𝑉𝑒 is the volume of the macroscale finite element.

The main point of equations (2.11) to (2.14) is that, by scaling the size of the

RVEs with the factor 𝑤


̅𝛼 , the internal virtual work for the microscale FE

analysis in FE2 can be made equal to that of the macroscale FE analysis.

Using equation (2.11) for 𝛿𝑊𝑖𝑛𝑡 means that the LHS of equation (2.5) is now

an expression in terms of the degrees of freedom (DOF) of the microscale RVE

26
FE simulations while the Right Hand Side (RHS) remains a function of the

DOFs of the macroscale FE model as shown in the following equation:

𝑤𝛼 𝐽𝛼
∑∑( ∫ 𝛿𝑢̃ 𝜎̃ 𝑑𝑉) = ∫ 𝛿𝑢𝑖 𝑏𝑖 𝑑𝑉 + ∫𝛿𝑢𝑖 𝑡𝑖 𝑑𝑠
𝑉𝛼 𝑉𝛼 𝑖,𝑗 𝑖𝑗 𝑉 𝑠
𝑒 𝛼 𝑒

(2.15)

Performing the usual finite element discretization, equation (2.15) in the finite

element space becomes:

̃ ∗IJ 𝐝̃𝐉 δ𝐝̃I


δWint = 𝐊 ̃ 𝑱 𝜹𝒅
̃ ∗𝑰𝑱 𝒅
}⇒𝑲 ̃ 𝑰 = 𝒇𝐾 𝛿𝒅𝐾
δWext = 𝐟𝐊 δ𝐝𝐾

(2.16)

̃ ∗ is the overall stiffness matrix assembled from all the microscale RVE
Here, 𝑲

FE where the stiffness matrix from each RVE is scaled by 𝑤


̅𝛼 . The vector of

̃ while 𝒇 and 𝒅 are the vectors


microscale nodal displacements is denoted by 𝒅

of nodal forces and displacements of the macroscale FE mesh.

In many multiscale methods, including FE2, the microscale FE analysis of the

RVEs are boundary value problems, i.e., the displacements of nodes on the

boundaries of the RVE are coupled to the displacement field within the

macroscale finite elements. For first order homogenization schemes, the FE

nodal displacements at the macroscale are linearly related to the microscale

̃ and 𝒅 can be linked through a matrix of


nodal displacement. As such, 𝒅

constants, 𝑳, that represents the scale transition relationship:

̃𝑰
𝒅𝐾 = 𝑳𝐼𝐾 𝒅

(2.17)

27
̃ , we
Inserting (2.17) into (2.16) and eliminating the virtual displacements 𝛿𝒅

obtain:

̃ 𝑱 = 𝑳𝑰𝑲 𝒇𝑲
̃ ∗𝑰𝑱 𝒅
𝑲

(2.18)

Equation (2.18) shows how both levels of FE2 simulation are combined into a

single FE simulation at the microscale level with a stiffness matrix that is scaled

by 𝑤
̅𝛼 and microscale nodal forces that are mapped from the macroscale nodal

forces through matrix 𝑳. This is depicted in the flowcharts shown in Figure 2.2.

(a) (b) (c)

Figure 2.2 Solution algorithms for (a) usual FE simulation, (b) classical FE2
and (c) Direct FE2. f ̃ is the vector of microscale nodal forces

Figure 2.2(a) shows the usual iterative calculations required in a nonlinear FE

simulation where homogenized constitutive relations are available. In usual FE2

where homogenized constitutive equations are not available, a nested iterative

loop is required to obtain the homogenized stress from RVE calculations as

shown in Figure 2.2(b). In Direct FE2 as shown in Figure 2.2(c), the nested loop

is eliminated by solving for the microscale degrees of freedom and before using

equation (2.17) to obtain the macroscale degrees of freedom.

28
Implementation techniques

This Direct FE2 approach can be implemented on commercial FE platform

entirely at the pre-processing stage through two key steps – (i) defining 𝐿 and

̃ ∗ in equation (2.18).
(ii) scaling the RVEs to obtain 𝐾

Many FE codes accept linear MPCs [108] – lines of instructions to impose linear

relationships on the displacements of groups of nodes. Collectively, these

instructions are the same as equation (2.17). Hence, MPCs can be defined to tie

the nodes from the macroscale FE mesh to the nodes from the microscale mesh

on the boundaries of the RVEs as will be described in Section 2.2.3 and

Appendix E. By using MPCs available on commercial FE platforms to

impose the above kinematic constraints, the matrix 𝑳 in equation (2.17) is

̃ ∗ can be obtained.
defined. Next it is shown how 𝑲

̃ ∗ is the stiffness matrix purely due to the microscale


As mentioned previously, 𝑲

FE mesh of the RVEs multiplied by 𝑤


̅𝛼 . With the macroscale mesh and

microscale mesh in the same FE model, the stiffness matrix calculated from

commercial codes will comprise of the stiffness from both the macroscale mesh

and the stiffness of the microscale mesh without any scaling. To circumvent

this, firstly, the macroscale finite elements are assigned a negligible stiffness so

that they make no contribution to the stiffness matrix. Secondly, the size of the

FE mesh of the RVEs relative to that of the macroscale is scaled up so that 𝑤


̅𝛼 =

1, i.e., the relative size of the RVEs to that of the macroscale elements in the FE

model is much larger than the physical relative sizes. From equation (2.13), this

means that:

29
Vα = wα Jα

(2.19)

If rectangular elements with 2 × 2 Gauss quadrature points are employed for

the macroscale FE in a two-dimensional FE2 analysis, equation (2.14) shows

that the size of the RVE meshes needs to be amplified so they are a quarter of

the size of the macroscale element. For two dimensional analyses, a simple way

to achieve this amplification is to specify a larger thickness for the finite

elements of the RVEs than the macroscale elements so that equation (2.19) is

achieved.

A near zero macro-scale element stiffness will not affect the RHS of equation

(2.5), which is not a function of any material constitutive relationship. By

substituting 𝛿𝑢𝑖 = 𝑁𝐼 𝛿𝑢𝐼𝑖 into equation (2.5), the RHS of equation (2.5) will

eventually lead to nodal forces 𝑓𝐼𝑖 (as shown in equation (2.6)) on the Neumann

boundaries. If these macro-scale nodal forces are applied directly at the macro-

scale nodes at the pre-processing stage, then there is no need to define any

macroscale finite elements in the commercial FE code and there will be no error

in the computation of internal virtual work. If, however, the nodal forces change

with nodal displacements, e.g. a pressure load on a surface that undergoes large

deformation, then it will be advantageous to define the macroscale finite

elements to leverage on the commercial FE platform’s in-built capability to

handle large deformation.

FE2 can be carried out directly at the microscale level FE analysis by scaling the

size of the RVEs according to equation (2.13) to satisfy 𝑤


̅𝛼 = 1 (which implies

equation (2.19)), and applying the appropriate boundary conditions to the RVE

30
through MPCs. In short, the macroscale level FE analysis can be eliminated in

FE2 calculations if the homogenized stresses from RVE calculations are to be

obtained using the Voigt-Taylor model or periodic boundary conditions which

are discussed in section 2.2.3.

Boundary conditions at the microscale

Having shown how the LHS of equation (2.5) for the macroscale FE analysis

can be determined from the microscale FE analysis and how the macro-scale

nodal forces are mapped to the microscale (in equation (2.18)), the scale

transition relationships are explained here. How the nodal displacements on the

boundaries of the RVEs are tied to the deformation of the macroscale FE defines

the transition from the macroscale to the microscale. The scale transition

relationship denoted by 𝑳 in the equations mentioned previously, define the

connection between macroscale to microscale displacements.

The microscale FE analysis of the RVEs are boundary value problems that can

be solved with appropriate boundary conditions depending on the chosen

homogenization scheme. The RHS of equation (2.5) can be reproduced in the

microscale FE analysis through the application of appropriate kinematic

relationships between the boundary conditions of the RVEs and the

displacement field within the macroscale finite elements, i.e. by the appropriate

definition of 𝑳.

Three classical boundary conditions [92] that are used to obtain homogenized

constitutive relations from RVEs are:

 Linear displacement boundary conditions (Voigt-Taylor model)

31
 Traction or constant stress boundary conditions (Reuss-Hill model)

 Periodic boundary conditions (PBCs)

All three satisfy the Hill-Mandel condition. However, periodic boundary

conditions (PBCs) are normally preferred because predictions from the Voigt-

Taylor and Reuss-Hill models correspond to the upper and lower bounds of

homogenized stiffness respectively [109].With PBCs, good predictions can be

obtained with smaller RVEs.

Displacement boundary conditions (Voigt-Taylor model) can be applied by

equating the displacement of all nodes on the boundary of the RVEs to the

interpolated displacement within the macroscale finite element:

𝐮
̃ J = NI (𝐱 J )𝐮I

(2.20)

̃𝐽 denotes the displacement of node J on the boundary of the RVE mesh,


Here, 𝒖

𝑁𝐼 and 𝒖𝐼 are the shape function and displacement of node I of the macroscale

finite element within which the RVE resides and 𝒙𝐽 is the point within the

macroscale finite element where node J lies. It should be noted that the Hill-

Mandel condition is satisfied for linear displacement boundary conditions only.

32
Figure 2.3 RVE at macroscale element integration point.

Periodic boundary conditions, which are normally preferred, can also be

prescribed for the RVEs. In Figure 2.3, the edge mid-points (not necessarily

nodes) of the RVEs are labelled as points T, B, L1 and R. The relative

displacements of the top and bottom points (T and B) can be tied to the nodal

displacements of the macroscale finite elements through:

̃ T−B = (NI (𝐱 T ) − NI (𝐱 B ))𝐮I


∆𝐮

(2.21)

The relative displacements of the left and right points are similarly constrained:

̃ R−L = (NI (𝐱 R ) − NI (𝐱 L ))𝐮I


∆𝐮

(2.22)

PBCs are applied by imposing the displacement of all nodes on the top boundary

of the RVE (denoted as node t) relative to their corresponding periodic node on

the bottom boundary (denoted as node b) is constrained by equation (2.21) i.e.:

1
not to be confused with the scale transition also denoted by L

33
𝐮 ̃ b = (NI (𝐱 T ) − NI (𝐱 B ))𝐮I
̃t − 𝐮

(2.23)

Similarly, the displacement of all nodes on the right boundary of the RVE

(denoted as node r) relative to their corresponding periodic node on the left

boundary (denoted as node l) is constrained by equation (2.22), i.e.:

𝐮 ̃ l = (NI (𝐱 R ) − NI (𝐱 L ))𝐮I
̃r − 𝐮

(2.24)

Although FE2 has been applied where there is no clear separation of length

scales between the size of the RVEs and the macroscale finite elements [90,103–

105], more often than not, it is employed for problems when the microscale

RVE is orders of magnitude smaller than the macroscale element so that the

microscopic length scale is much smaller than the characteristic length over

which macroscopic loads change [102]. In situations, where there is a clear

separation of length scales, i.e. the RVE size is orders of magnitude smaller than

the macro-element size, the shape function values, 𝑁𝐼 , will remain practically

unchanged within a RVE. Hence the boundary constraints in equations (2.20)

to (2.24) will not give rise to any displacement gradient. It is numerically more

robust to utilise the gradient of the shape function at the Gauss point,𝛁𝑁𝐼 . In

which case, we obtain:

𝐮 ̃ 0 = 𝛁NI (𝐱 0 ) ∙ (𝐱 J − 𝐱 0 )𝐮I
̃J − 𝐮

(2.25)

for the Voigt-Taylor model, and:

34
𝐮 ̃ b = 𝛁NI (𝐱 0 ) ∙ (𝐱 T − 𝐱 B )𝐮I
̃t − 𝐮

(2.26)

̃ 𝑙 = 𝛁𝑁𝐼 (𝒙0 ) ∙ (𝒙𝑅 − 𝒙𝐿 )𝒖𝐼


̃𝑟 − 𝒖
𝒖

(2.27)

for PBCs.

Here, 𝒙0 refers to the location of the Gauss point and the centre of the RVE

within the macroscale finite element.

Working with gradients as shown in equations ((2.25) to (2.27)) means that the

correct deformation is imposed on the RVEs regardless of the size of the RVEs.

Consequently, there is no requirement for the RVE nodes to coincide with the

actual material points inside the macroscale finite element, which was the case

shown in Figure 2.1. This is particularly useful for three dimensional Direct

FE2 analyses because the RVEs can now be sized accordingly to achieve the

required value of 𝑤
̅𝛼 in equation (2.13), which would result in RVEs of the same

size scale as the macroscale elements. For 3D analyses, if the length of the RVE

mesh is scaled by k in all 3 dimensions to satisfy equation (2.19), then the RHS

of equations ((2.20) to (2.27)) must also be scaled by k. The implementation of

these boundary conditions in AbaqusTM is described in section 2.3.2 and also

Appendix E.

Direct FE2 implementation in AbaqusTM

The Direct FE2 model comprises of only the microscale RVEs placed at the

macroscale integration points (Gauss points), which is connected to the

macroscale elements using MPCs. The method of implementing Direct FE2

35
involves: setting up the model- placing appropriately scaled RVEs (as per

equation (2.19)) at the macroscale integration points and defining the scale

transition relationships with interpolation functions as MPCs. In the following

sections the same is explained in detail. See Appendix D for flowchart detailing

the Direct FE2 model setup (also the python script for model setup) in Abaqus.

Setting up the Direct FE2 model domain

Figure 2.4 A FE2 macro scale-element: the dotted lines marks the boundary of
the macroscale element domain defined by the 4 nodes of the macro-element
defined as Reference point); The macroscale nodal displacements are ui, vi;
i=1,2,3,4; and the micro structure RVEs at the gauss-points; MPCs at the edge
nodes of the RVEs.

A single macroscale element with its Gauss point RVEs is shown in Figure 2.4.

If there is no need of conventional finite elements in the macroscale (as

discussed in section 2.2.2), the macroscale model can be built of non-elemental

nodes (Reference Points in AbaqusTM). The nodes of the model shown in Figure

2.4 are defined as reference points and the element nodal connectivity of that

model defines the macroscale elements of the FE2 model. The micro-structure

36
RVEs (with finite element mesh) are placed such that the RVE centres lie on

the Gauss point of the macroscale element for easy visualization and

interpretation of the stress and strain contours. The thickness of each RVE is

determined such that equation (2.19) is satisfied. The macroscale model can be

easily built in the AbaqusTM modelling suite and the Abaqus input file (.inp)

data can be used to define the macro scale element nodal connectivity.

Microscale boundary conditions: applying the MPCs

After completion of the model domain set up, the next step is the definition of

the scale transition relationships linking macro and micro elements as defined

by the matrix ‘L’ (defined in equation (2.17). MPCs[110] are applied to link the

RVE boundary nodes to the macroscale element nodes. At microscale, linear

displacement boundary conditions are applied to the integration point RVEs as

per equation (2.20). The periodic boundary conditions are applied using

equations ((2.23) and (2.24)). See Appendix A for an example with bilinear

shape functions. Equations ((2.25) to (2.27)) can be used when there is a clear

separation of length scales.

After setting up the model domain and constraining the microscale boundaries

through MPCs, the macroscale boundary conditions and constitutive material

models are assigned like any other FE model. The boundary conditions in the

FE2 problem will be generally applied to the macroscale element nodes which

define the geometry of the structure as will be seen for the cantilever beam in

the next chapter and the single macroscale element problems discussed in the

next section. Appendix D presents the python scripts for applying the MPCs.

37
Solution procedure in AbaqusTM

The Direct FE2 solution procedure in Abaqus is shown in Figure 2.5. The

solution procedure in Figure 2.5 is slightly different from that shown in Figure

2.2(c). This is how the Direct FE2 method will be solved if the MPCs are given

in the same form as equations (2.20) to (2.27). The MPCs can be given such that

all macroscale degrees of freedom are eliminated as discussed in section 2.2.1.

See Appendix E for more details about the scale transition relationship

definition through multi-point constraints (MPCs).

Figure 2.5 Flowchart depicting the Direct FE2 method solved using AbaqusTM

Single Direct FE2 macroscale element models

This section aims to examine whether a single Direct FE2 macroscale element

is equal to a straight edge boundary full Finite Element (FE) model or full FE

model with Periodic Boundary Conditions (PBCs). For a homogeneous

38
material, a single finite element’s behaviour needs to be captured and replicated

by a Direct FE2 macroscale element, which is shown in section 2.4.1. The

boundary conditions for an appropriate equivalent full FE model for composite

materials is discussed in section 2.4.2.

Table 2.1 Material properties

Elastic Poisson’s Yield Hardening Yield


Material
Modulus, Ration, Stress Exponent, Offset,
properties
𝐸 (GPa) 𝜈 σ0 (MPa) n α
AS4 Carbon
15* 0.2
Fibre [111]
3501-6 epoxy
4.275 0.35 172.4 5 0.428
matrix [112]
*transverse modulus is used since the fibre cross-sections are being modelled in 2D-plane stress

CPS4 plane stress elements [110] are used in the FE2 and FE models discussed

in this section. The material properties used in this section are listed in Table

2.1. The fibres are modelled as linear isotropic material and matrix is modelled

with a Ramberg Osgood model [113] which is described by the equation below:

𝑛−1
|𝜎|
𝐸𝜀 = 𝜎 [1 + 𝛼 ( 0 ) ]
𝜎

(2.28)

Comparing with single FE element for a homogenous material

The equivalence of single FE2 macroscale element with the RVEs constrained

by the linear displacement boundary conditions and a full FE element is

demonstrated in this section for a homogenous material. A single full FE plane

stress element CPS4 [110] with fibre properties (Table 2.1) is subjected to

transverse displacement of 3 μm in small deformation (Figure 2.6) and 38.1 µm

39
Figure 2.6 Deformed homogeneous single element models for a small
deformation. (Strain E11 refers to stress in the horizontal direction). The
deformed dimensions of the element are given along the edges. The vertical
reaction force in the fixed end for full FE model: 154807.6 µN and FE2 model:
154957.4 μN. All the dimensions are in µm.

in finite deformation (Figure 2.7) formulations respectively. The boundary

conditions are specified in Figure 2.6 top image which is identical to a cantilever

beam boundary conditions.

The RVEs in the FE2 macroscale element model are constrained by

displacement boundary condition as described in equation (2.20). The RVEs in

the Direct FE2 model are assigned thickness as per equation (2.19). The

40
respective FE2 single macroscale element models are also subjected to the same

boundary conditions. The deformed shape and forces on the fixed face for small

deformation and finite deformation formulations are shown in Figure 2.6 and

Figure 2.7 respectively.

Figure 2.7 Deformed homogeneous single element models for a large


deformation. The deformed dimensions of the element are given along the
edges. The vertical reaction force in the fixed end for full FE model:
2.140774N and FE2 model: 2.143697N. All the dimensions are in µm.

The difference in forces and deformed lengths is less than 0.15% and 0.06%

respectively. The differences in the deformed lengths and forces can be

attributed to truncation error and discretization error as several nodal

displacements are solved for in the Direct FE2 macroscale element model

41
compared to the four nodes solved for in the full FE single element model. The

next section compares the FE2 microscale boundary conditions for a

heterogeneous material.

Heterogeneous single macroscale element models

(a) (b)

Figure 2.8 (a) 10x10 Full FE model; (b) Equivalent FE2 model with 4 RVEs at
the gauss points and the macro scale nodes (modelled as reference points in
Abaqus).

A comparison of Direct FE2 and full FE models is carried out for different

heterogeneous macroscale element sizes in this section. The FE2 model is given

periodic constraints (equations (2.23) and (2.24)). To determine the appropriate

boundary conditions for an equivalent full FE model containing composite like

heterogeneous materials, the full FE model is given two types of external edge

constraints discussed in sections 2.4.2.2 and 2.4.2.3.

An identical strain state is imposed, for different macroscale element sizes. The

size of a macroscale element is defined as the number of RVEs per side of the

square reference full FE model, equivalent to the FE2 macroscale element. A

10x10 model (as shown in Figure 2.8 (a)) has a macroscale element size of 10

42
(10 RVEs per side) and a 80x80 full FE model will have a macroscale element

size of 80. The RVE consists of a single fibre of diameter 5μm with fibre volume

fraction of 0.6 placed non-symmetrically in a square unit cell, as shown in

Figure 2.8 (b).

The FE2 model in Figure 2.8 (b) consists of only four RVEs in every case. The

only difference between various FE2 models equivalent to a 10x10 and 80x80

is the absolute positions and thickness given to the RVEs as per equations

((2.13) and (2.19)) and the positions of the macroscale nodes.

Table 2.2 Boundary displacements given to various model sizes; x and y are
the displacements shown in Figure 2.9

Model Size(L) x(µm) y (µm)

10x10 (57.2µm) 5 2.5

15x15 (85.8µm) 7.5 3.75

20x20 (114.4µm) 10 5

40x40 (228.8µm) 20 10

80x80 (457.6µm) 40 20

The boundary conditions of the full FE model and FE2 model are stated in

Figure 2.9(a) and Figure 2.9(b) respectively. The displacements as per the

model size are listed in Table 2.2 to ensure identical strain states. A constant

macroscopic strain state is ensured for comparison across different model sizes.

Before we proceed to compare different FE2 and full FE models, the different

microscale boundary conditions in FE2 macroscale elements for a

heterogeneous material are compared and discussed.

43
(a)

(b)

Figure 2.9 (a) Boundary conditions of the full FE model; (b) boundary
conditions in the FE2 model.

Comparison of Direct FE2 microscale boundary conditions for

heterogeneous materials

It is well known that the displacement boundary condition (equation (2.20))

gives the upper bound of homogenized elastic moduli [109] as can also be seen

44
30000

25000

20000
Shear Force()

15000

10000

5000
periodic boundary conditions
displacement boundary condition

0
0 2 4 6
(a) Displacement(m)

(b)

Figure 2.10 FE2 microscale boundary conditions: displacement boundary


conditions (equation ((2.20)) vs periodic boundary Conditions (equations
((2.23) and (2.24)) for (a) a simple shear deformation and (b) tensile
deformation.

45
in Figure 2.10. A 20x20 FE2 macroscale element model (with fibre centred

RVEs-not the non-symmetric RVEs in Figure 2.8(b)) is simulated in simple

shear (𝑥 = 𝑦 = 5𝜇𝑚 in Figure 2.9) for both microscale boundary

conditions,namely displacement (equation (2.20)) and PBCs (equations (2.23)

and (2.24)) applied to the microscale RVEs. The shearing force vs displacement

of the bottom side is plotted in Figure 2.10(a). Elastic fibre properties (from

Table 2.1) and epoxy properties (𝐸 = 3350MPa, 𝜈 = 0.35 [111]) are used for

the fibre and matrix.

Similar behaviour is observed for tension as can be seen in Figure 2.10(b). The

displacement boundary conditions cannot capture the deformed shape of the

heterogeneous unit cells inside the bulk of full FE model as shown in Figure

2.11 for a simple shear deformation. The deformed RVE from the FE2 model

with periodic constraints is also shown for comparison in Figure 2.11(c).

Therefore, the response of all FE2 integration point RVEs are constrained to the

macroscale nodes with periodic conditions (equations (2.23) and (2.24)) in the

FE2 macroscale element models discussed after this section and the rest of the

thesis unless specified otherwise.

(a) (b) (c)

Figure 2.11 Deformed RVE shapes in: (a) the bulk of full FE model; (b) FE2
macroscale element with displacement boundary condition (equation (2.20)
and (c) FE2 macroscale element with periodic boundary conditions (equations
(2.23) and (2.24)).

46
The following subsections demonstrate the importance of separation of length

scales by comparing single FE2 macroscale element models and equivalent full

FE models. For comparing the FE2 models with full FE models for different

sizes, as mentioned above, the FE2 macroscale elements are given periodic

constraints whereas the reference full FE model’s exterior boundaries are

ensured to remain straight in section 2.4.2.2 and periodic in section 2.4.2.3.

Straight edge constraints for the full FE model

A heuristic approach to determine the appropriate boundary conditions for an

equivalent full FE model containing composite like heterogeneous materials

would be to compare the FE2 model with a full FE model whose exterior

boundaries lie along the same plane. This might be too restrictive for

heterogeneous materials, but it still is a good starting point as a 2D macroscale

FE2 element has 4 nodes which can be joined by straight lines. Therefore,

constraints are applied to ensure that the reference full FE model edges remain

straight. Multi-point constraints equations [110] are employed to ensure straight

exterior edges of the full FE models as described briefly in Appendix B.

The horizontal force on the bottom edge is measured and plotted for various

models in Figure 2.12 (a). A quick glance at the Figure 2.12 (a) might be

convincing that the boundaries of a macroscale FE2 maybe behaving like those

in the straight edge constrained full FE model. After taking a closer look at the

differences between the forces in the two models plotted in Figure 2.12 (b), an

explanation as to why the differences are larger in small sized models than the

large models, needs further investigation.

47
(a)

(b)

Figure 2.12 (a)The force vs normalized displacment behaviiour of differently


sized macroscale element models (b) The difference between the forces in full
FE model and FE2 model with macroscale element size.

As can be seen from Figure 2.12(b), the force difference between the full FE

model and FE2 model decreases with the increase in macroscale element size.

This could be because of the dominance of stress concentration (as seen in

48
Figure 2.13) due to the straight exterior edge constraint in the small sized full

FE models (10x10, 15x15) compared to larger full FE models.

(a) (b)

Figure 2.13 (a) Full FE model edge effect due to straight exterior edge
constraint (b) deformed RVE in FE2 macroscale element with periodic
boundary conditions.

This hypothesis is tested by changing the straight edge boundary constraints of

the full FE model. Periodic boundary conditions are applied on the full FE

model’s boundary and discussed in the next section.

Full FE heterogeneous model with Periodic Boundary Conditions

(PBCs)

PBCs are employed on the exterior edges of the full FE model as per Appendix

C. The same boundary conditions on the differently sized full FE models with

PBCs applied to their edges are compared with the respective FE2 macroscale

element models.

49
(a)

(b) (c)

Figure 2.14 (a) Full FE (10x10 RVEs)Model with PBCs in which no edge
effect is visible; (b) deformed FE2 RVE used in the presented analysis (c) a
deformed RVE from a 10x10 FE2 model with fibre centered RVEs to show
the effect of the RVE mesh on the stress contours.

The edge effect is no longer present in the full FE reference model as observed

from Figure 2.14. It is observed that the stress contours are similar throughout

in the full FE reference and FE2 macroscale element model. The differences in

the forces between the full FE and FE2 models are nearly zero as can be seen in

Figure 2.15. This implies that a single Direct FE2 element behaves in a similar

manner to the full FE model with PBCs. This also demonstrates the importance

of separation of length scales as it observed that differences in response exist in

the absence of PBCs and those differences decrease as we increase the

macroscale element size as seen in Figure 2.15.

50
Figure 2.15 Comparing errors between full FE with PBCs and FE2 and full FE
with straight edge constraints and FE2.

Conclusions

It is shown through the Hill-Mandel homogenization condition that FE2

methods, which requires two concurrent FE calculations performed at the

microscale and macroscale, can be reduced to a single microscale FE calculation

by appropriately weighing the energy contributions from each RVE together

with a matrix L that describes the kinematic scale transition. How the RVE

microscale finite elements and matrix L are related to the macroscale finite

elements are presented. As these relationships can be explicitly defined based

on macroscale mesh at the reference state and the choice of interpolation

functions, the RVE weights and matrix L (Appendix E) can be determined at

the pre-processing stage of the FE analyses. Consequently, Direct FE2 can be

carried out on commercial FE platform as a single standalone FE simulation. A

noteworthy outcome is that the material libraries and many in-built features of

51
the commercial code become immediately available as will be shown in the next

chapter. This method can be readily and easily extended to 3D and can be

coupled with full FE domains wherever FE2 domains are insufficient to capture

the physical structural behaviour (stress concentration) as observed in straight

exterior edge constrained full FE model.

A FE2 macroscale element is shown to be equivalent to a single full FE element

for homogenous materials. The FE2 macroscale element with periodic boundary

conditions is shown to exhibit behaviour similar to a full FE heterogeneous

model with PBCs applied to the boundaries. This chapter explained and

validated the Direct FE2 method. As just 4 RVEs can replace 6400 RVEs (in

80x80 model) in a FE2 macroscale element, the method is computationally

attractive to model large scale heterogeneous structures. The next chapter will

discuss the application of Direct FE2 to model a cantilever beam and angle plied

composite laminates.

52
Application of Direct FE2 to 2D beams and 3D

composite laminates

Having validated the Direct FE2 method with single macroscale elements, in

this chapter the application of Direct FE2 method to model a 2D composite

cantilever beam problem is demonstrated. The results are compared against

brute force full Finite Element (FE) computation models to validate and

benchmark the Direct FE2 method. The in-built material libraries of commercial

codes are demonstrated to be naturally available with Direct FE2 through

examples involving large deformation, plasticity and viscoelasticity problems

for a composite. The use of this method with irregular quadrilateral elements

and cohesive interaction in composite material is also illustrated.

One of the key contributions of this work is the demonstration of the method to

model composite laminates, which will be presented and compared with the

classical lamination theory [114]. The chapter will conclude with the modelling

of the elastic behaviour of a multi-directional composite laminate pipe section

to demonstrate the method’s capability to model structures bottom up from the

constituents.

Effect of macroscale mesh

The cantilever beam is modelled as a carbon fibre reinforced (CRP) composite

with fibres running perpendicular to the plane of deformation in AbaqusTM. A

Direct FE2 model of the beam and a full FE model of the beam are shown in

Figure 3.1. CPS4 [110] continuum plane stress elements are used in all the

simulations.

53
(a) (b)

Figure 3.1 Cantilever beam (a) FE2 model with 2x20 macroscale elements with
the macro-element boundaries (b) full FE model.

Figure 3.2 RVE used in the simulations.

Figure 3.2 shows the FE mesh of the RVE used in the simulations. The RVE

contains a single fibre of 5µm diameter and has a fibre volume fraction of 0.6.

Due to symmetry, the RVE in Figure 3.2 requires less elements for applying the

periodic constraints. Most FE2 analyses in this chapter are conducted using this

symmetric RVE, unless specified. The effect of RVE architecture and

microscale mesh in Direct FE2 is discussed in section 3.3. The material

properties of the fibre and matrix used in this section are given in Table 3.1

below.

Table 3.1 Material properties.

Material Young’s Modulus, E(MPa) Poisson’s ratio, ν

Epoxy matrix 3350 0.35

Carbon Fibre 230000 0.2

54
FE2 macroscale mesh convergence study with a small cantilever beam

with geometric non-linearity

A FE2 cantilever beam that is equivalent to a full FE cantilever beam constructed

with 200 RVEs (RVE shown in Figure 3.2) in length and 20 RVEs in width is

used for this study. The beam is discretized with different sized macroscale

meshes to study the convergence properties of the Direct FE2 method. The same

RVE with the discretization shown in Figure 3.2 is used with its thickness scaled

as per the macroscale element to satisfy equation (2.19).

The beam is deflected by one-third its length at the free end. The macroscale

element mesh is refined till convergence with a full Finite element (FE) solution

is achieved. The stress contours of the various macroscale element mesh near

the fixed end is shown in Figure 3.3(a-e). As the meshes are refined, the stress

limits approach closer to the full FE model. The peak stresses in the full FE

beam model are observed close to the beam boundaries. However, since there

are no RVEs along the beam boundary, the peak stress values will be lower in

the FE2 model, which can be seen for the model with 8 macroscale elements in

the thickness direction (Figure 3.3(e)). The force displacement behaviour is

plotted in Figure 3.3(f). Four FE2 macroscale meshes are modelled, which have

1, 2, 4 and 8 macroscale elements along the beam thickness in the transverse

direction. In Figure 3.3, the FE2 model that satisfactorily converges with full FE

model contains 8 macroscale elements in the thickness.

For linearly interpolated macroscale elements, the FE2 model requires 2560

RVEs in 8x80 FE2 model seen in Figure 3.3(e) to converge with the full FE

55
3500
Full FE reference
1x10-linear
2x20-linear
3000
2x20-linear-finer mesh
4x40-linear
8x80-linear
2500
Force(N)

2000

1500

1000

500

0
0 100 200 300 400

Displacement(m)

Figure 3.3 Stress contours of region closest to the fixed end of the beam (a)
full FE, (b) 1x10 FE2 linear macroscale elements, (c) 2x20 FE2 linear
macroscale elements, (d) 4x40 FE2 linear macroscale elements, (e) 8x80 FE2
linear macroscale elements; (f) Force-displacement behaviour for linear elastic
epoxy and fibre with geometric non-linearity cantilever beam of various linear
interpolated FE2 models and full FE model.

56
model that contains 4000 RVEs. This is due to the presence of high transverse

strain gradients in the cantilever beam.

It can be observed that the Finite element mesh used for the micro scale RVEs

affect the force displacement. The peak force at 381µm deflection differs by

0.7% in the 2x20 FE2 models for two different mesh sizes used for RVE shown

in Figure 3.3 (f). The effect of the RVE mesh is discussed in section 3.3

Effect of beam size on macroscale mesh requirements

The beam used in section 3.1.1 is shown in Figure 3.4(b). This section will study

the effect of beam size on macroscale mesh requirements.

(a)

(b)

Figure 3.4 Relative size of the cantilever beam models used in this work (a)
Large beam on top with size of -2000RVEs(in Figure 3.2) x 200 RVEs with a
8x80 macroscale element mesh and (b) Short beam on the bottom with size of
200RVEs x 20 RVEs with a 2x20 macroscale element mesh.

The effect of beam size on the method performance and the macroscale mesh

requirements is studied by modelling a beam (2000 RVEs x 200 RVEs in size

shown in Figure 3.4(a)) that is 100 times (10 times length and 10 times width)

larger in size compared to the short beam (Figure 3.4(b)). The model size

parameters of the large beam and the small beam used here are listed in Table

3.2.

57
8 macroscale elements are needed for the Direct FE2 short beam to converge

with the full FE reference solution. The large beam is also meshed with the same

macroscale discretization, which means that same number of macroscale

elements is required to model the larger cantilever beam. Both beams are

deflected by one-tenth its length at the free end and compared with the

corresponding full FE results in Figure 3.5. RVEs are constrained with periodic

boundary conditions using linear interpolation functions for the macroscale

elements.

Table 3.2 Size of the beam model.

Beam Size RVE length: RVE length:8x80


beam length macroscale
element size
Large beam 2000RVEs x 200 1:2000 1:25
(Figure 3.4(a)) RVEs
Small beam 200RVEs x 20 RVEs 1:200 1:2.5
(Figure 3.4(b))

The full FE models and the FE2 models behave similarly as seen in Figure 3.5.

The linearly interpolated FE2 models appear to be stiffer than the full FE

models. However, the difference is merely 1.7% with respect to the full FE

models. It is to be noted that both the large and the small beam require the same

number of macroscale elements, which is 640 (8x80) in this case. Only the

thickness of the RVEs in the FE2 model will vary as per equation (2.19). This

mesh convergence study has been done with a composite beam. The method

will be exact for a full FE homogenous material beam with the same number of

macroscale elements as the Direct FE2 beam models as will be discussed and

shown in section 3.2.1. For composite material beams, this difference between

58
the FE2 models and full FE models can be reduced by using higher order

interpolation functions as discussed in section 3.2.2.

Figure 3.5 Normalized force vs normalized displacement: comparison of large


and short beam behaviour.

Demonstration with different matrix constitutive models

The large cantilever beam (Figure 3.4 (a)) used for the simulations described in

this section is shown in Figure 3.6(a) with a quarter of a macroscale element

contrasted with its equivalent full FE region to show the difference in the

computations required between a full FE and a Direct FE2 model of the same

composite beam. The following sections demonstrate the immediate and easy

availability of in-built material constitutive libraries in commercial software,

which this FE2 implementation makes possible without requiring any major

59
change to the FE2 models except the different properties demanded by the

constitutive models. This is shown with the Abaqus FE software.

(a)

(b) Direct FE2 (c) Full FE

Figure 3.6 (a) Cantilever beam with a 8x80 macroscale element mesh with an
integration point region of a highlighted macroscale element represented in
Direct FE2 (thickness scaled RVE) and full FE (156.25 RVEs); (b) A RVE at
FE2 macroscale element integration point; (c) A full FE region equivalent to
single integration point RVE in 8x80 Direct FE2 model ( 𝑤 ̅𝛼 = 156.25) as
per equations (2.13, 2.19) for this macroscale discretization).

Linear elastic matrix model with geometric non-linearity

The first test case is of a homogeneous elastic cantilever beam subjected to large

deformation. The results from the FE2 model of the beam are compared to those

of a full FE model for validation. The FE2 beam comprises of 8 × 80 macroscale

elements. Linear interpolants are employed with 2 × 2 Gauss points per

elements. The RVE mesh shown in Figure 3.2 is used for each Gauss point. In

total, the FE2 beam contains 2560 RVEs. To construct the reference full FE

60
beam model, each macroscale element in the FE2 beam model is replaced by

25 × 25 RVE meshes, i.e., the full FE model contains a total of 400,000 RVEs.

The thickness specified for the elements in the FE2 model is 156.25 times that

for the full FE model to satisfy equation (2.19).

Figure 3.7 Homogeneous beam: Validating the Direct FE2 method.

For both FE models, all elements are assigned the properties of epoxy to model

a homogeneous elastic beam. A vertical displacement equal to one third of the

beam length is applied to the free end of the beam. A nonlinear analysis is

performed to account for the large deformation. Two FE2 analyses are

performed – one with linear displacement boundary conditions (equation (2.20))

on the RVEs and another with periodic boundary conditions (equations 2.23 and

2.24). The calculated load-displacement graphs from both FE2 analyses and the

full FE model (with the same number of elements as the number of FE2

61
macroscale elements) are plotted in Figure 3.7. It can be seen that all three force-

deformation curves are practically coincident.

Since the beam is homogeneous, very close agreement with the full FE model

is expected since the homogenized stresses from the RVEs would be exactly the

same if the material properties are directly assigned to the macroscale elements.

The converged full FE solution is also plotted for comparison. The full FE

model is slightly more compliant due to the much larger degrees of freedom

compared to the macroscale mesh of the FE2 beam model.

In the second test case, the large deformation of an elastic composite beam is

investigated. The models are identical to those of the first test case except that

the fibres and matrix are assigned their corresponding properties as listed in

Table 3.1. Again, the FE2 simulations were performed with linear displacement

(equation 2.20) and periodic boundary conditions (equations 2.23 and 2.24).

The converged full FE homogeneous behaviour from Figure 3.7 is also plotted

for comparison. The computation time needed for the full FE composite beam

model increased with increased deflection due to the smaller step increments

required for stability. Hence, the full FE computation was stopped when the

beam deflection was about 70% of that attained in the FE2 computations. The

load-displacement plots from all 3 simulations is shown in Figure 3.8 and it is

observed that the FE2 model with linear displacement boundary condition is

excessively stiff. This is not surprising since the Voigt-Taylor model gives the

upper bound for homogenized stiffness [109]. This is further exacerbated by the

small size of the RVE used.

62
In contrast, the FE2 load-displacement plot obtained with the more commonly

used periodic boundary conditions for RVEs are in good agreement with the full

FE model with only a very slight over prediction of the stiffness. It should be

noted that the FE2 simulation required only 733s using 8 processors to complete,

whereas the full FE model with around 27 million elements took about 24 hours

with 20 processors to attain 60% of the beam deflection.

Figure 3.8 Periodic vs Linear Displacement Boundary Conditions (BCs):


Homogeneous and composite materials (FE2 models contain 8x80 macroscale
elements).

Higher order interpolation function

The Direct FE2 can also be implemented with higher order shape functions. The

previous two FE2 test cases of homogeneous and composite beams are repeated

using quadratic shape functions. The 4-node rectangular macroscale elements

for the previous simulations are replaced by 8-node macroscale elements and

63
the number of Gauss points per element remains at 2 × 2, i.e., both linear and

quadratic FE2 models have the same number of integration points and same

number of macroscale elements (8 × 80). Periodic boundary conditions are

applied on the RVEs. The quadratic periodic Direct FE2 force displacement

behaviour for homogeneous and composite models are plotted along with the

linear periodic FE2 and full FE reference in Figure 3.9.

Figure 3.9 Effect of macroscale element interpolation on Force-Displacement:


quadratic interpolated SE converges faster than linear SE (FE2 models contain
8x80 macroscale elements).

As seen in Figure 3.9, the full FE solutions for both the homogeneous and

composite beams are indistinguishable from the FE2 model with quadratic

interpolation. This is because the interpolation functions plays the role of scale

transition relationships in the Direct FE2 model as stated before. For the

cantilever problem, the quadratic scale transition gives better approximations of

64
the nodal displacements than the linear scale transition. This is because it

interpolates deformed beam domain better as more macroscale nodes are being

used for solving the same domain, which is also observed as the domain in finite

element computations is discretized with more nodes(mesh refinement).

The micro-stress contours in the constituents are also captured better with

quadratically interpolated macroscale elements with respect to the full FE

reference stresses as can be seen in Figure 3.10 which show the stresses at 60%

of the deflection (0.18* beam length).

Figure 3.10 Stress contours in (a) linear FE2; (b) quadratic FE2; (c) Full FE
reference at free end deflection of 0.18*beam length.

For the same material domain modelled, the model with quadratic interpolation

converges better at large displacement towards the full FE solution for the

composite beam compared to the model with linear interpolated FE2 model as

can be seen from the stress limits in Figure 3.10. The vertical displacements at

an equivalent point from the full FE (Figure 3.11 (a)) and Direct FE2 models

(Figure 3.11 (b)) are compared in Figure 3.11 (c) to show the effect of the

interpolation functions used on the computed nodal displacement.

65
(a)

(b)

(c)

Figure 3.11 Equivalent points near the coordinate system for displacement
comparison in (a) Full FE reference; (b) Direct FE2 model (c) vertical
displacement comparison of Direct FE2 model with full FE model- linear
interpolation vs quadratic interpolation.

The displacements of the quadratically interpolated macroscale elements are

within an error of 1% (0.5-0.86%) with respect to the correpsonding

displacements calculated in the reference converged full FE model as shown in

Figure 3.11(c), whereas the displacement errors for the linear interpolated

macroscale elements ranges from 4.75-5.12%. This shows the effect of the

interpolation functions used on the computed displacments in FE2 model.

66
Increasing the number macroscale integration points

A cantilever beam model with quadratic macroscale elements used in section

3.2.2 is used to demonstrate the convergence and consistency of the Direct FE2

method as more integration (Gauss) points are used. RVEs are placed at 3x3

Gauss points and the results for cantilever with 8 node quadratic elements for

composites is shown below compared with the 2x2 Gauss points quadratic

macroscale Direct FE2 model used in the section 3.2.2.

Figure 3.12 Composite cantilever beam force displacement showing


consistency and convergence as more Gauss points are used with macroscale
quadratic elements.

Increasing the number of macroscale integration points also requires using

Equation (2.19) to recalculate the volume scaling factor for each of the RVEs

(since the integration point weights are different). Increasing the macroscale

integration points and using the right volume scaling factor gives consistent

results as expected and can be seen in Figure 3.12.

67
This can be used to advantage in cases where RVEs closer to the macroscale

element boundaries can capture the micro strains (that can be volume averaged

to calculate the macrostrains) closer to the edges of the structure.

Non-linear plastic matrix model with geometric non-linearity

The method’s capability to handle material non-linearity in addition to

geometric non-linearity is studied by assigning a isotropic hardening plasticity

constitutive model for the matrix. Now to investigate and benchmark models

with material non-linearity with full FE, the same beam shown in Figure 3.6 (a)

is used. The fibre and matrix are given the properties listed in Table 3.3 and

Table 3.4.

Table 3.3 Material elastic properties

Material Young’s Modulus, E(MPa) Poisson’s


ratio, ν
Epoxy matrix[115] 3900 0.39
Carbon Fibre 230000 0.2

Table 3.4 Matrix plastic properties [115]

Yield stress Plastic strain


29 0
59 0.02
93 0.049
95.5 0.058

The force displacement behaviour of the different FE2 models is compared with

reference FE model in Figure 3.13. The 8x80 quadratically and linearly

interpolated FE2 model predicts initial behaviour very similar to that of the full

FE model, as can be seen from Figure 3.13.

68
3000
Force(N)

2000

1000

Full FE
FE2 with linear macroscale elements
FE2 with quadratic macroscale elements

0
0 1000 2000 3000 4000
Displacement(m)

Figure 3.13 Load-displacement plots for large deformation of elastic-plastic


composite beam.

The full FE simulation was terminated at a displacement of about 500m due

to excessive deformation of the elements at the fixed edge of the cantilever

beam. This is shown in Figure 3.14 (d). It is well known that the FE2 method

cannot capture highly localized deformation unless the RVE size is identical to

the macroscale element size. As such, the FE2 simulations erroneously

continues past the peak load. As can be anticipated, the peak load predicted by

the quadratic FE2 is slightly lower than that predicted by the linear FE2 model.

This is expected as quadratic interpolation allows strain gradients similar to the

full FE model as discussed in section 3.2.2 compared to linear interpolation,

which is why the quadratic FE2 peak is below the peak force predicted by the

linear FE2 model. However, it should be highlighted that the Direct FE2 model

69
can also be coupled with full FE mesh regions in case a full field analysis of any

critical regions of the model is necessary as elaborated in Appendix F.

Figure 3.14 Equivalent plastic strain contours of RVEs closest to top left of
beam for (a) linear FE2, (b) quadratic FE2 and (c) full FE; (d) Excessive
deformation of elements in full FE simulation in grey at top left edge of beam.

Equivalent plastic strain2 contours of the RVE closest to the top left of the beam

for the FE2 models are shown in Figure 3.14 (a) and (b) at the displacement

𝑡
2
Defined as 𝜀̅𝑝𝑙 = 𝜀̅𝑝𝑙 |0 + ∫0 𝜀̅𝑝𝑙̇ 𝑑𝑡, where 𝜀̅𝑝𝑙 |0 is the initial equivalent plastic strain and 𝜀̅𝑝𝑙̇ is
the equivalent plastic strain rate. Here time is just the way the load increment is accounted for
as the analysis is static and time has no meaning in the analysis carried out here

70
where the full FE simulation is terminated. The FE2 plastic strain contours in

Figure 3.14 (a) and (b) agree well with the full FE contour seen in Figure 3.14

(c).

Viscoelastic matrix model

In this section, a composite beam with strain-rate dependent material model is

simulated. Geometric non-linearity is not included to save computational time

on the full FE models. The matrix is modelled as a viscoelastic material. The

relaxation modulus of the matrix of the composite is described by the Prony

series [116]:

𝑛
−𝑡
𝐸(𝑡) = 𝐸0 − ∑ 𝐸𝑖 (1 − 𝑒 𝜏𝑖 )
𝑖=1

(3.1)

Here, 𝐸0 is the instantaneous modulus of the matrix given in Table 3.5 and 𝜏𝑖 is

the relaxation time. To approximate a Maxwell model using a one term Prony

series, i.e., 𝑛 = 1, we use 𝐸1 = 0.999𝐸0 together with a relaxation time of 𝜏1 =

39.15𝑠 [117]. The fibre is modelled as an elastic material with properties listed

in Table 3.5.

Table 3.5 Material properties

Material Young’s Modulus, E0 (MPa) Poisson’s ratio, 𝜈


Epoxy matrix 4082 0.311
Carbon Fibre 230000 0.2

The beam is loaded to a transverse deflection of about a tenth of the beam length

in 0.3s and unloaded to initial position in 0.7s

71
5000
Full FE
FE2 with linear macroscale elements
FE2 with quadratic macroscale elements
4000
Force(N)

3000

2000

1000

0
0 250 500 750 1000 1250
Displacement(m)
(a) -1000

(b) (c) (d)

Figure 3.15 (a) Load-displacement plots of viscoelastic composite beam.


Residual stresses (MPa) at the end of loading cycle within the RVEs of Direct
FE2 models with (b) linear and (c) quadratic macroscale elements and (d) full
FE model.

The composite beam exhibits hysteresis as seen in Figure 3.15. The Direct FE2

models with linear or quadratic interpolation are able to capture the behaviour

along with the material’s hysteresis exhibited by the computationally more

expensive full FE models. The beam axial residual stresses in the constituents

of both the linearly and quadratically interpolated FE2 models are in good

agreement with the full FE model as can be seen in Figure 3.15 (b)-(d).

72
Effects of microscale RVE: fibre arrangement and microscale

mesh size

The effect of RVE architecture for the same fibre arrangement is discussed. The

RVEs used are shown in Figure 3.16. The fibres in the RVEs in Figure 3.16 (c,

d, and (e) are symmetrically distributed whereas those in Figure 3.16 (b) and

asymmetric. The force-displacement behaviour of both RVEs are plotted in

Figure 3.16 (a). The difference in the force in the non-symmetric RVE model is

between 3.5 to 4.3% with respect to the full FE reference. The difference may

be due to fibre arrangement or the different element size of the symmetric and

asymmetric RVEs or a combination of both factors.

The small deviation in the load displacement behaviour may arise due to the

variation in the location of the fibres in different RVEs. This is observed since

the RVE size is small. Terada et al. (2000) [95] have shown that larger RVEs

become less sensitive to RVE architecture. With increase in RVE size, this

effect of RVE architecture will become negligible as also observed by [103].

The stresses in the non-symmetric and symmetric RVEs are compared in Figure

3.16 (b) and (c) respectively. It is observed that there exists a difference in the

peak stress values. This is because the stress contours are a function of element

size. As we reduce the element size of the symmetric RVE mesh in Figure 3.16

(d)-(e), we can see that the micro-stress contour of the symmetric RVE become

comparable to that of the non-symmetric RVE in Figure 3.16 (b).

The mesh sensitivity of the FE2 models is shown by refining the mesh of the

symmetric RVEs whose stress contours are shown in Figure 3.16 (c)-(e). The

force displacement behaviour differs by a maximum of 4% upon refining the

73
mesh progressively from a coarse meshed RVE in Figure 3.16 (c) to the refined

mesh RVE in Figure 3.16 (e).

(a)

(b) (c)

(d) (e)

Figure 3.16 (a) Effect of RVE architecture: FE2 models contain 8x80
macroscale elements on force-displacement behaviour; Axial stress contours at
deflection of one-third beam length in models with: (b) non-symmetric RVE
with fine mesh; (c) Symmetric RVE with coarse mesh; (d) symmetric RVE
with intermediate mesh; (e) symmetric RVE with fine mesh.

74
The refined mesh of the symmetric RVE shown in Figure 3.16 (e) and the non-

symmetric RVE FE2 models’ force-displacement behaviour are within 0.5%

relative difference of each other which implies that the difference may be due

to the mesh size rather than the fibre arrangement. It is to be noted that the fibre

arrangement is still a simple square arrangement for both the symmetric and

asymmetric RVEs.

Modelling the fibre matrix interface using cohesive elements

The cohesive model [118] is widely used to model interfaces between two

different material phases. This section discusses the use of cohesive formulation

to model the fibre matrix interface in the RVEs used in the previous sections.

Since cohesive calculations are computationally expensive, the smaller beam is

used with 8x80 macroscale elements which is shown in Figure 3.4 (b).

The matrix and the fibre interface are modelled with a cohesive interaction in

Abaqus to study the effect of interface properties between the fibre and matrix.

Cohesive interactions requires a traction separation law given by the equation

[118,119].

𝜏𝑖 = (1 − 𝑑𝑚 )𝐾𝑝 𝛥𝑖 , 𝑖 = 𝐼, 𝐼𝐼 𝑎𝑛𝑑 𝐼𝐼𝐼,

(3.2)

where 𝜏𝑖 and 𝛥𝑖 are the components of the traction vector 𝜏 and separation

vector 𝛥 respectively, and i= 𝐼, 𝐼𝐼 𝑎𝑛𝑑 𝐼𝐼𝐼, denote the three modes of loading,

𝑑𝑚 is the damage factor characterizing the irreversible damage process with

𝑑𝑚 = 0 and 𝑑𝑚 = 1 representing the intact and fully failed states respectively.

75
Separation onset (delamination or debond initiation) is frequently predicted by

the quadratic criterion [120,121]:

2 2 2
𝜏 𝜏 𝜏𝐼𝐼𝐼
( 𝑐𝐼 ) + ( 𝐼𝐼
𝑐) +( 𝑐 ) =1
𝜏𝐼 𝜏𝐼𝐼 𝜏𝐼𝐼𝐼

(3.3)

The traction separation law defines both initiation and damage evolution of the

interface failure. Three parameters determine the traction separation law,

namely the penalty stiffness K, cohesive strength and fracture energy [122]. The

penalty stiffness determines the initial elastic material behaviour before damage

onset, which is a numerical parameter that is given a very high value to prevent

interface compliance. The strength is the maximum stress that the interface can

sustain, and the fracture energy is the material property that quantifies the

energy needed for crack propagation during the damage evolution (the area

under traction separation curve). Moreover, the strength also determines the

failure onset and material softening process. A suitable value for K should

neither be too low nor too high. The former gives rise to inaccuracy due to

interface compliance, while the latter affects the convergence of results [122].

Table 3.6 Traction separation law parameters

Property Value
Cohesive Behaviour (Traction Stiffness Coefficient: Knn=106 µN/µm3,
separation) Kss=Ktt=106 µN/µm3

Initiation (Interface strength)


Damage [123] τI= τII = τIII =39.1MPa
Damage Evolution
Fracture Energy = 100 (µN* µm) / (µm)2

76
(a)

(b)

(c)

70
Full FE model model with c interface strengt
Full FE model model with 2c interface strength
60 quadratic FE2 model with 2c interface strength
quadratic FE2 model with c interface strength

50
Force ()

40

30

20

10

0
0 10 20
(d) Displacement(m)

Figure 3.17 Comparison of Direct FE2 (left) and full FE (right) cantilever
beam axial stress contours near the fixed end top left corner with(deformation
magnified 100 times) (a) 2c debond strength at 18 μm free end deflection ; (b)
c debond strength at 6 µm free end deflection before debonding; (c) c debond
strength after debonding at 9µm free end deflection and (d) Force deflection of
a beam (Figure 3.4(b)) with fibre-matrix debonding.

77
The properties used for the cohesive interaction is given in the Table 3.6 and

the matrix and fibre properties from Table 3.1.

The beam is deflected transversely, and the force deflection and stress contours

are observed in Figure 3.17. The FE2 model uses quadratic macroscale elements.

The stresses in the FE2 model and full FE model are very similar before

debonding sets in the models (Figure 3.17(b)). The debonding occurs at

different instances in FE2 and full FE models as can be seen in Figure

3.17(a).For a debonding strength of c =39.1 MPa, the debonding occurs at 9µm

free end deflection for FE2 model, while in the full FE model the debonding

occurs at 12µm free end deflection.

The point of debonding will be sensitive to the number of RVEs placed in the

macroscale element. The number of RVEs can be increased by refining the

macroscale mesh or increasing the number of Gauss points in the existing

macroscale mesh.

This difference between FE and FE2 models can be attributed to the weighting

of the microscale RVEs in the FE2 model. The FE2 models will be able to

capture uniform deformations-elastic or otherwise until localisation occurs as

seen in section 3.2.3. FE2 macroscale elements cannot capture localized

debonding since a debond on a FE2 RVE means that all the RVEs in that

macroscale integration volume region in the full FE model have debonded. This

is why refining the macroscale mesh will improve results given appropriate

shape functions are used. Alternatively, these debonding regions can be

modelled in full FE to ensure appropriate modelling and model the rest of the

structure with FE2 till a suitable Direct FE2 implementation that can handle

78
localization is realized. An adaptive remeshing at the macroscale upon

debonding is a potential and exciting future direction to extend this method but

is out of scope for this dissertation.

Examples with irregular elements: benchmarking irregular

elements with full FE composite material model

Direct FE2 can also be implemented where the macroscale elements do not have

regular geometric shapes such as rectangles. The scaling for each microscale

RVE in irregular elements should be accounted for separately as per the

equation (2.13). Whereas, the scaling factors for all integration point RVE

inside a rectangular element are the same for 2×2 quadrature points, the scaling

factors are different for different quadrature points even within the same

element if it is not a rectangle. This is because the Jacobian in equation (2.13)

has a different value for each quadrature point.

A composite material with fibre and matrix properties given in Table 3.1 is used

to benchmark with full FE. The deformed FE and FE2 models are shown in

Figure 3.18. A tensile force (as shown in Figure 3.18 (b)) is applied to turbine

blade shaped 2D model with irregular elements with quadratic macroscale

elements accounting for geometric non-linearity.

The deformed full FE and FE2 stress contours shown in Figure 3.18 are in

agreement. The peak force-displacement behaviour of full FE and the FE2

model is similar with an error in the peak force of 0.032%. In case of higher

order curved boundary model regions meshed with higher order elements, the

inverse iso-parametric mapping can be carried out iteratively [124] to impose

the multi-point constraint equations (2.20), (2.23) and (2.24).

79
(a)

(b)

Figure 3.18 (a) Full FE model reference (b) FE2 model with quadratic
interpolated irregular 2D macroscale elements with PBCs on the RVE and
tensile displacement applied on the macroscale nodes.

Modelling laminates in 3D

In this section, the groundwork to model composite structures bottom-up from

their constituents will be laid and explored further in Chapter 5. Before

modelling structures, a Direct FE2 model of a laminate will be compared with

classical laminate theory [114] to validate the modelling methodology.

Modelling methodology

The model comprises of only the microscale RVEs meshed with C3D8 elements

[110] placed at the macroscale element integration points (gauss points) as


80
shown in Figure 3.19. A single RVE is modelled and placed at the appropriate

locations with the help of a python script (2D example given in Appendix D).

Different RVEs can be placed at different locations if required. The RVEs are

oriented as per the angle of the ply they belong to. This means any laminate

(with any stacking sequence) can be modelled with the same RVE designed for

a unidirectional (0°) ply as demonstrated in Figure 3.19.Custom designed

laminate RVEs as per stacking sequence shown in Figure 3.20 will not be

required for modelling laminates with this method.

Figure 3.19 RVEs placed as per ply orientation-RVE white region represent
the fibres and green region represent the matrix.

(a) (b)

Figure 3.20 Conventional modelling of laminates from constituents involves


laminate specific RVEs for multi directional laminates for (a) 0°/90° laminate
and (b) [+55°/-55°] laminate.
81
The surface nodes of each RVE are connected to the macroscale element nodes

with linear interpolation functions employing multi-point constraints (MPCs)

similar to the 2D models, which is also carried out using a python script (2D

example given in Appendix D). The interpolation functions plays the role of

scale transition.

The FE2 laminate modelling comprises of 2 steps- (i) building the macroscale

model with ply oriented RVEs at the macroscale integration points and

(ii)applying MPCs to connect the surface nodes of the RVEs with the

macroscale nodes with linear interpolation functions. The next section describes

the validation of the Direct FE2 laminate modelling methodology with the

classical laminate theory (CLT).

Benchmarking a quasi-isotropic FE2 laminate model with CLT

The modelled quasi-isotropic laminate [0°/45°/-45°/90°]s has 0.1 mm thick plies

making the laminate 0.8 mm thick. The FE2 laminate is modelled by placing the

RVEs as per orientation of the ply, which the macroscale element belongs to

and connecting the RVEs with the macroscale nodes as per equations 2.23 and

2.24. The quasi isotropic laminate is built with the RVE shown below. A coarse

meshed RVE is used to reduce computation time. The RVE has a fibre volume

fraction of 0.6 and fibre diameter of 15µm.

The laminate is validated with classical laminate theory (CLT) for three loading

cases- normal strains εxx=0.02 and εyy=0.02; and shear strain εxy=0.04. Each case

is compared with CLT calculations and the stresses in the local ply directions

are tabulated below.

82
Figure 3.21 RVE used for modelling the quasi-isotropic laminate. RVE white
region represent the fibres and green region represent the matrix. A coarse
mesh is used to reduce computation time.

The material properties used for modelling the laminate is given in the table

below. Both the fibre and matrix region of the RVE are given the properties in

Table 3.7, which are the same inputs for CLT calculations.

Table 3.7 Elastic properties of the carbon fibre ply – T300-BLS914C

E11 E22 E33 ν12 ν13 ν23 G12 G13 G23


(MPa) (MPa) (MPa) (MPa) (MPa) (MPa)
138100 11000 11000 0.2774 0.2774 0.539 5438 5438 3527

σ11, σ22 and σ12, the normal stresses in the x,y and shear stress in xy plane in the

local ply coordinate system respectively, are calculated by volume averaging in

the FE2 model for RVEs of each ply individually and compared with the

corresponding CLT calculations. The micro stresses in each ply RVE are shown

in Figure 3.22 for εxx=0.02. The volume averaged stresses calculated (on the

RVEs in each ply) in the local ply co-ordinate system are given in Table 3.8,

Table 3.9 and Table 3.10 for the three cases.

83
Figure 3.22 The local stresses in each ply when the laminate is loaded in the x
–direction, εxx=0.02 . Each RVE’s fibre and matrix regions are given ply
properties in Table 3 7. The laminate global coordinate system is shown in
bottom left hand corner.

Table 3.8 Comparison of FE2 model and CLT solutions for global laminate
normal strain εxx=0.02

FE2 model (local stresses) CLT (Local stresses)

Ply S11 S22 S12 S11 S22 S12

0° 2779.04 61.40 0.00 2779.03 61.40 0.00

45° 1420.23 141.38 -108.76 1420.22 141.38 -108.76

-45° 1420.23 141.38 108.76 1420.22 141.38 108.76

90° 61.40 221.36 0.00 61.40 221.36 0.00

90° 61.40 221.36 0.00 61.40 221.36 0.00

-45° 1420.23 141.38 108.76 1420.22 141.38 108.76

45° 1420.23 141.38 -108.76 1420.22 141.38 -108.76

0° 2779.04 61.40 0.00 2779.03 61.40 0.00

84
Table 3.9 Comparing FE2 calculated stresses with CLT predictions for global
laminate normal strain εyy=0.02

FE2 model (local stresses) CLT (Local stresses)

Ply S11 S22 S12 S11 S22 S12

0° 61.40 221.36 0.00 61.40 221.36 0.00

45° 1420.23 141.38 108.76 1420.22 141.38 108.76

-45° 1420.23 141.38 -108.7 1420.22 141.38 -108.7

90° 2779.04 61.40 0.00 2779.03 61.40 0.00

90° 2779.04 61.40 0.00 2779.03 61.40 0.00

-45° 1420.23 141.38 -108.7 1420.22 141.38 -108.7

45° 1420.23 141.38 108.76 1420.22 141.38 108.76

0° 61.40 221.36 0.00 61.40 221.36 0.00

Table 3.10 Comparing FE2 calculated stresses with CLT predictions for global
laminate normal strain εxy=0.02

FE2 model (local stresses) CLT (Local stresses)

Ply S11 S22 S12 S11 S22 S12

0° 0.00 0.00 217.52 0.00 0.00 217.52

45° 2717.64 -159.95 0.00 2717.63 -159.95 0.00

-45° -2717.64 159.95 0.00 -2717.63 159.95 0.00

90° 0.00 0.00 -217.52 0.00 0.00 -217.52

90° 0.00 0.00 -217.52 0.00 0.00 -217.52

-45° -2717.64 159.95 0.00 -2717.63 159.95 0.00

45° 2717.64 -159.95 0.00 2717.63 -159.95 0.00

0° 0.00 0.00 217.52 0.00 0.00 217.52

85
The difference between the stresses calculated using CLT and the FE2 models

are negligible when both the CLT and the FE2 laminate model are given the

same ply properties in Table 3.7. The orientation data of the ply in the FE2 model

is obtained from the orientation of the RVE and no ply orientation data is given

explicitly as separate input to the FE2 model.

Comparing the homogenized elastic properties of a single 3D

macroscale element with a single RVE

Table 3.11 Material properties of the fibre and matrix used for 3D single
element and pipe structure (section 3.6.4)

Material Young’s Modulus, E(MPa) Poisson’s ratio, ν


Epoxy matrix 3350 0.35
Carbon Fibre 74000 0.2

Table 3.12 Homogenized elastic properties of the RVEs and FE2 macroscale
elements

Elastic Moduli RVE FE2 Difference


(MPa) macroscale (%)
element
E11 45602.17 45542.85 -0.13008
E22 14833.30 14830.46 -0.01915
E33 14833.30 14830.36 -0.01982
ν12 0.25 0.25
ν13 0.25 0.25
ν23 0.26 0.26
G12 4556.54 4558.78 0.04916
G13 4556.54 4558.76 0.048721
G23 3246.399 3247.62 0.037611

86
The elastic constants of a single macroscale element and single RVE are

calculated by homogenization. The RVE in Figure 3.23 (c) is used. The fibre

and matrix regions of each RVE are given their respective constituent properties

in Table 3.11.

(a)

(b) (c)

Figure 3.23 (a) Shear stresses in the FE2 macroscale element when sheared in
the YZ plane (plane of this page and the fibre direction is x pointing out of the
plane of the page); (b) Shear stresses in the single RVE when sheared in the
YZ plane; (c) RVE mesh used for the calculations.

87
As can be seen from Table 3.12, the elastic properties of the FE2 macroscale

element and single RVE differ by less than 0.14%. A single FE2 macroscale

element can be considered as a coarse FE element in which the constitutive

relations emerge naturally from its RVE microstructures placed at the

macroscale integration points.

Modelling a tube structure from constituent bottom-up with Direct FE2

Each ply is modelled with RVEs that are placed as per the ply angle of the

laminate specification. Filament wound pipe with stacking sequence [+55/-55]8

configuration is modelled as concentric cylindrical plies. The full FE laminate

model are given the homogenized properties of the constituents - glass fibres

and epoxy given in Table 3.12.

(a) (b)

Figure 3.24 Scaling in 3D Direct FE2 pipe model with (a) cubic RVE in Figure
3.21 with fibre diameter of 15 µm and (b) cubic RVE volume scaled so that
volume of the RVE is same as macroscale integration point volume with
scaled fibre diameter of 1260 µm to satisfy equation (2.19).

The fibre and matrix regions of each RVE are given their respective constituent

properties in Table 3.11. The RVEs can be modelled as per their real/true

88
physical size and scaling of the forces can be carried out in post processing as

per equation (2.13). In 2D, this scaling was handled by assigning a higher

thickness to satisfy that equation. In 3D, as stated earlier, the results (forces) can

be scaled in the post-processing stage after modelling the micro-structure with

their true length scales or instead the constitutive properties can be scaled.

Alternatively, the RVEs can be blown up so that equation (2.19) is satisfied and

the forces need not be scaled in post processing. If the length of the RVE mesh

is scaled by k in all 3 dimensions to satisfy equation (2.19), then the RHS of

equations ((2.20) to (2.27)) must also be scaled by k.

(a) (b)

Figure 3.25 RVE used in the elastic pipe simulations with (a) coarse mesh; (b)
fine mesh.

In this chapter and the subsequent chapter, all the 3D FE2 models presented and

discussed will involve RVEs that are scaled up in volume such that the micro-

structures retain their morphology, but are dilated in volume so that it is equal

to the volume of the macroscale integration point- the RVE is representing as

observed in Figure 3.24. The model in Figure 3.24 (a) is built with a coarse

meshed RVE with fibre diameter of 15 µm shown in Figure 3.21 and that in

Figure 3.24 (b) has the same RVEs scaled as per the equation (2.19). The RVEs

will be interpenetrating each other but will not interact with one another as

89
contact interactions are not accounted for. The model in Figure 3.24 (b) is only

used with scaled RVEs with meshes shown in the Figure 3.25.

(a)
15

10
Torque (kNm)

FE2 with constituent properties-coarse mesh


FE2 with homogenized properties-coarse mesh
FE2 with constituent properties-fine mesh
Full FE model
0
0.000 0.005 0.010 0.015
(b) Shear strain

Figure 3.26 (a) Boundary conditions for the pipe section twisting; (b) Elastic
behaviour of pipe structure: Comparing full FE model and FE2 model.

The pipe has a mean inner diameter of 52.43 mm and mean outer diameter of

57.93 mm (with a stacking sequence of [+55/-55]8). The boundary conditions

90
are shown on the full FE model and the FE2 model are shown in the Figure

3.26(a). A torque is applied on the nodes of the annular positive (+) z face.

When both the FE2 model and full FE model are given the same input, that is

when both the fibre and matrix regions are assigned the homogenized ply

properties, the FE2 model behaves very similarly to the full FE model as can be

seen in Figure 3.26(b). When given the respective constituent properties, the

FE2 model with coarse meshed RVEs (in Figure 3.25 (a)) is more compliant than

the full FE model. As we refine the mesh of the RVE as shown in Figure 3.25

(b), the results of the FE2 is closer to the full FE model with homogenized ply

properties. This implies there is a potential and possibility of using this method

to model structures from constituent level upwards without having to

characterize the intermediate ply as hinted in Chapter 1. This will be explored

further in chapter 5.

Summary

In this chapter it is demonstrated that the Direct FE2 implementation can be

extended to incorporate geometric non-linearity, material non-linearity and

viscoelasticity without any difficulty and can produce similar macroscopic and

microscopic results to computationally expensive brute force FE models.

Higher order interpolation functions improve the FE2 model performance as the

macroscopic interpolation functions serve as the scale transition relations,

which implies that selection of appropriate interpolation function is critical. The

method has also been shown to be able to operate with irregular shaped elements

and cohesive interactions to model interfaces – up to the point of localised

debonding.

91
A method to model 3D laminates, with the Direct FE2, bottom-up from its

constituents has been proposed and validated with classical laminate theory

(CLT). To the best of the author’s knowledge, nobody has proposed such an

approach for modelling laminates from the constituent level. One huge

modelling advantage is that there would be no need to build laminate specific

RVEs as carried out conventionally.

The groundwork for modelling composite structures from their constituents has

been laid and demonstration of the method to capture the elastic behaviour of

structural component such as a pipe has been shown in section 3.6.4 in which

the FE2 model has been shown to be equivalent to the full FE model. As has

been highlighted in the introduction, predicting failure of heterogeneous

structures from constituents is difficult but a blooming research field and in the

following chapters, some composite pipe experiments carried out on Stewart

platform-based setup and an attempt to model the experiments with this Direct

FE2 laminate modelling methodology is discussed.

92
Experiments on composite structures: Glass

Reinforced Epoxy (GRE) pipes under tension, torsion and

bending loads

This chapter is a diversion from the Direct FE2 method, which describes the

experiments performed on GRE composite pipes carried out using a Stewart

platform-based test-set up at Hamburg University of Technology, Germany.

The platform allows load application in three axes (x,y,z) and three rotations

(Ɵx, Ɵy and Ɵz).

The tests were carried out as part of a larger effort to determine the performance

of underground GRE pipe networks, which are typically used for fuel transport.

Other than gaining valuable hands-on experience in composites testing and test

design, the tests presented an opportunity to understand the failure of an actual

composite structure under various loading scenarios. The test data will be a

source of reference when FE2 is implemented in 3D in the next chapter.

The following sections describes the experiments conducted on filament-wound

GRE pipes to study and compare the different failure modes that occur upon

tension, bending and torsion loads on the pipes. Different failure mechanisms,

characteristic of various loadings are identified from the damaged surfaces. The

pipe specifications, test set-up, specimen preparation, loading details and the

experimental measurements and observations are also discussed here.

Pipe specifications and material details

Manufacturer specifications of the GRE pipe used in this study are given in

Table 4.1.
93
Table 4.1 GRE pipe specifications

Manufacturer Ameron
Model Bondstrand
Inner diameter (mm) 104±1
Thickness (mm) 5.5±0.5
Longitudinal Modulus (GPa) 11.1
Hoop Modulus (GPa) 25.2
Longitudinal Strength (MPa) 59
Hoop Strength (MPa) 165

The pipe inner surface and thickness are shown in Figure 4.1 . The filament

winding direction is clearly visible, and the ply interfaces can be observed. The

laminate stacking sequence is determined to be [+55/-55]16. The glass fibres

and matrix material specifications are not mentioned by the pipe manufacturer.

The fibre diameter is observed to be around 15 μm when measured under the

microscope and the fibre volume fraction is determined to be between 0.5 and

0.6 by doing simple image processing on the micrographs.

Figure 4.1 Composite pipe coupon section: (a) inner diametrical surface of the
pipe- +55°/-55° filament windings are clearly visible; (b) through thickness
section view which shows the ply interfaces.

A brief review of work carried out in the past

Considerable research efforts have been directed towards the study of GRE pipe

failure analysis for identifying and recording failure mechanisms

experimentally. The failure mechanisms, identified from micrographs of failed

filament-wound composite tubes under combined loadings, are found to be

94
Table 4.2 Past research in GRE pipe failure analysis

Pipe layup Inner dia* Loading Objective Significant findings

(mm) condition

David Wall Jr, and Multiple 76 Torsion Comparison of experimental Experimental strengths correlate with yield
Michael Card [128] stacking strengths with torsional predictions when unrealistically large transverse
(1971) sequences buckling predictions and tensile and shear strength of UD* plies are used.
material strength predictions

Highton[130] ±75° 100 Combinations of Tested 50 specimens to obtain Predictions with non-linear in-plane shear properties
(1985) internal pressure experimental failure envelope and residual thermal stresses agree reasonably with
and axial experimental predictions
tensile/compress
ive loads

Mistry [131] (1992) ±55° 100, External Studied collapse behaviour of Numerical predictions agree with experimental
pressure, Axial filament-wound tubes. FPF* measurements for buckled specimens. When FPF
200 compression and Buckling loads predicted occurs before buckling, the residual strength prevented
with special purpose FE* buckling
program.

95
Soden[132] ±45°, ±55°, 100 Internal pressure Studied leakage and fracture Fracture and leakage vary significantly with ratio of
±75° and axial strength applied hoop to axial stresses
(1993) tension/compres
sion

Zhao and Pang[129] [90/0]n, 98.5 Torsion Analytical and experimental Failure analysis based on maximum strain failure
(1995) [±60]n, study to investigate elastic and criteria deviate from the experimental results
[±45]n failure behaviour

Bai[127,133,134] ±55° 60 Internal pressure Damage envelope predictions Micro-cracking and delamination are main failure
(1998) and axial with micromechanical modes based on loading conditions.
tension/compres modelling were in good
sion and agreement with microscopy Predicted FPF loads were underestimated for the
combined observations. pressure dominated loading
loading

Rousseau [±556] 60 Hoop:axial Studied the influence of Increasing the degree of weaving due to the presence
[135](1999) stress ratio (2:1) interweaving of the fibres of fibre crossovers, can induce premature weeping of
inside the helical wound layers the pipe for the case of close-ended internal pressure
on damage growth.

Beakou and 55° Pressure loaded Theoretically examined 55° as Concluded that the Young’s modulus of the matrix and
Mohamed [136] and axially the optimum winding angle transverse strength of the ply are the most important
(2001) loaded pressure using CLT* for stress analysis parameters that resulted in variations in optimum
vessels and Tsai-Wu failure criterion winding angle

96
Mertiny [137] [±603], 38.1 Biaxial tensile Investigated performance of Considering the [±603] as the baseline, the other
[±45, stress ratios multi-angle filament- wound multi-angle configurations showed improved strength
(2004) ±602], pipes under hoop: axial ratio of 2
[±30, ±602]

Meijer and Elliyin [±603] 50.8 Multi-axial Studied the stress and strain Mechanical properties predicted with CLT matched
[138] (2008) stress states failure envelopes for multi- with experimental observations. failure modes
axial stress states and identified -axial ten/comp structural failure, weepage,
categorised five failure modes local leakage, burst

A.E. Antoniou et al [±45]2 28 Combined Aimed to simulate complex The failure locus in the effective axial-shear stress
[139] torsion and stress states encountered in a plane was derived experimentally. The correlation
tension/compres wind turbine rotor blade established between the ratio of transverse normal and
(2009) sion in-plane shear stress in the principal coordinate ply
system and the elastic shear modulus, suggested a
strong dependence on the combined tube loading

Martins [140,141] [±454], 102 Internal pressure Progressive failure analysis in Both leakage and burst failure modes were observed
(2012,2013) [±554], FE employing a strain-based
[±604], continuum damage
[ ±754] formulation

*Dia: diameter; GRE: Glass Reinforced Epoxy; GRV: Glass Reinforced Vinyl Ester; GRP: Glass Reinforced Polyester; FE-Finite Element, FPF- First Ply Failure, LPF- Last Ply Failure, CLT- Classical Laminate
Theory, UD-Unidirectional

97
dependent on the applied stress ratio [125–127]. Some previous studies of GRE

pipe failure are summarized in Table 4.2. Extensive work has been carried out

in studying failure under the axial and internal pressure loading conditions

compared to torsional loading conditions [128,129].

Experiments

Hexapod test set-up

The test was conducted on a custom designed multi-purpose Hexapod testing

facility. The Hexapod test rig is based on the Stewart platform principle using

six hydraulic cylinders that move a 2.5-ton ring structure as shown in Figure

4.2.

Figure 4.2 Hexapod test rig at Hamburg University of Technology, Germany


with fixtures and test specimen.

This setup allows for both force and displacement-controlled loading in all

translational and rotational degrees of freedom (DOFs). The forces can be

98
measured using a customized six DOF load cell placed underneath the hexapod

ring. The load cell has a measuring range of 250 kN in axial direction, 150 kN

in lateral direction and 40 kNm in all rotatory DOFs. The loading setup is shown

in the Figure 4.2. The upper fixture plate is attached to the Hexapod ring, while

the lower fixture plate is mounted on the six DOF load cell. The specifications

of the Hexapod machine are given in the Table 4.3.

Table 4.3 Loading specifications for Hexapod test facility

Parameters Maximum Loading Limit

Combined Angles +/-10°


Combined Deflections +/-300 mm
Frequency Range up to 30 Hz
Forces up to 500 kN
Torques up to 40 kNm
Velocity up to 1 m/s

Specimen preparation and mounting

The clamping configuration is detailed in Figure 4.3. Pins had to be used to

transmit torque as suitable adhesive that could be used to transmit torque just

through the pipe surface could not be found. For the tests with torsional loading,

certain measures are adopted for the clamping configuration, to improve torque

transmission and to prevent premature loading pin-hole failure3 (28 mm

diameter pin as seen in Figure 4.3 (a)) that can occur in the clamping region

shown in Figure 4.3 (c). These measures include the use of three 10 mm pins

(as seen in Figure 4.3 (a)) and rubber sheets with different thickness during

3
Pinhole loading failure was observed during the preliminary experiments carried out to
evaluate the loading capacity of the fixture. Experimental results presented here did not have
this clamping failure due to the improved gripping of the extra pins and 3mm neoprene rubber
sheets

99
clamping. A 3 mm neoprene rubber sheet is placed between the mandrel and the

inner surface of the pipe, while a 4 mm neoprene rubber sheet is placed between

the outer surface of the pipe and the clamping collar (Figure 4.3(b)).

(a) (b)

(c)

Figure 4.3 (a) Additional pins used for transmission of torque and (b)
Neoprene rubber sheet cross-section configuration used for clamping (without
the pins) (c) Pin hole failure.

Figure 4.4 Strain gauge configuration A used for tension, bending and strain
gauge configuration B used for torsion.

100
The strains were measured using seven to eight linear strain gauges (HBM 1-

LY18-6/120) attached to each test specimen at different positions as shown in

Figure 4.4 at three height levels (¼ , ½, and ¾ of the pipe length). A (HBM 1-

LY68-6/120) strain rosette was also attached at the middle of the pipe (½ of the

pipe length).

The strain gauge configurations used for the tests depend on the loading

condition of the pipe specimen and are shown in Figure 4.4. For torsional tests,

strain gauges were placed along the fibre direction (55° to pipe axis) and

perpendicular to the fibre direction to measure torsional strain whereas strain

gauges for bending and axial tests were placed along cylindrical axial direction

of the pipe. The test specimens involving torsional load employ the strain gauge

Configuration B in Figure 4.4 and the Configuration A was used in the other

cases where no torsion is applied.

Loading details

The different loads on the pipe are shown in Figure 4.5. The loading conditions

and loading rates of each of the pipe are summarized in Table 4.4.

The bending tests were initially carried out with uniform bending moment in

the specimen. This was achieved by minimizing the transverse shear force that

was present when the top clamp was rotated to induce a bending moment. The

hexapod machine has an in-built feature with which a force (in a specific

direction) at a specific magnitude can be constrained. The hexapod machine can

automatically adjust the position of the specimen to meet the constraint. This

sometimes limits the machine displacement range and cannot be used in all

101
Figure 4.5 The different loading conditions applied on the pipe: Py is the axial
tensile force, Ty is the torsional moment, Mz is the bending moment and Fx is
the transverse shear force in the pipe.

cases. When it is not possible to reduce the shear to an acceptable level, shear

force compensation is carried out in the calculation of the bending moment.

The pipe could not be loaded in bending without shear force. Therefore, the

resulting moment M at the point of failure in the specimen is calculated with the

equation below when the shear force is present during bending.

M  M measured  P    Fx  L f

(4.1)

where M measured is the bending moment measured at the bottom of the pipe, P is

the tension; 𝐹𝑥 is the shear force acting in the pipe, 𝐿𝑓 is the height of the point

102
of failure from the bottom and 𝛿 is the lateral/transverse shift of the pipe axis

from initial machine axis at the point of failure.

Table 4.4 Experimental test details

Length 1st Loading


Test
(m) Loading
Test type No. Loading Type Loading Rate
Level
1* 0.75 Tension To Failure 1mm/min
Tension 2* 0.75 Tension To Failure 1mm/min
3 1.5 Tension To Failure 250mm/min
Bending 4** 1.5 Bending To failure 1.2o/min
` 5 1 Bending To Failure 1.2o/min
6 1 Torsion (–z) To Failure 247.8°/min
Torsion
7 1 Torsion To Failure 247.8°/min
The pipe specimens are 1.5m in length; * The pipe specimens are 0.75m in length,
** The pipe specimens are 1m in length to increase the torsional moment applied to the
specimen.

Observations

Table 4.5 Failure strengths of the pipe under various loading conditions

Failure strengths
Torsion
Tensile Bending Shear
Test Specimen shear
Length strength, axial stress stress,
No. No. strength
(m) (P/A) (My/I) (Fx/A)
(Tr/J)
MPa MPa MPa
MPa
1 0.75 58.68 - - -
2 Tension 0.75 59.26 - - -
3 1.5 62.56 - - -
4 1.5 - 86.19 - 4.67
Bending
5 1 - 93.95 - 9.86
6 1 - - -140.02 -
Torsion
7 1 - - 141.35 -

103
Damage in composites can be classified as intra-laminar (fibre failure, matrix

failure, fibre debond) and inter-laminar damage (such as delamination). The

various failure modes observed in the failed specimens, experimental

measurements and its analyses are described in this section. The failure loads

are listed in Table 4.5.

The stress-strain behavior of the pipes and the failed speximens are plotted with

the load tests in the subsequent sections. The stresses are calculated using the

mean value of inner and outer diameters of the pipe calculated from the pipe

samples. The stress-strain behaviour is plotted in this work rather than force-

displacement. The hexapod cross head displacement will include the machine

cross head deformation however small it is. Strain gauges record only the pipe

deformations, which will also prove useful to compare with simulations of the

experiments to be discussed in the next chapter. Similarly torsional shear stress

vs shear strain and axial stress due to bending moment vs axial strains plots will

be presented for the same reason stated here.

Tensile loading of the pipe specimens

One pipe of length 1.5m (referred to as long pipe) was tested on the hexapod.

Two pipes of 750mm length (referred to as short pipes from now) were tested

in tension on a Shimadzu universal testing machine equipped with a 250 kN

load cell. The test setup is shown in Figure 4.6.

The long pipe was tested at a higher loading rate of 250 mm/min to check for

effect of length and loading rate on failure strength. The strength increased

slightly by 6% for the longer specimen as can be calculated from strength values

listed in Table 4.5. It is unlikely that the strength increase is due to the length of

104
the pipe specimen as specimen size only increases the probability of failing

earlier due to the presence larger number of defects as pointed out by Griffith

[142]. The strength increase is most likely due to higher strain rate employed

rather than the increase in length. Also, the modulus at the higher strain rate is

almost identical to the modulus at lower strain rates. The effect of the strain rate

on modulus is not observed in this experiment.

Figure 4.6 In-house tensile tests setup with Shimadzu machine. (A) Pipe
specimen inserted into the mandrel assembly; (B) Collars clamped the pipe
specimen with rubber sheet; (C) Entire test setup onto the Shimazhu machine;
(D) Protective shield to prevent any flying chips and digital video recorder to
record the entire test.

For the long pipe tension test, the axial strains along the gauge length of the pipe

is observed to be uniform until the peak failure load as seen from the digital

image correlation (DIC) image of the specimen in Figure 4.7. This further

confirms the uniformity of the loading along the gauge length in the pipe while

loaded in tension.

105
(a) (b)

Figure 4.7 DIC image showing the uniformity of the tensile load throughout
the gauge length of the pipe (a) before peak failure load; (b) after peak failure
load with the formation of the strain localization bands along the filament
winding direction.

The tensile stress-strain behaviour is plotted in Figure 4.8(a). The tensile

behaviour of three pipes is observed to be consistent and the effect of loading

rate upon the elastic or failure behaviour is not easily observed in the stress

strain graph. It is difficult to ascertain the percentage difference as the strain

readings of the top and bottom strain gauges are slightly different as can be seen

in stress strain plot of Figure 4.8(a).

Fibre pull-out is visible in the final stages after the matrix failure. Figure 4.8 (b)

and (c) shows the short pipe before and after the failure. Figure 4.8 (d), (e) and

(f) shows matrix crazing–whitened bands along the regions of strain localization

before fibre pull out and fibre failure occurs.

106
70

60
Tensile stress (MPa)
50

40

30

20
short pipe-expt1
10 short pipe-expt2
long-pipe-bot-strain guage
long-pipe-top-strain guage
0
0.00 0.02
(a) Tensile strain ()

(b) (c) (d) (e)

(f)

Figure 4.8 (a) Stress strain behaviour of composite pipe in tensile load
(b) short pipe section before failure; (c) and (d) short pipe failed specimens;
(e) and (f) failed region in long pipe very similar to the damage pattern
observed in the short pipe test at a lower loading rate.

107
Torsion

The torsion tests were carried out on pipes of 1 m length with a gauge length of

500 mm on the hexapod setup. The strain gauges were applied as per

configuration B in Figure 4.4. Initially, having faced difficulties in torqueing

the pipe at low twisting rates and due to the limitation of rotation of hexapod

cross-head, 1 m long pipes with a gauge length of 500 mm was tested at high

rate of 247.8°/min, which is equivalent to a loading rate of 250 mm/min. For

the same twisting angle θ, a long pipe was strained (torsionally sheared) less

compared to a short pipe (γ=rθ/L).

The shear stress is calculated using the expression Tr/J, where T is the torsional

moment, r is the mean pipe diameter and J is the polar moment of inertia. The

shear strain in the pipe is calculated from equation 4.2 with strains measured

using the strain rosette (schematic in Figure 4.9) glued at the centre of the pipe

as depicted in pipe strain gauge configuration B in Figure 4.4 .

Figure 4.9 Strain rosette with strain gauge P, Q and R; θ=45°; ϕ=45°.

The shear strain (γ) is calculated using the following expression:

𝛾 = 2𝜀𝑄 − 𝜀𝑃 − 𝜀𝑅

(4.2)

108
where εQ, εP and εR are the strain measured by strain gauges P Q and R

respectively.

Figure 4.10 Torsional shear stress vs shear strain of the pipes.

The pipe’s behaviour is plotted in Figure 4.10. The failed pipe specimens are

shown in Figure 4.11 (a), (b) and (c).

When torqued in the positive (+) z direction (top circumferential face twisted

anti-clockwise), the outermost ply was in compression, which can be clearly

seen from the local ply buckling visible on the outside in Figure 4.11(b). The

pipes shown in Figure 4.11(a) have the outermost plies in tension and the inner

plies can be observed to buckle inside (Figure 4.11(c)).The dominant failure

mode in torsion appears to be controlled by the failure of the fibres- fibre tension

failure with fibre pull out (in Figure 4.11(a)) and local ply buckling that most

109
likely is initiated by fibre compression triggered by fibre microbuckling as can

be seen in Figure 4.11(b) and (c).

Figure 4.11 Torsion pipe failure modes: external visible failure dependent on
the relative direction between the torque and state of stress in the outermost
ply; pipes in (a) test 6 were twisted in the –z direction in which the outermost
ply was in tension; (b) failed specimen in test 7 twisted in +z direction in
which outermost ply was in compression (c) local ply buckling failure of the
inner plies observed inside the torqued (-z) pipe.

Bending

The bending tests were conducted on two pipes - one of 1.5 m length and another

of 1 m length on the hexapod setup. This was carried out to see whether length

affects the bending failure strength. The strain gauges were glued as per

110
configuration A shown in Figure 4.4. The pipe were loaded at a rate of 1.2°/min

till the failure.

These pipes tested with bending moment were failed in the presence of a shear

force as machine displacement limitations prevented from failing the pipe with

a pure bending moment. Due to the presence of shear force, a gradient of

bending moment was present along the length. This can be seen from the strains

measured using the strain gauges along the gauge length of the pipe as a drop in

the strain readings can be observed. The strains vs the rotation along the tensile

side of the pipe is shown in Figure 4.12. The top strain gauge reading is negative

because the effective bending moment is in the opposite direction of the applied

bending moment.

0.016
0.014
0.012
0.010
Longitudinal Strains

0.008
0.006
0.004
0.002
0.000
0 2 4 6 8
-0.002 Rotation of top clamp(°)
-0.004 top- tensile side
center-tensile side
-0.006 bottom-tensile side

Figure 4.12 Axial strains recorded along the gauge length of the tensile side of
pipe during the bending test 4 (in Tables 4.4 and 4.5).

111
The longitudinal stress vs the axial strain is shown in Figure 4.13. The moments

used to calculate the failure stress is determined with equation 4.1. The

longitudinal bending stress is calculated from the expression σb=My/I, where σb

is the axial bending stress, y is the perpendicular distance from the neutral plane,

and I is the second moment of inertia with respect to the bending axis. The

strains are plotted with the strain gauges close to the region of failure, which are

the axial strain gauges close to the bottom clamp. The moment at the bottom

most strain gauge along the gauge length of the pipe is used to calculate the

stress plotted in Figure 4.13.

100
Bending longitudinal stress (MPa)

80

60

40

20
1m pipe(Test 5) at 0.25m from pipe bottom
1.5m pipe(Test 4) at 0.375m from pipe bottom
1.5m pipe(Test 4)- stress at 0.25m vs strain at 0.375m
0
0.000 0.002 0.004 0.006 0.008 0.010 0.012 0.014 0.016
Strain

Figure 4.13 Bending stress vs strain (longitudinal direction).

112
The failed specimens are displayed in Figure 4.14. The failed specimens initially

show traces of matrix crazing as visible in Figure 4.14 (a) on the tensile side

(similar to the one observed in the tensile tests).

(a) (b) (c)

(d) (e)

Figure 4.14 Failed regions upon bending by failure in test 4 on (a) tensile side,
(b) and (c) compressive side; failed regions in test 5 (d) matrix compression on
the compressive side and (e) delamination upon ultimate failure.

The bending failure stress is in between tensile (60 MPa as can be seen from the

graph) and compressive strengths of the pipe (143 MPa which is determined

from coupon tests4 ). This implies that the failure of the pipe is due to the shear

force rather than the bending. The failure is due to shear rather than bending

moment is further supported by the fact that the shorter pipe failed with a higher

shear force compared to the longer pipe.

4
(compressive strength determined separately from coupon tests[151]). In the same paper, the
coupon tensile strength is found to be 50MPa as opposed to the 60MPa from the pipe tests
presented here. This may be due to free edge effects (edges created while machining coupon
from pipe) in the coupons as per ASTM D3039/D 3039M standard or due to inconsistent
manufactured pipe samples.

113
Summary

The experimental study of the failure of Glass Reinforced Epoxy (GRE)

composites pipes under tension, torsion and bending have been described in this

chapter. In the conducted tests the failure mechanisms for different loading

conditions have been identified and described. It is noted that the number of

tests conducted is limited due to the cost involved in shipping the specimens

and fees involved in using the hexapod tester.

Conducting these tests presented a challenge since not many equipment are

capable of conducting such pipe experiments. There are no standards available

for reference to the best of the author’s knowledge for such large-scale specimen

testing. However, the tests proved to be a valuable experience in jig and fixture

design and training to use the hexapod for pipe testing. Such tests are new to

TUHH also. The failure mechanisms identified are summarized here.

For pipes loaded axially in tension, matrix crazing along the filament winding

direction precedes final failure, which is due to fibre tensile failure along with

fibre pull-out.

Fibre failure appears to be the dominant mode of failure in torsion. Both local

ply buckling and fibre pull-out is visible in pipes damaged by torsion. The

direction of applied torque is readily identifiable by observing the external

surfaces of the failed pipe sections. Distinct failure signature is observed, which

depends on the direction of applied torque and stacking layup. This can be

useful in failure investigations of a pipe network.

114
The pipes could not be failed by applying a pure bending moment. Therefore, a

transverse shear force was present during failure. This makes it difficult to

determine the individual contribution of the shear force and bending moment to

failure. The failure in bending also appears to begin in the matrix as evident

from the extensive white bands due to crazing. However final failure is

characterized by fibre pull out and delamination. This is not surprising since in

bending, one side of the pipe is in tension and other is in compression and

similar failure mechanisms being observed is expected.

115
Direct FE2 analysis of composite structures –

simulation of GRE composite pipe experiments

This chapter will discuss the modelling of the experiments performed on glass

reinforced epoxy (GRE) composite pipes described in Chapter 4. The Direct

FE2 method used to analyse the pipe experiments might be able to help us co-

relate structural behaviour to micro-structural behaviour as the micro-structures

are modelled. Tension, bending and torsion experiments are numerically

modelled with the Direct FE2 laminate modelling methodology described in

section 3.6. This is performed to determine the method’s ability and

shortcomings in applying it to predict structural behaviour up to the point where

strain/damage localization occurs in the GRE composite pipes tested. The

Direct FE2 model of the pipe, boundary conditions and the observations made

from the FE2 simulations are discussed.

Modelling of the composite pipes

The pipes tested were 0.75 m to 1.5 m in length and it would be preferable to

model the whole pipe. But due to computational considerations only 10 mm of

the pipe is modelled as can be seen in Figure 5.1(d). It is similar to the model

described in section 3.6.4. But here the constituents are given non-linear

material properties to model the uniform inelastic behaviour before failure

localization.

Each ply is modelled with RVEs of GRE with a 0.6 fibre volume fraction. The

RVEs are placed as per the ply angle of the laminate specification. Filament

116
wound pipe with stacking sequence [+55/-55]8 is modelled as concentric

cylindrical plies as described in section 3.6.4.

All the 3D FE2 models presented and discussed will involve RVEs that are

scaled up in volume such that the micro-structures retain their morphology, but

are dilated in volume so that it is equal to the volume of the macroscale

integration point to satisfy equation (2.19) - as can be observed in Figure 5.1(d).

Details for this volume scaling have been discussed in Section 3.6.4. The true

fibre diameter is 15 µm, which after volume scaling to satisfy equation (2.19),

becomes 1260 µm. The RVE mesh in Figure 5.1 (e) is referred to as coarse mesh

RVE and Figure 5.1 (f) is referred to as fine meshed RVE from now on.

Figure 5.1 FE models of pipe at: (a) macroscale level, (b) ply-level, (c)
constituent level, (d) Direct FE2 pipe model, (e) coarse mesh RVE and (f) fine
mesh RVE.

117
Constitutive modelling of fibre and matrix

Some researchers have determined the constituent strengths from the ply

strengths to replicate failure envelopes of individual plies in single RVE

micromechanical models [143,144]. This is justified by claims that the in-situ

behaviour of the constituents is different from the available data from the

experiments performed on the bulk constituents.

The fibre is modelled as linear elastic with properties given in Table 5.1. The

fibre failure is modelled with a maximum strain criterion in tension and

compression. Upon reaching a critical strain (in Table 5.1), the Young’s

modulus of the fibre is switched to zero using a user defined subroutine

USDFLD. The material elastic properties are given in Table 5.1.

Table 5.1 Material elastic properties [145]

Material Young’s Poisson’s Tensile Tensile Compressive Compressive


Modulus, ratio, ν Strength failure Strength failure strain
E(MPa) (MPa) Strain (MPa) (%)
(%)

Glass
74000 0.2 2150 2.905 1450 1.959
Fibre

The matrix inelasticity is modelled with a Mohr-Coulomb model used in

micromechanical modelling [123,146] of composites. The Mohr-Coulomb

model in addition to the elastic properties requires the dilation angle, friction

angle and cohesive strength. The Mohr-Coulomb model has been used to model

brittle polymers like epoxies [123,146]. The Mohr-Coulomb criterion states that

yielding will occur on a given plane when the shear stress (τ) exceeds the

118
cohesive stress of the material plus the frictional force acting along failure plane

such that:

𝜏 = 𝑐 − 𝜎𝑛 𝑡𝑎𝑛𝜙

(5.1)

Here c is the cohesion yield stress, 𝜎𝑛 is the normal stress acting on the failure

plane and ϕ is the angle of friction. In Mohr’s stress plane (σ-τ), the yield surface

can be expressed in terms of maximum and minimum principal stresses:

𝑓(𝜎1 , 𝜎3 ) = (𝜎1 − 𝜎3 ) + (𝜎1 + 𝜎3 )𝑠𝑖𝑛𝜙 − 2𝑐 𝑐𝑜𝑠𝜙 = 0

(5.2)

The angle of internal friction and the cohesion stress can be related to the tensile

and compressive strengths of the materials through the following expressions:

𝑐𝑜𝑠𝜙
𝜎𝑡 = 2𝑐
1 + 𝑠𝑖𝑛𝜙

(5.3)

𝑐𝑜𝑠𝜙
𝜎𝑐 = 2𝑐
1 − 𝑠𝑖𝑛𝜙

(5.4)

The epoxy matrix(E = 3350MPa,ν = 0.35 [111])is assigned a dilation angle of

0.1 and the friction angle of 15° and a cohesive strength of 39.1 MPa which is

equivalent to a tensile strength of 60 MPa and a compressive strength of 101

MPa [123].

119
Tension

The boundary conditions for modelling the tension experiments is shown in

Figure 5.2. A displacement along the positive z direction is given on the annular

face of the model until the onset of damage localization was detected in the

model. The analysis was performed with a Direct FE2 model with coarse meshed

RVEs and another model with fine meshed RVEs shown in Figure 5.1 (e) and

(f) respectively.

Figure 5.2 Boundary conditions for tension pipe test on the FE2 macroscale
elements.

The tensile force vs tensile strain behaviour of the Direct FE2 model is shown

in Figure 5.3. The experimental measurements and the FE2 model results agree

reasonably well. The elastic modulus of the composite pipe model would be

affected by the fibre volume fraction of the RVE. The volume fraction of 0.6

used in the RVE is the upper bound for the pipe as was determined from a simple

image processing of the micrographs of the pipe sample. The FE2 model with

fine mesh RVEs is observed to be slightly stiffer initially, which was also

observed in the Direct FE2 2D models discussed in section 3.3. Failure occurs

120
in the matrix and is controlled by the matrix tensile strength. The fibres were

not loaded close to their critical strains or stresses listed in Table 5.1 at failure.

70

60
Tensile stress (MPa)

50

40

30

20 short pipe-expt1(Test 1)
short pipe-expt2(Test 2)
long-pipe-bot-strain guage(Test 3)
10 long-pipe-top-strain guage(Test 3)
FE2 -coarse mesh RVEs
FE2 -fine mesh RVEs
0
0.000 0.005 0.010 0.015 0.020 0.025
Tensile strain ()

Figure 5.3 Tensile stress vs tensile strain: Comparing FE2 model response with
pipe tensile experiments.

The RVEs’ orientations are as per the ply orientations which are either +55° or

-55°. The outermost ply is at the bottom in each of the images in Figure 5.4. The

axial stress contours are shown in Figure 5.4 (a) and (b). The axial stress in the

fibres and matrix before plastic strain sets is shown in Figure 5.4 (a). They are

uniform, whereas after matrix plasticity sets in, the stress in the fibres become

non-uniform as can be seen in Figure 5.4 (b). The RVEs in both the +55° and -

55° plies are in tension which can be seen in the stress contours shown in Figure

5.4 (a) and (b). The equivalent plastic strain contours before and after the matrix

yields in the RVEs are shown in Figure 5.4 (c) and (d). The equivalent plastic
1
strain(𝜀̅𝑝𝑙 ) is calculated as 𝜀̅𝑝𝑙 = ∫ 𝑐 𝜎: 𝑑 𝜀 𝑝𝑙 , where 𝜎 is the stress tensor, 𝜀 𝑝𝑙

121
is the plastic strain tensor and c is cohesion yield stress for the Mohr-Coulomb

plasticity model[110].

Figure 5.4 Axial Stress in the fibres and matrix of the +55° and -55° plies at
(a) 0.0025 axial strain(25 µm axial displacement) (b) 0.01 axial strain(100 µm
axial displacement) and (b) equivalent plastic strain contours in the fine
meshed RVEs in tension at (c) 0.0025 axial strain(25 µm axial displacement)
and (d) 0.01 axial strain(100µm axial displacement).

Torsion

Boundary conditions for modelling the torsion experiments is shown in Figure

5.5 and the pipe experimental behaviour is compared with the model predictions

122
in Figure 5.6. A torsional moment is applied on the positive z face and the

negative z face displacement is completely constrained in x, y and z directions.

Figure 5.5 Boundary conditions for torsion pipe test on the FE2 macroscale
elements.

The torsional elastic modulus obtained from the coarse mesh differs by 5%

relative to the fine mesh models. The fine mesh model is closer to the

experimental elastic modulus compared to the coarse mesh model. This trend of

fine mesh Direct FE2 models being slightly stiffer than coarse mesh Direct FE2

models was also pointed out in Chapter 3 and in FE2 tensile simulations

discussed in section 5.2. The slight non-linearity observed in the experiments is

also captured by the Direct FE2 predictions.

It is also observed that the FE2 model behaviour in positive and negative torsion

are the same as observed in the experiments. The outermost ply RVE is the one

on the bottom left in all the images in Figure 5.7. The outermost ply is in tension

when torqued negatively (clockwise) as can be seen in Figure 5.7 (c) and in

123
compression when torqued positively (anti-clockwise) in Figure 5.7 (d), which

confirms that the orientations in the Direct FE2 pipe model are the same as the

experimental specimen. Another point to note is that the matrix in the plies with

fibres in tension (red in S33 contours) do not undergo plastic strain (and are in

compression) whereas the matrix in plies with fibres in compression undergo

plastic strain and are in tension. This detailed stress state of the material phases

are observed as a result of the fibre and matrix being explicitly modelled in FE2,

i.e., the stress state of the limiting constituent can be determined, which is

essential to tune structural behaviour with microstructure. It should also be

noted that in contrast to a single stress state at a point in homogenized ply

models, it is possible to have tensile stress in fibre and compressive stress in

matrix or vice versa only in models that account for microstructure explicitly.

200

180

160
Shear Stress (MPa)

140

120

100

80

60
Test 6(-z Torque)
40 Test 7(+z Torque)
FE2 fine mesh(+z torque)
20 FE2 coarse mesh(-z torque)
FE2 coarse mesh(+z torque)
0
0.00 0.01 0.02
Shear Strain

Figure 5.6 Stress- strain behavior of pipe in torsion: Comparing FE2 model
simulation results with experiments.

124
Figure 5.7 All contours are at macroscale shear strain of 0.01346. Plastic stress
contours when torqued in (a) -z direction and (b) +z direction. The
corresponding axial stresses in the (c) outermost ply fibres are in tension when
torqued in -z direction and the (d) outermost ply (bottom left) fibres are in
compression when torqued in the +z direction. (e) Tensile stresses in the
matrix and the plies with fibres in compression, which causes the plastic strain
observed in Figure 5.7(a).

The FE2 model predicts a very high failure value because the in-situ effects on

the failure properties of the constituents have not been accounted for and also

due to the lack of an appropriate progressive fibre failure model [147–149]. The

experimental observations reported in Chapter 4 suggest that failure under

torsional loading is fibre controlled. It is therefore necessary to incorporate

progressive failure of the fibres in the RVE of the FE2 model in order to

accurately predict failure. The next section will discuss about updating the

125
compressive failure strain of the fibre to account for fibre microbuckling in the

FE2 model.

Modelling torsion with microbuckling of fibres

The fibre micro-buckling is modelled by assigning a buckling strain which is

lower than the compressive failure strain of the fibres. The microbuckling strain

is determined from micromechanical models in which an initial fibre

misalignment is modelled with a sinusoidal curve (Figure 5.8).

Figure 5.8 Sinusoidal function used to model initial misalignment and the FE
model of the RVE used for the simulation to determine the microbuckling
strain.

The misalignment angle is not known with certainty, and is assumed to be

3°[39]. Periodic Boundary Conditions(PBCs) are applied and simulations are

carried out as per [150] and the fibre failure is determined by its bulk failure

properties (in Table 5.1). Fibre debonding is not considered in this simulation.

For a 3° initial misalignment, the buckling strain was found to be 1.5% which

did not vary even upon doubling the initial length (L in Figure 5.8) of the RVE

which was assumed as 1000μm. The fibre failure strain in compression is now

updated with this buckling strain, which is slightly less than the compressive

strain tabulated in Table 5.1.

126
200

150
Shear Stress (MPa)

100

50
FE2 coarse model with 1.5% microbuckling strain(+z torque)
Test 6(-z torque)
Test 7(+z torque)
FE2 coarse model with 0.4% microbuckling strain(+z torque)
FE2 coarse model with 0.4% microbuckling strain(+z torque)
0
0.00 0.01 0.02
Shear Strain

Figure 5.9 Torsion with fibres modelled to fail at a preassigned microbuckling


strain rather than the compressive strain.

The fibres in the outermost ply fail first when torqued in the positive z direction.

However, even with the microbuckling strain obtained from the microbuckling

micro RVE model of 1.5%, the simulations predict 1.5 times higher failure load

of the pipe compared to the experimental measurement. This may be due to the

fact that the initially assumed fibre misalignment of 3° is conservative and might

be greater in the experimental specimens. This microbuckling strain is predicted

with a RVE of 0.6 fibre volume fraction. When assigned a microbuckling strain

of 0.4%, the FE2 model underpredicts the point of damage or strain localization

as shown in Figure 5.9. This indicates the sensitivity of the buckling strain on

the model behaviour. There are other failure mechanisms, such as fibre

debonding, matrix failure and delamination, which will have to be accounted

for in the model in order to correctly predict the microbuckling strain. At

present, these have not been considered.

127
Bending

The boundary conditions used for simulating the bending experiments are

shown in Figure 5.10. A bending moment is applied on the positive z face and

the negative z face is completely constrained in x, y and z directions.

As mentioned while describing the experiments, a transverse shear stress is

present while testing the pipes in bending and the shear force is applied along

with bending moment on the loading face of the model as shown in Figure 5.10.

The model behaviour is compared with experiments in Figure 5.11. Since the

transverse shear stress is lower relative to the longitudinal stress, the models

Figure 5.10 Boundary Conditions for bending.

exhibit similar behaviour. Models with higher transverse shear stress exhibit

higher moments as expected (from equation (4.1)).

128
The stress contours of the RVEs from the plies are shown in Figure 5.12. The

RVEs hardly show any plastic strain even in the presence of shear force (Figure

5.12(b).

120
Bending longitudinal macroscale stress (zz)(MPa)

100

80

60

40

1m pipe(Test 5) at 0.25m from pipe bottom


1.5m pipe(Test 4) at 0.375m from pipe bottom
20 1.5m pipe(Test 4)- stress at 0.25m vs strain at 0.375m
Direct FE2 coarse mesh with zero y z:zz
Direct FE2 coarse mesh with 0.054 y z:zz(Test 4)

Direct FE2 coarse mesh with 0.1 y z:zz(Test 5)


0
0.0000 0.0075 0.0150
Strain

Figure 5.11 Bending longitudinal stress- strain behaviour: Comparing FE2


models with experiments.

The presence of the transverse shear stress affects the axial σzz stress (S33) and

τxy and τyz shear stresses (S12 and S23) as can be seen from Figure 5.12. The

matrix plastic strain are negligible as shown in Figure 5.12 (a) and (b). The τxy

(S12) stress increases (Figure 5.12(c) and (d)) and τyz (S23) stress decreases in

magnitude when acting positively ( when shear stress and face normal are of the

same sign) on a face and both S12 and S23 increases when acting negatively on

a face in response to transverse shear stresses. The axial stresses are mildly

affected as can be seen in Figure 5.12 (g) and (h).

129
Figure 5.12 Stress contours of the tensile side RVEs in bending at a bending
axial stress of 39.8MPa (point of divergence of FE2 model behaviour). The top
right RVE in each image is the outermost ply. The left column of contours-(a),
(c), (e) and (g)-has zero transverse shear stress and the right column-(b), (d), (f)
and (h)-has 0.1 τyz : σzz as reported in bending Test no. 5 in Chapter 4.Plastic
strain contours in (a) and (b);Shear stress in xy in (c) and (d); Shear stress in yz
in (e) and (f); normal(longitudinal) stress in axial direction in (g) and (h).

The Direct FE2 model is expected to capture bending since it was able to model

tensile tests satisfactorily. The failure stress during the bending tests were higher

than the tensile strength. However, in bending, apart from tensile loads,

compressive loads are also present on the other side which may not be captured

130
appropriately by the FE2 model. To investigate this, the Direct FE2 pipe model

being discussed here is simulated in compression and compared with

compression experiments performed on coupons cut out from the pipe [151].

The model used is shown in Figure 5.2 and instead of a tensile displacement on

the loading face, a compressive displacement is applied.

200
Compressive stress (MPa)

150

100

UC1
50 UC2
UC3
UC4
Fully intact Direct FE2(coarse mesh)
completely delaminated Direct FE2
0

0.00 0.02 0.04


Compressive strain

Figure 5.13 Compression: FE2 vs experiments: UC1,UC2,UC3 and UC4 refer


to compression coupon experiments [151].

Delamination was observed during the tests. The results are compared in Figure

5.13. In addition to the fully intact Direct FE2 model above, a completely

delaminated FE2 model is also simulated in which 2 sets of macroscale nodes

are present at the ply interfaces--one set for each ply which can deform and

displace independent of one another. It can be seen that the completely

delaminated model behaviour is very similar to experimental coupon response

to compression. This suggests that in order for the Direct FE2 model to correctly

capture failure in compression, some form of delamination criterion needs to be

131
incorporated. This explains why the current model is not able to replicate failure

due to bending as observed in experiments.

Using cohesive elements along with 3D Direct FE2 model

This section intends to highlight the ability to use cohesive zone model along

with the Direct FE2 model to capture delamination and possibly fibre debond as

demonstrated in 2D in section 3.4 The problem with using it here in the current

model is that the macroscale elements are too large to capture appropriate

delamination behaviour. Having only a single element along the axial direction

of the pipe is not encouraging to use cohesive elements to model delamination.

The pipe could not be modelled with convenient number of elements in the axial

direction due to computational constraints. Cohesive elements COH3D8 can be

added in between the macroscale elements to capture delamination as can be

seen in the Figure 5.14.

Figure 5.14 Cohesive elements (COH3D8) used with the Direct FE2 model:
Cohesive element Damage variable SDEG (Overall scalar stiffness
degradation) contours are shown.

For appropriately sized models like coupons, Direct FE2 with cohesive zone

models can be used for delamination modelling similar to the modelling of fibre

matrix debond in 2D to potentially capture all fibre reinforced polymer

132
composite failure modes- debond (CZM), matrix failure, fibre failure and

delamination (with CZM). However, the parameters and the models to be used

to capture these failure modes and the interactions between them accurately

needs more research effort.

Discussion

The Direct FE2 method has been used to model a macrostructure such as the

composite pipe and it captures the elastic behaviour of the structure

satisfactorily as can be seen in the model’s behaviour in the loading cases

discussed in the previous sections. A question as to why the Direct FE2 with

fine meshed RVEs is stiffer than coarse mesh RVEs remains, but it is to be noted

that as we refine the mesh, the model behaviour becomes closer to the

homogenized properties as shown in section 3.6.

Whether it reasonable to model just 10mm length of 1000 mm pipe to predict

its behaviour depends on the loading conditions and the objective of the model.

If it is intended to model post-strain localization, obviously a 10 mm model is

not sufficient. The aim of this chapter is to capture the pipe behaviour up to the

point of localization. Being constrained by computational resources and time, a

10 mm model was used to mixed degree of success. It is effective for

axisymmetric loading conditions, such as torsion and tension - for which

modelling even a section of the circumference would be sufficient as can be

seen from the Figure 5.15 (a) and (b).

133
70

60

Tensile stress (MPa) 50

40

30

20
FE2 -coarse mesh RVEs
10 FE2 -coarse mesh RVE(20mm-thick)
FE2 -coarse mesh RVE(5mm-thick)

0
0.000 0.005 0.010 0.015
(a) Tensile strain ()

200

180

160
Shear Stress (MPa)

140

120

100

80

60

40
FE2 coarse mesh-5mm thick
20 FE2 coarse mesh-20mm thick
FE2 coarse mesh-10mm thick
0
0.00 0.01 0.02

(b) Shear Strain

Figure 5.15 Effect of model length for (a) pipe tension; (b) pipe torsion.

But for bending, having just a 10 mm model is debatable and there is no real

justification to be made. It is assumed that a section with the bending moment

along with the transverse shear forces measured in the pipe being modelled is

representative of the pipe. For bending without shear force, the model length

134
does not seem to affect the longitudinal stress as can be seen in Figure 5.16 (a).

Figure 5.16 (b) shows the effect of model length on bending with shear force.

The difference between the Direct FE2 model and the experiments can be

attributed to a combination of the following possibilities: (a) lack of information

about the structural or manufacturing defects in the FE2 model, (b) an

approximate constitutive relation of the matrix, (c) not modelling progressive

fibre failure due to lack of a good model [147–149], (d) not accounting for

debonding and delamination as their contributions to tensile failure are not

understood well and (e) not modelling and accounting for the exact structure of

the pipe due to filament winding. A good model of fibre failure needs to be able

to capture a complicated process, which may involve individual fibre breakages

at random locations which eventually join up through extensive debonding or

matrix cracking, taking the statistical size effect (due to randomness of strength)

into account. The problem is further complicated by the fact of size effects in

composites [152,153] and in-situ effect on matrix failure strength [154]. Using

bulk strengths is debatable since it is known that with increase in size, strength

decreases due to the presence of larger number of defects as compared to small

sized specimens or structures as observed in experiments with composite

materials [152,153].

135
Bending longitudinal macroscale stress (zz)(MPa)
100

50

Direct FE2 coarse mesh with zero y z:zz


Direct FE2 coarse mesh with zero y z:zz-20mm long
Direct FE2 coarse mesh with zero y z:zz(Test 5)-5mm long
0
0.000 0.005 0.010
(a) Strain
Bending longitudinal macroscale stress (zz)(MPa)

100

50

Direct FE2 coarse mesh with 0.1 y z:zz(Test 5)


Direct FE2 coarse mesh with 0.1 y z:zz(Test 5)-20mm long

Direct FE2 coarse mesh with 0.1 y z:zz(Test 5)-5mm long

0
0.000 0.005 0.010
(b) Strain

Figure 5.16 Effect of model length for (a) pipe bending without shear force;
(b) pipe bending with shear force.

136
Efforts have been directed towards modelling of composite behaviour and

prediction of its failure as evident from the world-wide failure exercises [155–

157] and numerous other researches. The predictions deviate from the

experimental results. In spite of extensive research, understanding and

predicting failure of the composite pipes remains a challenge [158]. Theoretical

prediction of failure for composite pipes and composites in general is still an

ongoing research area, which is evident based on outcomes of the world-wide

failure exercises conducted for composites [155,156]. Failure in composite

laminates is an accumulation of micro-level failure events. The difficulty in

predicting their failure arises in relating these micro-level failure events to

consequent changes in material response.

Summary

The Direct FE2 laminate modelling methodology presented in section 3.6 is

applied for prediction of the structural response of glass fibre reinforced epoxy

composite pipes loaded on Stewart platform-based test ring described in

Chapter 4. The inputs for the model include the constituent properties. Even

though not considered here, debonding and delamination can be modelled with

cohesive elements and the method is shown to be capable of being used along

with the cohesive zone model, but not applied here due to computational

constraints. The developed 3D composite model is employed for the modelling

of the composite pipe experiments. The predictions are contrasted with

experimental measurements.

The Direct FE2 performance in tension is easily captured because of uniform

non-linearity due to the matrix undergoing plastic deformation throughout the

137
pipe length as can be seen in the form of matrix crazing along the pipe length

up to the point of ultimate failure and the tensile behaviour of the pipe is matrix

controlled.

The non-linear behaviour of pipe in torsion is well captured. Being fibre

dominated and with only bulk failure properties, the model predicts a very high

failure load. However, if microbuckling can be modelled and also controlled by

buckling strain, then we can tune the model to predict reasonably the onset of

damage localization. However, this method needs to be verified and validated

further experimentally before being adopted.

Bending experiments have also been satisfactorily captured in the regime upto

the axial tensile failure stresses. Beyond that, the model overpredicts pipe

strength because other potential failure modes such as delamination are not

considered and fibre failure have not been modelled effectively due to lack of

good fibre failure models [147–149].

The model presented here might enable to design composite structures from the

basic material constituents, after a complete validation with experiments. The

model predicts that by using an epoxy with a higher tensile strength will give a

pipe with higher axial tensile strength. This is useful, but this needs to be

validated experimentally. This insight from the model is supported with some

experimental observations of matrix crazing observed in the pipes during the

tensile experiments as shown in Figure 4.8. However as mentioned previously,

these model insights need to be further validated experimentally.

138
Conclusions and Future Work

The contributions of this thesis is summarized in the following points below.

The conclusions and future work are discussed in sections 6.1 and 6.2

respectively.

 Implemented and benchmarked a robust concurrent multi-scale model,

called Direct FE2 in 2D (Chapter 2) and 3D.

 A direct constituent to laminate model has been built and demonstrated

to be useful for homogenization at the laminate level, bypassing the

intermediate ply level. Plies emerge directly as result of the relative

orientation of the microstructures and the scale transition relationships

between the RVEs and the macroscale elements.(Section 3.6)

 Demonstrated the potential to predict heterogeneous structural behavior,

effectively eliminating an intermediate homogenization required in

hierarchical models and scaling up single unit cell homogenization to

the structural level.(Chapter 5)

Conclusions

A concurrent multi-level Finite Element Method called Direct FE2 has been

described and implemented for both 2D and 3D and has some notable features

that are recalled here. The Direct FE2 method implemented, unlike traditional

FE2 implementations, does not require programming for information/data

interchange between the micro and macro scales because finite element

calculations for both the scales are carried out in the same finite element

139
simulation as the macro scale nodes are coupled to the microscale nodes through

multi-point constraints. The method proposed here does not require a separate

definition of the microscale boundary conditions as the microscale RVE

boundary nodes are linked to the macroscale element nodes through multi-point

constraints and the microscale boundary conditions arise naturally because of

this strong coupling between the macroscale and the microscale.

The Chapters 2 and 3 establish that Direct FE2 is a viable method to model

heterogeneous materials like composites by comparing Direct FE2 models with

brute force full FE models. The Direct FE2 models are shown to have

considerable computational cost savings with respect to brute force full FE

model. The experimental study of the failure of Glass Reinforced Epoxy

composites pipes under tension, torsion and bending with a Stewart platform

based experimental setup have been described in chapter 4. In the conducted

tests the failure mechanisms for different loading have been identified and

described.

The Direct FE2 method is used to model composite structures from their

constituents as has been presented in section 3.6 and has been used to model

pipe experiments as discussed in Chapter 5. Plies emerge directly as result of

the relative orientation of the microstructures and the scale transition

relationships between the RVEs and the macro model. The method offers the

potential of modelling failure mechanisms like matrix failure, fibre failure and

interphase failure if the appropriate models and material properties are

determined. The model is able to predict the linear structural response observed

in all the loading cases in Chapter 5 as well as uniform non-linear behaviour as

140
observed in modelling tensile experiments. The Direct FE2 models of the pipe

cannot capture microbuckling of the fibres in the RVE as we cannot model the

minimum length required to observe microbuckling in the Direct FE2 models.

The model can be tuned to capture failure modes such as microbuckling with a

user subroutine and delamination with cohesive models as discussed in sections

5.3 and 5.5 respectively.

For modelling complete engineering size structures from their microscale

constituents (in this work we modelled a 10mm tall annular ring to represent a

1000mm tall pipe), the immense computational costs comes along with the

difficulties of experimental characterization at smaller scales [18]. Still the

method provides a lens to zoom into microscale events occurring in response to

macroscale inputs (loads) which is necessary if we want to completely

understand and characterize heterogeneous materials and their use in critical

applications such as high pressurized fuel containers, etc. The method can

contribute to a virtual composite testing paradigm in which structural

components are analysed directly from their constituents once the following

hurdles are overcome - constituent properties and model determination,

interface characterization and computational constraints for modelling large

structures.

Future directions

The future possibilities are categorised as possibilities for the Direct FE2

method, which will be outlined first and secondly future work for the application

of the method for modelling composite structures.

141
Future possibilities for Direct FE2

Comparing the Direct FE2 with a traditional FE2 simulation for studying the

computational efficiency and time savings would be of interest. The first step

towards realizing this would be to implement a traditional FE2 implementation,

which could be a potentially challenging future endeavour.

The Direct FE2 method in this thesis has been used and implemented only on

continuum 2D plane stress elements and 3D C3D8 elements. Extending this to

shell elements might be very useful in modelling and designing thin composite

plates and layered structures with heterogeneous microstructures. If Direct FE2

can be implemented in an explicit framework, it would be useful to study

multiscale wave propagation in heterogeneous structures used in critical

application such as aircraft and spacecraft.

Implementing the Direct FE2 method to model crystalline materials with their

granular structure might be of interest for the crystal plasticity community.

Future work for modelling composite structures with Direct FE2

The method can be used to investigate whether in–situ behavior of composite

plies can be captured in the Direct FE2 laminate model with a refined macroscale

mesh. Adaptive remeshing of macroscale elements to capture post localization

behavior will be another interesting area to explore to capture damage and

failure behavior of heterogeneous materials more comprehensively.

For realistic modelling of structures, developing Direct FE2 models with various

realizations of random unit cells in a single model might be of interest to the

composite community. The only hurdle in such an effort would be constraint of

142
computational resources since each realization’s data needs to be given as input

to the Direct FE2 model.

Apart from multiscale modelling, it has been shown that the Direct FE2 method

can be used for scaling up single unit cell homogenization to the structural level

and potentially can be used for macrostructural design from the microstructures

eliminating an intermediate homogenization needed to model large structures.

143
APPENDIX A: APPLYING MICROSCALE BOUNDARY
CONDITIONS IN DIRECT FE2

Displacement (u,v) at any point (X,Y) inside the macroscale element as shown
in Figure A.1 is given by:

u= 𝑁1 ∗ 𝒖𝟏 + 𝑁2 ∗ 𝒖𝟐 + 𝑁3 ∗ 𝒖𝟑 + 𝑁4 ∗ 𝒖𝟒

v= 𝑁1 ∗ 𝒗𝟏 + 𝑁2 ∗ 𝒗𝟐 + 𝑁3 ∗ 𝒗𝟑 + 𝑁4 ∗ 𝒗𝟒

(A.1)

The interpolation functions should be able to capture the appropriate


deformation of the material being modelled. It should also be noted that the
interpolation function that link the micro-structure (Gauss point RVEs) to the
macro-structure (macroscale element nodes) play the role of scale transition.

Figure A.1: A FE2 macro scale-element: the dotted lines defines the
macroscale: 4 nodes of the macro-element defined as Reference point); The
macroscale nodal displacements are ui, vi; i=1,2,3,4; and the micro structure
RVEs at the gauss-points; MPCs at the edge nodes of the RVEs.

The inverse iso-parametric mapping (converting the global co-ordinates into


natural local coordinates), X,Y->𝜉, 𝜂 is carried out. Assuming a bilinear
interpolation function to for scale transition, equation (A.1) becomes:

144
1 1
u= 4 ∗ (1-𝜉) ∗ (1-𝜂) ∗ 𝒖𝟏 + 4 ∗ (1 + 𝜉) ∗ (1-𝜂) ∗ 𝒖𝟐 +

1 1
∗ (1 + 𝜉) ∗ (1+𝜂) ∗ 𝒖𝟑 + ∗ (1 − 𝜉) ∗ (1 + 𝜂) ∗ 𝒖𝟒
4 4

1 1
v= 4 ∗ (1-𝜉) ∗ (1-𝜂) ∗ 𝒗𝟏 + 4 ∗ (1 + 𝜉) ∗ (1-𝜂) ∗ 𝒗𝟐 +

1 1
∗ (1 + 𝜉) ∗ (1+𝜂) ∗ 𝒗𝟑 + ∗ (1 − 𝜉) ∗ (1 + 𝜂) ∗ 𝒗𝟒
4 4

(A.2)

Equation A.1 is used to apply linear displacement boundary conditions


expressed in equation (2.20)

Figure A.2 RVE at Integration (Gauss) Point α.

The Periodic boundary conditions are applied as stated in equations ((2.23) and
(2.24)) as described below:
𝟒

𝒖𝑅𝑉𝐸𝛼
𝑟 − 𝒖𝑅𝑉𝐸𝛼
𝑙 = ∑(𝑵𝑰 (𝒙𝑹 , 𝒚𝑹 ) − 𝑵𝑰 (𝒙𝑳 , 𝒚𝑳 )) 𝑢𝐼 𝒆
𝑰=𝟏

𝒗𝑅𝑉𝐸𝛼
𝑟 − 𝒗𝑅𝑉𝐸𝛼
𝑙 = ∑(𝑵𝑰 (𝒙𝑹 , 𝒚𝑹 ) − 𝑵𝑰 (𝒙𝑳 , 𝒚𝑳 )) 𝑣𝐼 𝒆
𝑰=𝟏

(A.3)

where, R and L denote corresponding nodes on the right and left RVE edges in
Figure A.2 and 𝑵𝑰 is the shape functions for macroscale element node I as

145
shown in Figure A.1. These MPCs are needed to be applied for all the Gauss
point RVEs (for RVE𝛼, where 𝛼=1,4 in Figure A.1).

Similarly, MPCs are applied for the top and bottom edges of each Gauss point
RVE as shown in the equations:
𝟒

𝒖𝑅𝑉𝐸𝛼
𝑡 − 𝒖𝑅𝑉𝐸𝛼
𝑏 = ∑(𝑵𝑰 (𝒙𝑻 , 𝒚𝑻 ) − 𝑵𝑰 (𝒙𝑩 , 𝒚𝑩 )) 𝑢𝐼 𝒆
𝑰=𝟏

𝒗𝑅𝑉𝐸𝛼
𝑡 − 𝒗𝑅𝑉𝐸𝛼
𝑏 = ∑ 𝑵𝑰 (𝒙𝑻 , 𝒚𝑻 ) − 𝑵𝑰 (𝒙𝑩 , 𝒚𝑩 ) 𝑣𝐼 𝒆
𝑰=𝟏

(A.4)

146
APPENDIX B: STRAIGHT EDGE BOUNDARY CONSTRAINTS

Figure B.1 Straight Edge Boundary Constraint illustration for FE model.

Face2 horizontal edge between V1-V2:


𝑥𝐴 −𝑥𝑣2
𝑢𝐴 = 𝑢𝑣2 + (𝑢 − 𝑢𝑣2 )
𝑥𝑣1 − 𝑥𝑣2 𝑣1

𝑥𝐴 −𝑥𝑣2
𝑣𝐴 = 𝑣𝑣2 + (𝑣 − 𝑣𝑣2 )
𝑥𝑣1 − 𝑥𝑣2 𝑣1

Face1 vertical edge between V1-V4:


𝑦𝐵 − 𝑦𝑣4
𝑢𝐵 = 𝑢𝑣4 + (𝑢 − 𝑢𝑣4 )
𝑦𝑣1 − 𝑦𝑣4 𝑣1

𝑦𝐵 − 𝑦𝑣4
𝑣𝐵 = 𝑣𝑣4 + (𝑣 − 𝑣𝑣4 )
𝑦𝑣1 − 𝑦𝑣4 𝑣1

Constraints are similarly applied for the other two faces to ensure that the edges
remain straight.

147
APPENDIX C: PERIODIC BOUNDARY CONDITIONS (PBC)

Figure C.1 Periodic Boundary Conditions illustration for FE model.

In figure C.1; 1,2,3,4 denotes the vertex nodes of the RVE. RP-1 and RP-2-
reference points, denoted as 𝑅𝑃𝐻 , and 𝑅𝑃𝑉 respectively.

RP-1 is 𝑅𝑃𝐻 for constraining the horizontal sides- top edge between vertices 2-
1(denoted by T) and bottom edge between vertices 3-4 (denoted by B) of length
𝐿𝑥 .

RP-2 is 𝑅𝑃𝑉 for constraining the vertical sides- right edge between vertices 1-4
(denoted by R) and left edge between vertices 2-3 (denoted by L) of length 𝐿𝑦 .

The displacements of the reference points in x,y direction is denoted by:


𝑅𝑃𝑉 𝑅𝑃𝑉 𝑅𝑃𝐻 𝑅𝑃𝐻
𝛿𝑥 , 𝛿𝑦 , 𝛿𝑥 , 𝛿𝑦

𝑅𝑃𝑉 𝑅𝑃𝑉 𝑅𝑃𝐻 𝑅𝑃𝐻


The forces on the RPs are denoted by 𝐹𝑥 , 𝐹𝑦 , 𝐹𝑥 , 𝐹𝑦 and given as:
𝑅𝑃𝑉 𝑅𝑃𝑉
𝐹𝑥 = 𝜎𝑥 𝐿𝑦 , 𝐹𝑦 = 𝜏𝑥𝑦 𝐿𝑦

𝑅𝑃𝐻 𝑅𝑃𝐻
𝐹𝑥 = 𝜏𝑥𝑦 𝐿𝑥 , 𝐹𝑦 = 𝜎𝑦 𝐿𝑥

where, the global stress states are given as: [ 𝜎𝑥 𝜎𝑦 𝜏𝑥𝑦 ]

148
Constraints on the displacement of edges nodes:
𝑅𝑃𝑉
𝑢𝑖𝑅 − 𝑢𝑖𝐿 = 𝛿𝑥 , 𝑖 = 1,2,3, … , 𝑛𝑉

𝑅𝑃𝑉
𝑣𝑖𝑅 − 𝑣𝑖𝐿 = 𝛿𝑦 , 𝑖 = 1,2,3, … , 𝑛𝑉

𝑅𝑃𝐻
𝑢𝑖𝑇 − 𝑢𝑖𝐵 = 𝛿𝑥 , 𝑖 = 1,2,3, … , 𝑛𝐻

𝑅𝑃𝐻
𝑣𝑖𝑇 − 𝑣𝑖𝐵 = 𝛿𝑦 , 𝑖 = 1,2,3, … , 𝑛𝐻

Constraints on the displacement of vertex nodes:


𝑅𝑃𝑉 𝑅𝑃𝐻
𝑣𝑣1 –𝑣𝑣3 – 𝛿𝑦 −𝛿𝑦 =0

𝑅𝑃𝑉 𝑅𝑃𝐻
𝑢𝑣1 – 𝑢𝑣3 – 𝛿𝑥 − 𝛿𝑥 =0

𝑅𝑃𝐻
𝑣𝑣2 – 𝑣𝑣3 − 𝛿𝑦 =0

𝑅𝑃𝐻
𝑢𝑣2 – 𝑢𝑣3 − 𝛿𝑥 =0

𝑅𝑃𝑉
𝑣𝑣4 – 𝑣𝑣3 - 𝛿𝑦 =0

𝑅𝑃𝑉
𝑢𝑣4 – 𝑢𝑣3 – 𝛿𝑥 =0

149
APPENDIX D: SETTING UP A DIRECT FE2 MODEL IN ABAQUS

Direct FE2 models can be setup in Abaqus by following the procedure detailed
in the flowchart shown in Figure D.1

Figure D.1 Implementation of Direct FE2 in Abaqus.

A python script for placing the microscale RVEs at the macroscale element
integration points is given below titled ‘GP_RVE_place.py’ (Step 2 in Figure
D1).

A python script to define the scale transition relationship between the


microscale and the macroscale Degrees of freedom (dofs) for linear
displacement BCs and PBCs for linear macroscale elements (Step 4 in Figure
D.1) is given in python scripts titled- ‘FE2-linear_displacement-MPCs.py’ and
‘FE2-PBCs-for-linear_macroscale_elements.py’ respectively.

150
GP_RVE_place.py
#user input
ele_connect=raw_input("name of element_ connectivity .dat file?")
FE_2_model_gen=raw_input("name of Fe_square abaqus model generation .py file?")
integration=input("input 0 for full integration or 1 for reduced integration FE_2 element:")
order=input("Does the macroscale model have linear or quadratic elements? Enter 1 for linear
macroscale elements or Enter 2 for quadratic macroscale elements:")
model=raw_input("Which model?")
Partname=raw_input("Which part?")

#empty sets to collect macroscale nodal corordinates and macroscale nodal connectivity
Nodes=[];
nodal_connectvity=[];

#coordinates of the Gauss integration points in natural coordinate system(-1,+1) for macroscale element
GPs=([[-3**-0.5,-3**-0.5,0],[3**-0.5,-3**-0.5,0],[3**-0.5,3**-0.5,0],[-3**-0.5,3**-0.5,0]],
[[0,0,0]])# Reduced

#directory in which .inp file is present


os.chdir('C:\\Temp\\Karthic_temp\\FE_2_FYP\\DirectoryName\\')
#keywords to identify the macroscale nodal coordinate data and macroscale nodal connectivity data
keyword=['*Node','*Element, type=S4R'];
f1=open('20x2000-quad.txt','r+')#input file name
inp=[]

#//////////////////////// read in all the lines in the input file


while 1:
line=f1.readline()
if not line:
break
line=line.strip()
inp.append(line)
#/////////////////////// search the keyword in inp, and assign the variables
def Search(inp,line,array):
array[:]=[]
a=inp.index(line);

if (inp[a].count("generate")==0):
for line_temp in inp[(a+1):]:
if (line_temp=='') or (line_temp.count("*")!=0):
break;
line_temp=(line_temp.replace(',','')).split()
array.append(line_temp)
else:
for line_temp in inp[(a+1):]:
if (line_temp=='') or (line_temp.count("*")!=0):
break
line_temp=(line_temp.replace(',','')).split()
for i in range(int(line_temp[0]),(int(line_temp[1])+1),int(line_temp[2])):
array.append(i)

Search(inp,keyword[0],Nodes);#node no, x,y,x


Search(inp,keyword[1],nodal_connectvity);#element no,N1,N2,N3,N4

#function to calculate shape functions for linear interpolation


def tsi_eta_shape_fn(tsi,eta):
N1=float(0.25*(1-tsi)*(1-eta))
N2=float(0.25*(1+tsi)*(1-eta))
N3=float(0.25*(1+tsi)*(1+eta))
N4=float(0.25*(1-tsi)*(1+eta))
return N1,N2,N3,N4

def tsi_eta_quad_shape_fn(tsi,eta):
N1=float(-0.25*(1-tsi)*(1-eta)*(1+tsi+eta))
N2=float(-0.25*(1+tsi)*(1-eta)*(1-tsi+eta))

151
N3=float(-0.25*(1+tsi)*(1+eta)*(1-tsi-eta))
N4=float(-0.25*(1-tsi)*(1+eta)*(1+tsi-eta))
N5=float(0.25*2*(1-tsi**2)*(1-eta))
N6=float(0.25*2*(1+tsi)*(1-eta**2))
N7=float(0.25*2*(1-tsi**2)*(1+eta))
N8=float(0.25*2*(1-tsi)*(1-eta**2))
return N1,N2,N3,N4,N5,N6,N7,N8

# opening the .py file with file name given as user_input stored in variable 'FE_2_model_gen'
RVE = open('%s.py'%(FE_2_model_gen),'w')

# print statements write into the file opened above


#writing all the preliminary import statements to write a python abaqus script
print>>RVE, "from abaqus import *"
print>>RVE, "from abaqusConstants import *"
a='''
#==========import modulus of abaqus
import part
import regionToolset
import displayGroupMdbToolset as dgm
import material
import section
import assembly
import step
import interaction
import load
import mesh
import job
import sketch
import visualization
import xyPlot
import displayGroupMdbToolset as dgm
'''
RVE.write(a)

#importing math modules to make available math functions like square root which is not used in this, but
to be included here in case needed for future modifications
a='''
#import math
#import numpy as np

#===================== construction of the FE2 model


a1 = mdb.models['%s'].rootAssembly
'''%(model)
RVE.write(a)
for i in range(0,len(Nodes)):
print>>RVE,'a1.ReferencePoint(point=(%E, %E,
%E))'%(float(Nodes[i][1]),float(Nodes[i][2]),float(Nodes[i][3]))
a='''a1.Set(referencePoints=(a1.referencePoints[mdb.models['%s'].rootAssembly.features['RP-
%s'].id], ), name='N%s-RP')\n'''%(model,str(i+1),str(i+1));
#line below to create sets of nodes if you have a macroscale mesh with Part-instance name Part-3-1
#print>>RVE,"a1.Set(name='N%s-RP', nodes=(a1.instances['Part-3-
1'].nodes[%s:%s],))"%(str(i+1),str(i),str(i+1))
RVE.write(a)

# loop around list of macroscale elements as per nodal connectivity data from macroscale element
number 1 to len(nodal_connectvity)
for i in range(0,len(nodal_connectvity)):
#macroscale element number
ele_no=int(nodal_connectvity[i][0]);

152
#macroscale nodes (node number in ni and coordinates Xi,Yi,Zi) of macroscale element number : ele_no
n1=int(nodal_connectvity[i][3])
n2=int(nodal_connectvity[i][4])
n3=int(nodal_connectvity[i][1])
n4=int(nodal_connectvity[i][2])

X1=float(Nodes[n1-1][1])
Y1=float(Nodes[n1-1][2])
Z1=float(Nodes[n1-1][3])

X2=float(Nodes[n2-1][1])
Y2=float(Nodes[n2-1][2])
Z2=float(Nodes[n2-1][3])

X3=float(Nodes[n3-1][1])
Y3=float(Nodes[n3-1][2])
Z3=float(Nodes[n3-1][3])

X4=float(Nodes[n4-1][1])
Y4=float(Nodes[n4-1][2])
Z4=float(Nodes[n4-1][3])

#need four more nodes for each macroscale element in case of quadratic macroscale elements as will be
indicated by user input stored in variable 'order'
if order==2:
n5=int(nodal_connectvity[i][7])
n6=int(nodal_connectvity[i][8])
n7=int(nodal_connectvity[i][5])
n8=int(nodal_connectvity[i][6])

X5=float(Nodes[n5-1][1])
Y5=float(Nodes[n5-1][2])
Z5=float(Nodes[n5-1][3])

X6=float(Nodes[n6-1][1])
Y6=float(Nodes[n6-1][2])
Z6=float(Nodes[n6-1][3])

X7=float(Nodes[n7-1][1])
Y7=float(Nodes[n7-1][2])
Z7=float(Nodes[n7-1][3])

X8=float(Nodes[n8-1][1])
Y8=float(Nodes[n8-1][2])
Z8=float(Nodes[n8-1][3])

#assigning the coordinates of the points where the RVEs are going to be placed as per full or reduced
order of integration for macroscale elements
GP_RVEs=GPs[int(integration)]

#loop from 1 to no. of Gauss point RVEs

#for placing identical RVEs(Partname) by creating various assembly instances and placing the RVEs at
macroscale gauss integration points
for ii in range(1,(len(GP_RVEs)+1)):

#next conditional statements calculate the x,y,z of the (0,0) of RVE to be placed in the assembly
based on whether order=1 or order=2
#if RVE center not at (0,0), we need to shift all RVE instances uniformly
if order==1:
N1,N2,N3,N4=tsi_eta_shape_fn(GP_RVEs[ii-1][0],GP_RVEs[ii-1][1]);
position_x=N1*X1+N2*X2+N3*X3+N4*X4
position_y=N1*Y1+N2*Y2+N3*Y3+N4*Y4
position_z=N1*Z1+N2*Z2+N3*Z3+N4*Z4
if order==2:
N1,N2,N3,N4,N5,N6,N7,N8=tsi_eta_quad_shape_fn(GP_RVEs[ii-1][0],GP_RVEs[ii-1][1]);

153
position_x=N1*X1+N2*X2+N3*X3+N4*X4+N5*X5+N6*X6+N7*X7+N8*X8
position_y=N1*Y1+N2*Y2+N3*Y3+N4*Y4+N5*Y5+N6*Y6+N7*Y7+N8*Y8
position_z=N1*Z1+N2*Z2+N3*Z3+N4*Z4+N5*Z5+N6*Z6+N7*Z7+N8*Z8

print>>RVE, '#=========When assembling '+'Macroscale ele-'+str(ele_no)+'-RVE @ GP-


'+str(ii)
a='''a = mdb.models['%s'].rootAssembly
'''%(model)
RVE.write(a)
#Partname='';Partname=str(Part-1-ele-%d-GP-%d
print>>RVE, "p = mdb.models['%s'].parts['"%(model)+str(Partname)+"']"

print>>RVE, "a.Instance(name='Part-1-ele-%s-GP-%s', part=p, dependent=ON)"


%(str(ele_no),str(ii))
print>>RVE, "p2 = a.instances['Part-1-ele-%s-GP-%s']" %(str(ele_no),str(ii))
a='''a = mdb.models['%s'].rootAssembly
a = mdb.models['%s'].rootAssembly
'''%(model,model)
RVE.write(a)
# orientation of RVEs in 2D
angle_x=0.0
angle_y=0.0
angle_z=0.0

#statements to place the RVE instance at GP[ii]


print>>RVE,'p2.rotateAboutAxis(axisPoint=(0.0,0.0,0.0),axisDirection=(1.0,0.0,0.0),angle=%s)'
%(str(angle_x))
print>>RVE,'p2.rotateAboutAxis(axisPoint=(0.0,0.0,0.0),axisDirection=(0.0,1.0,0.0),angle=%s)'
%(str(angle_y))
print>>RVE,'p2.rotateAboutAxis(axisPoint=(0.0,0.0,0.0),axisDirection=(0.0,0.0,1.0),angle=%s)'
%(str(angle_z))
print>>RVE, "p2.translate(vector=(%s,%s,%s))" %(str(position_x),str(position_y),str(position_z))
print>>RVE, "a = mdb.models['%s'].rootAssembly"%(model)
print>>RVE, "p2 = a.instances['Part-1-ele-%s-GP-%s']" %(str(ele_no),str(ii))
print>>RVE,''

#close the python script


RVE.close()

#opening a .dat file needed for nodal connectivity data for writing the MPCs to link the macroscale to
microscale
NC=open('%s.dat'%(str(ele_connect)),'w')
for i in range(0,len(nodal_connectvity)):
if order==1:#linear
print>>NC,'[%d, %d, %d, %d],'%(int(nodal_connectvity[i][3]),int(nodal_connectvity[i][4]),
int(nodal_connectvity[i][1]),int(nodal_connectvity[i][2]))
if order==2:#quadratic
print>>NC,'[%d, %d, %d, %d, %d, %d, %d,
%d],'%(int(nodal_connectvity[i][3]),int(nodal_connectvity[i][4]),
int(nodal_connectvity[i][1]),int(nodal_connectvity[i][2]),int(nodal_connectvity[i][7]),
int(nodal_connectvity[i][8]),int(nodal_connectvity[i][5]),int(nodal_connectvity[i][6]))

NC.close()

---end of code---

154
FE2-linear_displacement-MPCs.py
# script to apply MPCs for FE2 method by using the isoparametric formulation with
#linear displacement boundary condition at the microscale RVE boundaries
# Part and Model need to be named 'Part-1' and 'Model-1' respectively
# All macroscale nodes need to be named like N1-RP,N200-RP(where 1 or 200 is the macroscale node
number)
#importing abaqus modules for accessing through python script
from part import *
from material import *
from section import *
from assembly import *
from step import *
from interaction import *
from load import *
from mesh import *
from job import *
from sketch import *
from visualization import *
from connectorBehavior import *

import numpy as np
import math

session.journalOptions.setValues(recoverGeometry = COORDINATE)

Part = mdb.models['Model-1'].parts['Part-1']

a=mdb.models['Model-1'].rootAssembly

#macroscale nodal connectivity- need to give this as input- get from abaqus input(.inp) file of
#macroscale mesh

#use data from .dat file named NC generated in the GP_RVE_place.py

nodal_connectvity=np.array([[1,2,3,4]])
# to loop around macroscale nodal connectivity(all macroscale elements)
for i in range(0,int(np.shape(nodal_connectvity)[0])):
nodes=nodal_connectvity[i]
ele=i+1;#element number
X1=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[0])].xValue
X2=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[1])].xValue
X3=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[2])].xValue
X4=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[3])].xValue
Y1=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[0])].yValue
Y2=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[1])].yValue
Y3=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[2])].yValue
Y4=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[3])].yValue

for j in range(1,5):# loop to go around the gauss point RVEs in macroscale element no 'ele'
Faces = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Faces']
for ii in range(len(Faces.nodes)):#loop to go around the boundary nodes defined in the set Faces in
#the RVE GP-j
X=Faces.nodes[ii].coordinates[0]
Y=Faces.nodes[ii].coordinates[1]
# co-efficients of interpolation equations only for quadrilateral elements
#Dx=Ax*e+Bx*n+Cx*e*n ; Dy=Ay*e+By*n+Cy*e*n
#constant terms
Dx=float(4*X-(X1+X2+X3+X4))
Dy=float(4*Y-(Y1+Y2+Y3+Y4))

Ax=float(-X1+X2+X3-X4)
Ay=float(-Y1+Y2+Y3-Y4)

155
Bx=float(-X1-X2+X3+X4)
By=float(-Y1-Y2+Y3+Y4)

Cx=float(X1-X2+X3-X4)
Cy=float(Y1-Y2+Y3-Y4)

tsi=float(Dx/Ax);
eta=float(Dy/By);
assert abs(tsi)<=1 and abs(eta)<=1

#shape functions calculated at the boundary node ii in GP-j RVE.Faces


N1=float(0.25*(1-tsi)*(1-eta))
N2=float(0.25*(1+tsi)*(1-eta))
N3=float(0.25*(1+tsi)*(1+eta))
N4=float(0.25*(1-tsi)*(1+eta))

#defining set of the boundary node ii


a.Set(name='Ele'+str(ele)+'-RVE'+str(j)+'-Face_node-'+str(Faces.nodes[ii].label)+'-'+str(ii),
nodes=(
a.instances['Part-1-ele-'+str(ele)+'-GP-'+str(j)].nodes[Faces.nodes[ii].label-
1:Faces.nodes[ii].label],))

# MPC equation constraining x diaplacement of boundary node ii


mdb.models['Model-1'].Equation(name='Ele-'+str(ele)+'-GP-'+str(j)+'-Constraint_Face_nodes-1-
'+str(ii), terms=((1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face_node-'+str(Faces.nodes[ii].label)+'-'+str(ii), 1), (-N1,
'N'+str(nodes[0])+'-RP', 1), (-N2, 'N'+str(nodes[1])+'-RP', 1), (-N3,'N'+str(nodes[2])+'-RP', 1),
(-N4,'N'+str(nodes[3])+'-RP', 1)))
# MPC equation constraining y diaplacement of boundary node ii
mdb.models['Model-1'].Equation(name='Ele-'+str(ele)+'-GP-'+str(j)+'-Constraint_Face_nodes-2-
'+str(ii), terms=((1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face_node-'+str(Faces.nodes[ii].label)+'-'+str(ii), 2), (-N1,
'N'+str(nodes[0])+'-RP', 2), (-N2, 'N'+str(nodes[1])+'-RP', 2), (-N3,'N'+str(nodes[2])+'-RP', 2),
(-N4,'N'+str(nodes[3])+'-RP', 2)))
---end of code---

156
FE2-PBCs-for-linear_macroscale_elements.py
# script to apply MPCs for FE2 method by using the
# isoparametric formulation for linear macroscale elements with PBCs on the microscale RVEs

from part import *


from material import *
from section import *
from assembly import *
from step import *
from interaction import *
from load import *
from mesh import *
from job import *
from sketch import *
from visualization import *
from connectorBehavior import *

import numpy as np
import math

session.journalOptions.setValues(recoverGeometry = COORDINATE)
a=mdb.models['Model-1'].rootAssembly

#following can be uncommented if macroscale nodes sets N1-RP,N200-RP,etc are not defined
#where 1,200 are macroscale node numbers
#for i in range(1,730):
# a.Set(referencePoints=(a.referencePoints[mdb.models['Model-1'].rootAssembly.features
# ['RP-'+str(i)].id], ), name='N'+str(i)+'-RP')

#function to remove vertex node out of edge nodes set


def TakeVertexOut(face):
face.pop(0)
face.pop(-1)
return face

#arranges nodes in face as per coordinate- coordinate=0,1 is x,y respectively


def SortListOfNodes(face,coordinate):
newlist = []
oldlist = []
for ii in range(len(face.nodes)):
oldlist.append( face.nodes[ii].coordinates[coordinate])

orderedlist = sorted(oldlist)
for ii in range(len(oldlist)):
vecindex = oldlist.index(orderedlist[ii])
#newlist.append(oldlist[vecindex])
newlist.append(face.nodes[vecindex].label-1)

return newlist

## spits out shape functions given point coordinates X,Y


def vertex_shape_fn(P,X1,X2,X3,X4,Y1,Y2,Y3,Y4):
# co-efficients of interpolation equations only for quadrilateral elements
#Dx=Ax*e+Bx*n+Cx*e*n ; Dy=Ay*e+By*n+Cy*e*n
#constant terms
X=P.nodes[0].coordinates[0]
Y=P.nodes[0].coordinates[1]

Dx=float(4*X-(X1+X2+X3+X4))
Dy=float(4*Y-(Y1+Y2+Y3+Y4))

Ax=float(-X1+X2+X3-X4)
Ay=float(-Y1+Y2+Y3-Y4)

157
Bx=float(-X1-X2+X3+X4)
By=float(-Y1-Y2+Y3+Y4)

Cx=float(X1-X2+X3-X4)
Cy=float(Y1-Y2+Y3-Y4)

tsi=float(Dx/Ax);
eta=float(Dy/By);
assert abs(tsi)<=1 and abs(eta)<=1

N1=float(0.25*(1-tsi)*(1-eta))
N2=float(0.25*(1+tsi)*(1-eta))
N3=float(0.25*(1+tsi)*(1+eta))
N4=float(0.25*(1-tsi)*(1+eta))
return N1,N2,N3,N4

#macroscale nodal connectivity- need to give this as input- get from abaqus input(.inp) file of
macroscale mesh
#use data from .dat file named NC generated in the GP_RVE_place.py to define the nodal_connectivity
nodal_connectvity=np.array([[1, 2, 3, 4]])

# to loop around macroscale nodal connectivity


for i in range(0,int(np.shape(nodal_connectvity)[0])):
nodes=nodal_connectvity[i]
ele=i+1;#element number

#macroscale node x and y coordinates


X1=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[0])].xValue
X2=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[1])].xValue
X3=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[2])].xValue
X4=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[3])].xValue

Y1=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[0])].yValue
Y2=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[1])].yValue
Y3=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[2])].yValue
Y4=mdb.models['Model-1'].rootAssembly.features['RP-'+str(nodes[3])].yValue

for j in range(1,5):# loop to go around the gauss point RVEs in macroscale element no 'ele'
Face1 = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Face11']# Vertex11 and Vertex14
Face2 = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Face12']# Vertex11 and Vertex12
Face3 = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Face13']# Vertex12 and Vertex13
Face4 = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Face14']#Vertex13 and Vertex14

Vertex1 = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex11']#top right corner


Vertex2 = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex12']#top left corner
Vertex3 = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex13']#bottom left corner
Vertex4 = a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex14']#bottom right corner

#uncomment lines below for calculating the shape functions at the Gauss point if you want to
fix the Gauss point rather #than vertex 3 as done below
# GP=a.sets['Part-1-1-lin-'+str(ele)+'-1.GP'+str(j)]
# N1G,N2G,N3G,N4G=vertex_shape_fn(GP,X1,X2,X3,X4,Y1,Y2,Y3,Y4)
# calculating the shape functions at the mid-points of the edges
M1=a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.M11']
N1M1,N2M1,N3M1,N4M1=vertex_shape_fn(M1,X1,X2,X3,X4,Y1,Y2,Y3,Y4)

M2=a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.M12']
N1M2,N2M2,N3M2,N4M2=vertex_shape_fn(M2,X1,X2,X3,X4,Y1,Y2,Y3,Y4)

M3=a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.M13']
N1M3,N2M3,N3M3,N4M3=vertex_shape_fn(M3,X1,X2,X3,X4,Y1,Y2,Y3,Y4)

M4=a.sets['Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.M14']
N1M4,N2M4,N3M4,N4M4=vertex_shape_fn(M4,X1,X2,X3,X4,Y1,Y2,Y3,Y4)

158
# for fixing the Gauss point if you don't want to fix the vertex 3
# mdb.models['Model-1'].Equation(name='Ele-'+str(ele)+'-Constraint_GP'+str(j)+'-1', terms=((1.0,
# 'Part-1-1-lin-'+str(ele)+'-1.GP'+str(j), 1), (-N1G, 'N'+str(nodes[0])+'-RP', 1), (-N2G,
'N'+str(nodes[1])+'-RP', 1), (-N3G,'N'+str(nodes[2])+'-RP', 1),
# (-N4G,'N'+str(nodes[3])+'-RP', 1)))
# mdb.models['Model-1'].Equation(name='Ele-'+str(ele)+'-Constraint_GP'+str(j)+'-2', terms=((1.0,
# 'Part-1-1-lin-'+str(ele)+'-1.GP'+str(j), 2), (-N1G, 'N'+str(nodes[0])+'-RP', 2), (-N2G,
'N'+str(nodes[1])+'-RP', 2), (-N3G,'N'+str(nodes[2])+'-RP', 2),
# (-N4G,'N'+str(nodes[3])+'-RP', 2)))
N1v3,N2v3,N3v3,N4v3=vertex_shape_fn(Vertex3,X1,X2,X3,X4,Y1,Y2,Y3,Y4)
#for fixing vertex13
mdb.models['Model-1'].Equation(name='Ele-'+str(ele)+'-Constraint_V3'+str(j)+'-1',
terms=((1.0,
'Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex13', 1), (-N1v3, 'N'+str(nodes[0])+'-RP',
1),
(-N2v3, 'N'+str(nodes[1])+'-RP', 1), (-N3v3,'N'+str(nodes[2])+'-RP', 1),
(-N4v3,'N'+str(nodes[3])+'-RP', 1)))
# fix y displacement of vertex13
mdb.models['Model-1'].Equation(name='Ele-'+str(ele)+'-Constraint_V3'+str(j)+'-2',
terms=((1.0,
'Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex13', 2), (-N1v3, 'N'+str(nodes[0])+'-RP',
2),
(-N2v3, 'N'+str(nodes[1])+'-RP', 2), (-N3v3,'N'+str(nodes[2])+'-RP', 2),
(-N4v3,'N'+str(nodes[3])+'-RP', 2)))

#V2-V3 constraints
mdb.models['Model-1'].Equation(name='ConstraintV2-V3-1'+'Ele'+str(ele)+'-RVE'+str(j),
terms=((1.0,
'Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex12', 1),(-1.0, 'Part-1-ele-'+str(ele)+'-GP-
'+str(j)+'.Vertex13', 1),
(-(N1M2-N1M4),'N'+str(nodes[0])+'-RP', 1), (-(N2M2-N2M4),'N'+str(nodes[1])+'-RP', 1), (-
(N3M2-N3M4),'N'+str(nodes[2])+'-RP', 1),
(-(N4M2-N4M4),'N'+str(nodes[3])+'-RP', 1)))

mdb.models['Model-1'].Equation(name='ConstraintV2-V3-2'+'Ele'+str(ele)+'-RVE'+str(j),
terms=((1.0,
'Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex12', 2), (-1.0, 'Part-1-ele-'+str(ele)+'-GP-
'+str(j)+'.Vertex13', 2),
(-(N1M2-N1M4),'N'+str(nodes[0])+'-RP', 2), (-(N2M2-N2M4),'N'+str(nodes[1])+'-RP', 2), (-
(N3M2-N3M4),'N'+str(nodes[2])+'-RP', 2),
(-(N4M2-N4M4),'N'+str(nodes[3])+'-RP', 2)))

#V1-V3 constraints
mdb.models['Model-1'].Equation(name='ConstraintV1-V3-1'+'Ele'+str(ele)+'-RVE'+str(j),
terms=((1.0,
'Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex11', 1), (-1.0, 'Part-1-ele-'+str(ele)+'-GP-
'+str(j)+'.Vertex13', 1),
(-(N1M1-N1M3)-(N1M2-N1M4), 'N'+str(nodes[0])+'-RP', 1), (-(N2M1-N2M3)-
(N2M2-N2M4), 'N'+str(nodes[1])+'-RP', 1),
(-(N3M1-N3M3)-(N3M2-N3M4),'N'+str(nodes[2])+'-RP', 1),(-(N4M1-N4M3)-
(N4M2-N4M4),'N'+str(nodes[3])+'-RP', 1)))
mdb.models['Model-1'].Equation(name='ConstraintV1-V3-2'+'Ele'+str(ele)+'-RVE'+str(j),
terms=((1.0,
'Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex11', 2),(-1.0, 'Part-1-ele-'+str(ele)+'-GP-
'+str(j)+'.Vertex13', 2),
(-(N1M1-N1M3)-(N1M2-N1M4), 'N'+str(nodes[0])+'-RP', 2), (-(N2M1-N2M3)-
(N2M2-N2M4), 'N'+str(nodes[1])+'-RP', 2),
(-(N3M1-N3M3)-(N3M2-N3M4),'N'+str(nodes[2])+'-RP', 2),(-(N4M1-N4M3)-
(N4M2-N4M4),'N'+str(nodes[3])+'-RP', 2)))

#V4-V3 constraints
mdb.models['Model-1'].Equation(name='ConstraintV4-V3-1'+'Ele'+str(ele)+'-RVE'+str(j),
terms=((1.0,
'Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex14', 1),(-1.0, 'Part-1-ele-'+str(ele)+'-GP-
'+str(j)+'.Vertex13', 1),
(-(N1M1-N1M3),'N'+str(nodes[0])+'-RP', 1), (-(N2M1-N2M3),'N'+str(nodes[1])+'-RP', 1), (-
(N3M1-N3M3),'N'+str(nodes[2])+'-RP', 1),

159
(-(N4M1-N4M3),'N'+str(nodes[3])+'-RP', 1)))

mdb.models['Model-1'].Equation(name='ConstraintV4-V3-2'+'Ele'+str(ele)+'-RVE'+str(j),
terms=((1.0,
'Part-1-ele-'+str(ele)+'-GP-'+str(j)+'.Vertex14', 2), (-1.0, 'Part-1-ele-'+str(ele)+'-GP-
'+str(j)+'.Vertex13', 2),
(-(N1M1-N1M3),'N'+str(nodes[0])+'-RP', 2), (-(N2M1-N2M3),'N'+str(nodes[1])+'-RP', 2), (-
(N3M1-N3M3),'N'+str(nodes[2])+'-RP', 2),
(-(N4M1-N4M3),'N'+str(nodes[3])+'-RP', 2)))

# pairing the Edge nodes of each RVE


ParingFaces13 = []
ParingFaces13.append(TakeVertexOut(SortListOfNodes(Face1,1)))
ParingFaces13.append(TakeVertexOut(SortListOfNodes(Face3,1)))
# Face 2 and 4 (top-bottom)
ParingFaces24 = []
ParingFaces24.append(TakeVertexOut(SortListOfNodes(Face2,0)))
ParingFaces24.append(TakeVertexOut(SortListOfNodes(Face4,0)))

# loop to define sets on the vertical edges Face11 and Face 13 and periodic MPCs
for ii in range(len(ParingFaces13[0])):

# sets definition
a.Set(name='Ele'+str(ele)+'-RVE'+str(j)+'-Face13-mas_node-'+str(ParingFaces13[0][ii]+1)+'-
'+str(ii), nodes=(
a.instances['Part-1-ele-'+str(ele)+'-GP-
'+str(j)].nodes[ParingFaces13[0][ii]:ParingFaces13[0][ii]+1],))
a.Set(name='Ele'+str(ele)+'-RVE'+str(j)+'-Face13-sla_node-'+str(ParingFaces13[1][ii]+1)+'-
'+str(ii), nodes=(
a.instances['Part-1-ele-'+str(ele)+'-GP-
'+str(j)].nodes[ParingFaces13[1][ii]:ParingFaces13[1][ii]+1],))

mdb.models['Model-1'].Equation(name='SEle-'+str(ele)+'gRVE'+str(j)+'-
Constraint_Face13_nodes-1-'+str(ii), terms=((1.0,

'Ele'+str(ele)+'-RVE'+str(j)+'-Face13-mas_node-'+str(ParingFaces13[0][ii]+1)+'-'+str(ii), 1), (-
1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face13-sla_node-'+str(ParingFaces13[1][ii]+1)+'-'+str(ii), 1),
(-(N1M1-N1M3), 'N'+str(nodes[0])+'-RP', 1), (-(N2M1-N2M3), 'N'+str(nodes[1])+'-RP', 1), (-
(N3M1-N3M3),'N'+str(nodes[2])+'-RP', 1),
(-(N4M1-N4M3),'N'+str(nodes[3])+'-RP', 1)))

# y dsiplacement
mdb.models['Model-1'].Equation(name='SEle-'+str(ele)+'gRVE'+str(j)+'-
Constraint_Face13_nodes-2-'+str(ii), terms=((1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face13-mas_node-'+str(ParingFaces13[0][ii]+1)+'-'+str(ii), 2), (-
1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face13-sla_node-'+str(ParingFaces13[1][ii]+1)+'-'+str(ii), 2),
(-(N1M1-N1M3), 'N'+str(nodes[0])+'-RP', 2), (-(N2M1-N2M3), 'N'+str(nodes[1])+'-RP', 2), (-
(N3M1-N3M3),'N'+str(nodes[2])+'-RP', 2),
(-(N4M1-N4M3),'N'+str(nodes[3])+'-RP', 2)))

# loop to define sets on the vertical edges Face12 and Face 14 and periodic MPCs on these vertical
edges
for ii in range(len(ParingFaces24[0])):

# sets definition
a.Set(name='Ele'+str(ele)+'-RVE'+str(j)+'-Face24-mas_node-'+str(ParingFaces24[0][ii]+1)+'-
'+str(ii), nodes=(
a.instances['Part-1-ele-'+str(ele)+'-GP-
'+str(j)].nodes[ParingFaces24[0][ii]:ParingFaces24[0][ii]+1],))
a.Set(name='Ele'+str(ele)+'-RVE'+str(j)+'-Face24-sla_node-'+str(ParingFaces24[1][ii]+1)+'-
'+str(ii), nodes=(
a.instances['Part-1-ele-'+str(ele)+'-GP-
'+str(j)].nodes[ParingFaces24[1][ii]:ParingFaces24[1][ii]+1],))

160
mdb.models['Model-1'].Equation(name='SEle-'+str(ele)+'gRVE'+str(j)+'-
Constraint_Face24_nodes-1-'+str(ii), terms=((1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face24-mas_node-'+str(ParingFaces24[0][ii]+1)+'-'+str(ii), 1), (-
1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face24-sla_node-'+str(ParingFaces24[1][ii]+1)+'-'+str(ii), 1),
(-(N1M2-N1M4), 'N'+str(nodes[0])+'-RP', 1), (-(N2M2-N2M4), 'N'+str(nodes[1])+'-RP', 1), (-
(N3M2-N3M4),'N'+str(nodes[2])+'-RP', 1),
(-(N4M2-N4M4),'N'+str(nodes[3])+'-RP', 1)))
# y-displacement
mdb.models['Model-1'].Equation(name='SEle-'+str(ele)+'gRVE'+str(j)+'-
Constraint_Face24_nodes-2-'+str(ii), terms=((1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face24-mas_node-'+str(ParingFaces24[0][ii]+1)+'-'+str(ii), 2), (-
1.0,
'Ele'+str(ele)+'-RVE'+str(j)+'-Face24-sla_node-'+str(ParingFaces24[1][ii]+1)+'-'+str(ii), 2),
(-(N1M2-N1M4), 'N'+str(nodes[0])+'-RP', 2), (-(N2M2-N2M4), 'N'+str(nodes[1])+'-RP', 2), (-
(N3M2-N3M4),'N'+str(nodes[2])+'-RP', 2),
(-(N4M2-N4M4),'N'+str(nodes[3])+'-RP', 2)))

---end of code---

161
APPENDIX E: DEFINING THE SCALE TRANSITION
RELATIONSHIP (L MATRIX) THROUGH MPCS

The scale transition relationship in Direct FE2 is carried out by linking the
microscale nodal displacements (degrees of freedom) d ̃ with macroscale nodal
displacements d which is the L matrix in equation (2.17). Defining L in the form
as shown in equation (2.17) is difficult as the number of microscale degrees of
freedom outnumber the macroscale degrees of freedom. It is much easier to
define the inverse relation expressed by equation (2.17) that is relating the
microscale degrees of freedom though a transformation matrix to the less
numerous macroscale degrees of freedom. This is carried out by defining MPC
equations (2.20) and (2.25) for linear displacement boundary conditions and
equations (2.23), (2.24), (2.26) and (2.27) for periodic boundary conditions on
the RVE’s boundary dofs. The structure of the L matrix for a single macroscale
element (shown in Figure E.1) will be discussed after elaborating upon the
definition of the MPCs required to set up the Direct FE2 models as stated above
and referred to in section 2.3.2 and Figure D.1.

(a) (b)

Figure E.1 A macroscale element (with 4 nodes and corresponding 8 degrees


of freedom labelled) with simple 4 element microscale RVEs- R1,R2,R3 and R4
for demonstrating the scale transition relationship with MPCs (b) The RVE
Ri (with labelled 8 nodes and corresponding 16 degrees of freedom on the
RVE boundaries) inside the macroscale element in Figure E.1(a) needed to
define the scale transition relationship. The internal centre node is not
involved in the scale transition relationship.

162
Linear Displacement Boundary Conditions
The scale transition relationship for the macroscale element with linear
displacement boundary conditions imposed on the RVEs in Figure E.1 can be
written as
̃ = 𝑴𝑼
𝑼
[64x1] [64x8] [8x1]
(E.1)

The RVE boundary degrees of freedom 𝑼 ̃ 𝑎𝑛𝑑 𝑴 can be written by splitting the
equations as per the RVEs involved as shown below
̃ 𝑹𝟏
𝑼 𝑴𝑹𝟏
̃ 𝑹𝟐 𝑹𝟐
̃ = [𝑼 ]; 𝑴 = [𝑴 ]
𝑼
̃
𝑼 𝑹𝟑
𝑴𝑹𝟑
̃ 𝑹𝟒
𝑼 𝑴𝑹𝟒
(E.2)
̃ 𝑹𝒊
Each 𝑼 is a 16x1 matrix which makes 𝑼 ̃ a 64x1 matrix and each 𝑴 is a 𝑹𝒊

16x8 matrix which makes M a 64x8 matrix. 𝑼𝑰 (8x1 matrix) contains the
macroscale degrees of freedom as shown below
𝑢1
𝑣1
𝑢2
𝑣2
𝑼= 𝑢
3
𝑣3
𝑢4
[ 𝑣4 ]
(E.3)
For each RVE Ri, equation (E.1) can be written as

̃ 𝑹𝒊 = 𝑴𝑹𝒊 𝑼
𝑼
𝑅𝑖
𝑢̃𝑣3 𝑅𝑖
𝑁1 (𝑥𝑣3 ) 0 𝑁2 (𝑥𝑣3 𝑅𝑖
) 0 𝑅𝑖
𝑁3 (𝑥𝑣3 ) 0 𝑁4 (𝑥𝑣3 𝑅𝑖
) 0
𝑅𝑖 𝑅𝑖 𝑅𝑖 𝑅𝑖 𝑅𝑖
𝑣̃𝑣3 0 𝑁 1 (𝑥𝑣3 ) 0 𝑁 2 (𝑥 𝑣3 ) 0 𝑁 3 (𝑥 𝑣3 ) 0 𝑁4 (𝑥𝑣3 )
𝑅𝑖
𝑢̃𝑣1 𝑅𝑖
𝑁1 (𝑥𝑣1 ) 0 𝑁2 (𝑥𝑣1 𝑅𝑖
) 0 𝑅𝑖
𝑁3 (𝑥𝑣1 ) 0 𝑁4 (𝑥𝑣1 𝑅𝑖
) 0
𝑅𝑖 𝑅𝑖 𝑅𝑖 𝑅𝑖 𝑅𝑖
𝑣̃𝑣1 0 𝑁 (𝑥
1 𝑣1 ) 0 𝑁 (𝑥
2 𝑣1 ) 0 𝑁 (𝑥
3 𝑣1 ) 0 𝑁4 (𝑥𝑣1 )
𝑅𝑖
𝑢̃𝑣2 𝑅𝑖
𝑁1 (𝑥𝑣2 ) 0 𝑁2 (𝑥𝑣2 𝑅𝑖
) 0 𝑅𝑖
𝑁3 (𝑥𝑣2 ) 0 𝑁4 (𝑥𝑣2 𝑅𝑖
) 0 𝑢1
𝑅𝑖 𝑅𝑖 𝑅𝑖 𝑅𝑖 𝑅𝑖
𝑣̃𝑣2 0 𝑁1 (𝑥𝑣2 ) 0 𝑁2 (𝑥𝑣2 ) 0 𝑁3 (𝑥𝑣2 ) 0 𝑁4 (𝑥𝑣2 ) 𝑣1
𝑅𝑖
𝑢̃𝑣4 𝑁1 (𝑥𝑣4 ) 0 𝑁2 (𝑥𝑣4 )
𝑅𝑖 𝑅𝑖 0 𝑁3 (𝑥𝑣4 ) 0 𝑁4 (𝑥𝑣4 ) 0
𝑅𝑖 𝑅𝑖 𝑢2
𝑁 (𝑥 𝑅𝑖
) 𝑁 (𝑥 𝑅𝑖
) 𝑅𝑖 𝑅𝑖 𝑣2
0 𝑁3 (𝑥𝑣4 ) 0 𝑁4 (𝑥𝑣4 )
𝑅𝑖
𝑣̃𝑣4 0 1 𝑣4 0 2 𝑣4
= 𝑢3
𝑢̃𝑎𝑅𝑖 𝑁1 (𝑥𝑎𝑅𝑖 ) 0 𝑁2 (𝑥𝑎𝑅𝑖 ) 0 𝑁3 (𝑥𝑎𝑅𝑖 ) 0 𝑁4 (𝑥𝑎𝑅𝑖 ) 0
(𝑥 𝑅𝑖 ) (𝑥 𝑅𝑖 ) (𝑥 𝑅𝑖 ) 𝑅𝑖 𝑣3
𝑣̃𝑎𝑅𝑖 0 𝑁 1 𝑎 0 𝑁 2 𝑎 0 𝑁3 𝑎 0 𝑁 4 𝑎 )
(𝑥
0 0 𝑁3 (𝑥𝑏𝑅𝑖 ) 0 0 𝑢4
𝑢̃𝑏𝑅𝑖 𝑁1 (𝑥𝑏𝑅𝑖 ) 𝑁2 (𝑥𝑏𝑅𝑖 ) 𝑁4 (𝑥𝑏𝑅𝑖 )
𝑁 (𝑥 𝑅𝑖
) 𝑁 (𝑥 𝑅𝑖
) 𝑁 (𝑥 𝑅𝑖
) 𝑁 𝑅𝑖 [ 𝑣4 ]
𝑣̃𝑏𝑅𝑖 0 1 𝑏 0 2 𝑏 0 3 𝑏 0 4 𝑏 )
(𝑥
𝑢̃𝑐𝑅𝑖 𝑁1 (𝑥𝑐𝑅𝑖 ) 0 𝑁2 (𝑥𝑐𝑅𝑖 ) 0 𝑁3 (𝑥𝑐𝑅𝑖 ) 0 𝑁4 (𝑥𝑐𝑅𝑖 ) 0
(𝑥 𝑅𝑖 ) (𝑥 𝑅𝑖 ) (𝑥 𝑅𝑖 ) 𝑅𝑖
𝑣̃𝑐𝑅𝑖 0 𝑁1 𝑐 0 𝑁 2 𝑐 0 𝑁 3 𝑐 0 𝑁4 𝑐 )
(𝑥
𝑢̃𝑑𝑅𝑖 𝑁1 (𝑥𝑑𝑅𝑖 ) 0 𝑁2 (𝑥𝑑𝑅𝑖 ) 0 𝑁3 (𝑥𝑑𝑅𝑖 ) 0 𝑁4 (𝑥𝑑𝑅𝑖 ) 0
𝑅𝑖 𝑅𝑖 𝑅𝑖 𝑅𝑖
[𝑣̃𝑑𝑅𝑖 ] [ 0 𝑁 (𝑥
1 𝑑 ) 0 𝑁 (𝑥
2 𝑑 ) 0 𝑁 (𝑥
3 𝑑 ) 0 𝑁4 𝑑 )]
(𝑥
(E.4)

163
Equation (E.4) is the set of MPCs required to apply linear displacement
boundary conditions for a RVE Ri in the macroscale element in Figure E.1 (a).
Equation (E.1) is the complete set of MPC equations for all the RVEs of the
macroscale element. The L matrix can be obtained by transforming M in
equation (E.1).

Now the L matrix for linear displacement boundary conditions for the single
macroscale element can be obtained by choosing any 8 microscale degrees of
freedom which is equivalent to selecting any 8 equations in (E.1) and then
inverting the 𝑴𝒄 in the resulting system 𝑼 ̃𝒄 = 𝑴𝒄 𝑼 to obtain the element L
matrix corresponding to the microscale degrees of freedom in 𝑼 ̃𝒄 . As stated
above due to large number of microscale degrees of freedom, the L matrix is
not unique and depends on 𝑼 ̃𝒄 . Due to this large difference in size of 𝑼
̃ and U,
the global L matrix is sparse.
Periodic Boundary conditions

The MPCs for Periodic Boundary Conditions (PBCs) can be defined similar to
equation (E.1) as shown below
̃𝑷 = 𝑴𝑷 𝑼
𝑼
[48x1] [48x8] [8x1]
(E.5)

using the equations (2.23) and (2.24) modified as per Figure E.1(b) and given
below in equation (E.6)

̃ 𝑏 = (𝑁𝐼 (𝒙𝑑 ) − 𝑁𝐼 (𝒙𝑏 ))𝒖𝐼


̃𝑡 − 𝒖
𝒖

̃ 𝑙 = (𝑁𝐼 (𝒙𝑐 ) − 𝑁𝐼 (𝒙𝑎 ))𝒖𝐼


̃𝑟 − 𝒖
𝒖

(E.6)

Apart from equations (E.6), one node needs to be given constraints to prevent
rigid body motion of the RVEs while applying PBCs. For the macroscale
element in Figure E.1, the bottom left node (Vertex 3) of each RVE Ri is
constrained as per equation (2.20) as can be seen in the first two rows in the
system of equations (E.8). The vertices where both equations in (E.6) need to
be applied at the same node need to be handled separately as given in (E.7).
𝑢̃𝑣1 – 𝑢̃𝑣3 – ∆𝒖 ̃ 𝑑/𝑏 = 0
̃ 𝑐/𝑎 − ∆𝒖

𝑢̃𝑣2 – 𝑢̃𝑣3 − ∆𝒖
̃ 𝑑/𝑏 = 0

𝑢̃𝑣4 – 𝑢̃𝑣3 – ∆𝒖
̃ 𝑐/𝑎 = 0

̃ 𝑑/𝑏 = (𝑁𝐼 (𝒙𝑑 ) − 𝑁𝐼 (𝒙𝑏 ))𝒖𝐼


∆𝒖

̃ 𝑐/𝑎 = (𝑁𝐼 (𝒙𝑐 ) − 𝑁𝐼 (𝒙𝑎 ))𝒖𝐼


∆𝒖

(E.7)

164
where ∆𝒖 ̃ d/𝑏 𝑎𝑛𝑑 ∆𝒖̃ c/a are equivalent to equations (2.21) and (2.22)
respectively and rewritten as per Figure E.1 . The system of MPCs for a single
RVE Ri in the macroscale element in Figure E.1(a) is given in equation (E.8)
below.
𝑅𝑖
𝑢̃𝑣3
𝑅𝑖
𝑅𝑖
𝑁1 (𝑥𝑣3 ) 0 𝑁2 (𝑥𝑣3 𝑅𝑖
) 0 𝑁3 (𝑥𝑣3 𝑅𝑖
) 0 𝑁4 (𝑥𝑣3 𝑅𝑖
) 0
𝑣̃𝑣3 0 𝑁 (𝑥
1 𝑣3
𝑅𝑖
) 0 𝑁 (𝑥
2 𝑣3
𝑅𝑖
) 0 𝑁 (𝑥
3 𝑣3
𝑅𝑖
) 𝑅𝑖
0 𝑁4 (𝑥𝑣3 )
𝑅𝑖 𝑅𝑖
𝑢̃𝑣1 − 𝑢̃𝑣3 0 0 0 0
𝑅𝑖 𝑅𝑖 𝑢1
𝑣̃𝑣1 − 𝑣̃𝑣3 0 0 0 0 𝑣1
𝑢̃𝑣2𝑅𝑖 𝑅𝑖
− 𝑢̃𝑣3 0 0 0 0 𝑢2
𝑅𝑖
𝑣̃𝑣2 − 𝑣̃𝑣3𝑅𝑖
0 0 0 0 𝑣2
𝑅𝑖 𝑅𝑖
= 0 0 0 0 𝑢3
𝑢̃𝑣4 − 𝑢̃𝑣3
𝑅𝑖 𝑅𝑖 𝑣3
𝑣̃𝑣4 − 𝑣̃𝑣3 0 0 0 0
0 0 0 0 𝑢4
𝑢̃𝑐𝑅𝑖 − 𝑢̃𝑎𝑅𝑖 [ 𝑣4 ]
𝑣̃𝑐𝑅𝑖 − 𝑣̃𝑎𝑅𝑖 0 0 0 0
𝑅𝑖
𝑢̃𝑑 − 𝑢̃𝑏 𝑅𝑖 0 0 0 0
[ 𝑣̃𝑑𝑅𝑖 − 𝑣̃𝑏𝑅𝑖 ] [ 0 0 0 0 ]
(E.8)

The empty boxes are filled as per equations (E.6) and (E.7) and each entry
will be the difference of the shape functions. Equations (2.26) and (2.27) should
be used in case gradients of the shape functions are employed. Equation (E.8)
is the scale transition relationship that is needed to be defined as MPCs for
solving Direct FE2 problems with PBCs.

For obtaining the L matrix for macroscale element in Figure E.1 with PBCs, 8
equations (rows) in equation (E.5) need to be chosen so that the 8 macroscale
degrees of freedom can be replaced with 8(or more) microscale degrees of
freedom. If the equations of the fixed vertex v3 of every RVE Ri, (as shown in
first 2 rows of the matrices in equation (E.8)) are chosen resulting in 𝑼̃ 𝑷𝒄 =
𝑴𝑷𝒄 𝑼. For the macroscale element (in Figure E1) employing PBCs, L is
𝑅𝑖 𝑅𝑖
obtained by inverting 𝑴𝑷𝒄 effectively mapping 𝑼 (equation (E.3)) onto 𝑢̃𝑣3 , 𝑣̃𝑣3
{i=1,2,3,4}.This means that if the vertex 3 of every RVE is chosen in equation
(E.1), the L matrix for the macroscale element employing linear displacement
boundary condition and periodic boundary conditions is similar.

However, it is to be noted that the L matrix is not required to define the MPCs,
its inverse in the form of MPCs (matrix M) as demonstrated above is sufficient
to set up a Direct FE2 model. Since the number of microscale degrees of freedom
outnumber the macroscale degrees of freedom, the L matrix defined in equation
(2.17) is not unique and is sparse and depends on the choice of microscale
degrees of freedom to calculate the L matrix.

165
APPENDIX F: COUPLING DIRECT FE2 WITH FULL FE REGIONS

The problem discussed in section 3.2.4 is carried out with the regions close to
the fixed end of the beam top and bottom corners replaced with full FE regions
as shown in Figure F.1. The edges of the full FE region are ensured to remain
straight using the approach in Appendix-B so that macroscale element
boundaries are not violated.

(a)

(b)

Figure F.1 (a) Cantilever beam with a 8x80 Direct FE2 macroscale element
mesh with regions close to the fixed(left) end of the beam top and bottom
corners replaced with full FE regions (b) top left corner of the coupled FE2-
full FE beam model.

The strain contours in top left edge from full FE (same as Figure 3.14(d)) and
coupled FE-FE2 models are compared in Figure F.2 and the force displacement
behavior of the coupled model is shown in Figure F.3. The elements closed to
the fixed end in the coupled model are not deforming to the extent observed in
the full FE model which might be due to the smaller size of the full FE region
and also the effect of the straight edge boundaries on the full FE region to uphold
the Direct FE2 macroscale element boundaries.

The coupled model runs further than the reference full FE model upto 910µm
displacement compared to the 500µm in which the full FE simulation was
terminated. This may be due to the fact that the fixed end is not fully modelled.
However both models show the same non-linearity. Ensuring a small full FE
region to remain straight edged in the interior of the beam so closed to the fixed
end might not be physically realistic and reasonable. The size of these full FE
regions and the coupling between the full FE regions and Direct FE2 macroscale
elements regarding the boundary constraints of the full FE region need to be
investigated further to use the coupled Direct FE2 –FE model with confidence.

166
(a)

(b)

Figure F.2 Equivalent plastic strain contours of RVEs closest to top left of
beam for (a) coupled full FE-FE2 model and (b) Excessive deformation of
elements in the full FE model simulation in grey at top left edge of beam.

3000
Force(N)

2000

1000

Full FE
FE2 with linear macroscale elements
FE2 with quadratic macroscale elements
coupled linear FE2-full FE model
0
0 1000 2000 3000 4000
Displacement(m)

Figure F.3 Load-displacement plots for large deformation of elastic-plastic


composite beam.

167
BIBLIOGRAPHY

[1] L. Mishnaevsky, K. Branner, H.N. Petersen, J. Beauson, M. McGugan,


B.F. Sørensen, Materials for Wind Turbine Blades: An Overview, Mater.
(Basel, Switzerland). 10 (2017) 1285. doi:10.3390/ma10111285.
[2] P. Brøndsted, H. Lilholt, A. Lystrup, COMPOSITE MATERIALS FOR
WIND POWER TURBINE BLADES, Annu. Rev. Mater. Res. 35 (2005)
505–538. doi:10.1146/annurev.matsci.35.100303.110641.
[3] P.S. Veers, T.D. Ashwill, H.J. Sutherland, D.L. Laird, D.W. Lobitz, D.A.
Griffin, J.F. Mandell, W.D. Musial, K. Jackson, M. Zuteck, Trends in the
design, manufacture and evaluation of wind turbine blades, Wind Energy
An Int. J. Prog. Appl. Wind Power Convers. Technol. 6 (2003) 245–259.
[4] G. Savage, Composite Materials Technology in Formula 1 Motor
Racing., Honda Racing F1 Team. 91 (2008) 92.
[5] C. Soutis, Fibre reinforced composites in aircraft construction, Prog.
Aerosp. Sci. 41 (2005) 143–151.
doi:https://doi.org/10.1016/j.paerosci.2005.02.004.
[6] S. Dutton, D. Kelly, A. Baker, Composite materials for aircraft
structures, American Institute of Aeronautics and Astronautics, 2004.
[7] A.B. Strong, Fundamentals of composites manufacturing: materials,
methods and applications, Society of Manufacturing Engineers, 2008.
[8] V.V.V.& E. V.Morozov, Advanced Mechanics of composite materials,
Elsevier, 2007.
[9] A. Fawcett, 777 empennage certification approach, in: Proc. 11th. Int.
Conf. Compos. Mater. Gold Coast, Aust. July, 1997, 1997: pp. 14–18.
[10] B. Cox, Q. Yang, In Quest of Virtual Tests for Structural Composites,
Science (80-. ). 314 (2006) 1102 LP – 1107.
doi:10.1126/science.1131624.
[11] C. González, J.J. Vilatela, J.M. Molina-Aldareguía, C.S. Lopes, J.
LLorca, Structural composites for multifunctional applications: Current
challenges and future trends, Prog. Mater. Sci. 89 (2017) 194–251.
doi:https://doi.org/10.1016/j.pmatsci.2017.04.005.
[12] J. LLorca, C. González, J.M. Molina-Aldareguía, C.S. Lópes, Multiscale
Modeling of Composites: Toward Virtual Testing … and Beyond, JOM.
65 (2013) 215–225. doi:10.1007/s11837-012-0509-8.
[13] A.R. Melro, P.P. Camanho, F.M. Andrade Pires, S.T. Pinho,
Micromechanical analysis of polymer composites reinforced by
unidirectional fibres: Part I – Constitutive modelling, Int. J. Solids Struct.
50 (2013) 1897–1905. doi:https://doi.org/10.1016/j.ijsolstr.2013.02.009.

168
[14] Z. Xia, Y. Chen, F. Ellyin, A meso/micro-mechanical model for damage
progression in glass-fiber/epoxy cross-ply laminates by finite-element
analysis, Compos. Sci. Technol. 60 (2000) 1171–1179.
[15] C. Grufman, F. Ellyin, Numerical modelling of damage susceptibility of
an inhomogeneous representative material volume element of polymer
composites, Compos. Sci. Technol. 68 (2008) 650–657.
[16] Y. Zhang, Z. Xia, F. Ellyin, Viscoelastic and damage analyses of fibrous
polymer laminates by micro/meso-mechanical modeling, J. Compos.
Mater. 39 (2005) 2001–2022.
[17] P.P. Camanho, C.G. Dávila, S.T. Pinho, L. Iannucci, P. Robinson,
Prediction of in situ strengths and matrix cracking in composites under
transverse tension and in-plane shear, Compos. Part A Appl. Sci. Manuf.
37 (2006) 165–176.
[18] A. Forghani, M. Shahbazi, N. Zobeiry, A. Poursartip, R. Vaziri, 6 - An
overview of continuum damage models used to simulate intralaminar
failure mechanisms in advanced composite materials, in: P.P. Camanho,
S.R.B.T.-N.M. of F. in A.C.M. Hallett (Eds.), Woodhead Publ. Ser.
Compos. Sci. Eng., Woodhead Publishing, 2015: pp. 151–173.
doi:https://doi.org/10.1016/B978-0-08-100332-9.00006-2.
[19] K. Williams, R. Vaziri, Fnite element analysis of the impact response of
CFRP composite plates, in: Tenth Int. Conf. Compos. Mater. V. Struct.,
1995: pp. 647–654.
[20] K. V Williams, R. Vaziri, Application of a damage mechanics model for
predicting the impact response of composite materials, Comput. Struct.
79 (2001) 997–1011.
[21] R. Talreja, A continuum mechanics characterization of damage in
composite materials, Proc. R. Soc. London. A. Math. Phys. Sci. 399
(1985) 195–216.
[22] P. Ladevèze, O. Allix, J.-F. Deü, D. Lévêque, A mesomodel for
localisation and damage computation in laminates, Comput. Methods
Appl. Mech. Eng. 183 (2000) 105–122.
[23] M. Kashtalyan, C. Soutis, Analysis of composite laminates with intra-
and interlaminar damage, Prog. Aerosp. Sci. 41 (2005) 152–173.
[24] P. Maimí, P.P. Camanho, J.A. Mayugo, C.G. Dávila, A continuum
damage model for composite laminates: Part I–Constitutive model,
Mech. Mater. 39 (2007) 897–908.
[25] F.P. van der Meer, N. Moës, L.J. Sluys, A level set model for
delamination – Modeling crack growth without cohesive zone or stress
singularity, Eng. Fract. Mech. 79 (2012) 191–212.
doi:https://doi.org/10.1016/j.engfracmech.2011.10.013.

169
[26] S.T. Pinho, G.M. Vyas, P. Robinson, Response and damage propagation
of polymer-matrix fibre-reinforced composites: Predictions for WWFE-
III Part A, J. Compos. Mater. 47 (2013) 2595–2612.
[27] Z. Yuan, J. Fish, Are the cohesive zone models necessary for
delamination analysis?, Comput. Methods Appl. Mech. Eng. 310 (2016)
567–604. doi:https://doi.org/10.1016/j.cma.2016.06.023.
[28] F.-K. Chang, L.B. Lessard, Damage Tolerance of Laminated Composites
Containing an Open Hole and Subjected to Compressive Loadings: Part
I—Analysis, J. Compos. Mater. 25 (1991) 2–43.
doi:10.1177/002199839102500101.
[29] Z. Hashin, Failure Criteria for Unidirectional Fiber Composites, J. Appl.
Mech. 47 (1980) 329–334. doi:10.1115/1.3153664.
[30] F.-K. Chang, K.-Y. Chang, A progressive damage model for laminated
composites containing stress concentrations, J. Compos. Mater. 21
(1987) 834–855.
[31] M.J. Hinton, P.D. Soden, Predicting failure in composite laminates: the
background to the exercise, Compos. Sci. Technol. 58 (1998) 1001–
1010. https://doi.org/10.1016/S0266-3538(98)00074-8.
[32] C. Davila, D. Ambur, D. McGowan, Analytical prediction of damage
growth in notched composite panels loaded in axial compression, in: 40th
Struct. Struct. Dyn. Mater. Conf. Exhib., 1999: p. 1435.
[33] M.J. Hinton, A.S. Kaddour, P.D. Soden, Evaluation of failure prediction
in composite laminates: background to ‘part B’of the exercise, Compos.
Sci. Technol. 62 (2002) 1481–1488.
[34] M.J. Hinton, A.S. Kaddour, P.D. Soden, Evaluation of failure prediction
in composite laminates: background to ‘part C’of the exercise, Compos.
Sci. Technol. 64 (2004) 321–327.
[35] C. Davila, N. Jaunky, S. Goswami, Failure criteria for FRP laminates in
plane stress, in: 44th AIAA/ASME/ASCE/AHS/ASC Struct. Struct. Dyn.
Mater. Conf., 2003: p. 1991.
[36] S.W. Tsai, E.M. Wu, A General Theory of Strength for Anisotropic
Materials, J. Compos. Mater. 5 (1971) 58–80.
doi:10.1177/002199837100500106.
[37] K. V Williams, R. Vaziri, A. Poursartip, A physically based continuum
damage mechanics model for thin laminated composite structures, Int. J.
Solids Struct. 40 (2003) 2267–2300. doi:https://doi.org/10.1016/S0020-
7683(03)00016-7.
[38] R. Xu, C. Bouby, Z. Hamid, T. Ben Zineb, H. Hu, M. Potier-Ferry, 3D
modeling of shape memory alloy fiber reinforced composites by
multiscale finite element method, 2018.
doi:10.1016/j.compstruct.2018.05.108.

170
[39] S. Nezamabadi, M. Potier-Ferry, H. Zahrouni, J. Yvonnet, Compressive
failure of composites: A computational homogenization approach,
Compos. Struct. 127 (2015) 60–68.
doi:https://doi.org/10.1016/j.compstruct.2015.02.042.
[40] F. Feyel, J.-L. Chaboche, Multi-scale non-linear FE2 analysis of
composite structures: damage and fiber size effects, Rev. Eur. Des
Éléments Finis. 10 (2001) 449–472.
doi:10.1080/12506559.2001.11869262.
[41] T. Herwig, W. Wagner, On a robust FE2 model for delamination analysis
in composite structures, Compos. Struct. 201 (2018) 597–607.
doi:https://doi.org/10.1016/j.compstruct.2018.06.033.
[42] J. Schröder, M. Labusch, M.-A. Keip, Algorithmic two-scale transition
for magneto-electro-mechanically coupled problems: FE2-scheme:
Localization and homogenization, Comput. Methods Appl. Mech. Eng.
302 (2016) 253–280. doi:https://doi.org/10.1016/j.cma.2015.10.005.
[43] F. Feyel, A multilevel finite element method (FE2) to describe the
response of highly non-linear structures using generalized continua,
Comput. Methods Appl. Mech. Eng. 192 (2003) 3233–3244.
doi:https://doi.org/10.1016/S0045-7825(03)00348-7.
[44] C. Gonzalez, J. LLorca, Multiscale modeling of fracture in fiber-
reinforced composites, Acta Mater. 54 (2006) 4171–4181.
[45] L. Wu, L. Noels, L. Adam, I. Doghri, A multiscale mean-field
homogenization method for fiber-reinforced composites with gradient-
enhanced damage models, Comput. Methods Appl. Mech. Eng. 233
(2012) 164–179.
[46] J.S. Mayes, A.C. Hansen, Composite laminate failure analysis using
multicontinuum theory, Compos. Sci. Technol. 64 (2004) 379–394.
[47] P. Ladeveze, Multiscale modelling and computational strategies for
composites, Int. J. Numer. Methods Eng. 60 (2004) 233–253.
[48] C.C. Chamis, P.L.N. Murthy, P.K. Gotsis, S.K. Mital, Telescoping
composite mechanics for composite behavior simulation, Comput.
Methods Appl. Mech. Eng. 185 (2000) 399–411.
[49] J. Fish, Q. Yu, Multiscale damage modelling for composite materials:
theory and computational framework, Int. J. Numer. Methods Eng. 52
(2001) 161–191.
[50] P. Raghavan, S. Ghosh, Adaptive Multi-Scale Computational Modeling
of Composite Materials, C. Comput. Model. Eng. Sci. 5 (2004) 151–170.
[51] P. Raghavan, S. Ghosh, Concurrent multi-scale analysis of elastic
composites by a multi-level computational model, Comput. Methods
Appl. Mech. Eng. 193 (2004) 497–538.
doi:https://doi.org/10.1016/j.cma.2003.10.007.

171
[52] S. Ghosh, J. Bai, P. Raghavan, Concurrent multi-level model for damage
evolution in microstructurally debonding composites, Mech. Mater. 39
(2007) 241–266. doi:https://doi.org/10.1016/j.mechmat.2006.05.004.
[53] V. Šmilauer, C.G. Hoover, Z.P. Bažant, F.C. Caner, A.M. Waas, K.W.
Shahwan, Multiscale simulation of fracture of braided composites via
repetitive unit cells, Eng. Fract. Mech. 78 (2011) 901–918.
doi:https://doi.org/10.1016/j.engfracmech.2010.10.013.
[54] T. Mori, K. Tanaka, Average stress in matrix and average elastic energy
of materials with misfitting inclusions, Acta Metall. 21 (1973) 571–574.
doi:https://doi.org/10.1016/0001-6160(73)90064-3.
[55] I. Doghri, A. Ouaar, Homogenization of two-phase elasto-plastic
composite materials and structures: Study of tangent operators, cyclic
plasticity and numerical algorithms, Int. J. Solids Struct. 40 (2003) 1681–
1712. doi:https://doi.org/10.1016/S0020-7683(03)00013-1.
[56] G.W. Milton, R. V Kohn, Variational bounds on the effective moduli of
anisotropic composites, J. Mech. Phys. Solids. 36 (1988) 597–629.
doi:https://doi.org/10.1016/0022-5096(88)90001-4.
[57] L.J. Walpole, On bounds for the overall elastic moduli of inhomogeneous
systems—I, J. Mech. Phys. Solids. 14 (1966) 151–162.
doi:https://doi.org/10.1016/0022-5096(66)90035-4.
[58] B. Budiansky, Micromechanics, Comput. Struct. 16 (1983) 3–12.
doi:https://doi.org/10.1016/0045-7949(83)90141-4.
[59] R. Hill, A self-consistent mechanics of composite materials, J. Mech.
Phys. Solids. 13 (1965) 213–222. doi:https://doi.org/10.1016/0022-
5096(65)90010-4.
[60] C. Miehe, J. Schröder, M. Becker, Computational homogenization
analysis in finite elasticity: material and structural instabilities on the
micro- and macro-scales of periodic composites and their interaction,
Comput. Methods Appl. Mech. Eng. 191 (2002) 4971–5005.
doi:https://doi.org/10.1016/S0045-7825(02)00391-2.
[61] J. Fish, A. Suvorov, V. Belsky, Hierarchical composite grid method for
global-local analysis of laminated composite shells, Appl. Numer. Math.
23 (1997) 241–258. doi:https://doi.org/10.1016/S0168-9274(96)00068-
2.
[62] T. Asada, N. Ohno, Fully implicit formulation of elastoplastic
homogenization problem for two-scale analysis, Int. J. Solids Struct. 44
(2007) 7261–7275. doi:https://doi.org/10.1016/j.ijsolstr.2007.04.007.
[63] F. Feyel, J.-L. Chaboche, FE2 multiscale approach for modelling the
elastoviscoplastic behaviour of long fibre SiC/Ti composite materials,
Comput. Methods Appl. Mech. Eng. 183 (2000) 309–330.
doi:https://doi.org/10.1016/S0045-7825(99)00224-8.

172
[64] J.A. Hernández, J. Oliver, A.E. Huespe, M.A. Caicedo, J.C. Cante, High-
performance model reduction techniques in computational multiscale
homogenization, Comput. Methods Appl. Mech. Eng. 276 (2014) 149–
189. doi:https://doi.org/10.1016/j.cma.2014.03.011.
[65] R.J.M. Smit, W.A.M. Brekelmans, H.E.H. Meijer, Prediction of the
mechanical behavior of nonlinear heterogeneous systems by multi-level
finite element modeling, Comput. Methods Appl. Mech. Eng. 155 (1998)
181–192. doi:https://doi.org/10.1016/S0045-7825(97)00139-4.
[66] Z. Yuan, J. Fish, Toward realization of computational homogenization in
practice, Int. J. Numer. Methods Eng. 73 (2007) 361–380.
doi:10.1002/nme.2074.
[67] A. Tchalla, S. Belouettar, A. Makradi, H. Zahrouni, An ABAQUS
toolbox for multiscale finite element computation, Compos. Part B Eng.
52 (2013) 323–333.
doi:https://doi.org/10.1016/j.compositesb.2013.04.028.
[68] E. Tikarrouchine, G. Chatzigeorgiou, F. Praud, B. Piotrowski, Y.
Chemisky, F. Meraghni, Three-dimensional FE2 method for the
simulation of non-linear, rate-dependent response of composite
structures, Compos. Struct. 193 (2018) 165–179.
doi:https://doi.org/10.1016/j.compstruct.2018.03.072.
[69] F. El Halabi, D. González, A. Chico, M. Doblaré, FE2 multiscale in linear
elasticity based on parametrized microscale models using proper
generalized decomposition, Comput. Methods Appl. Mech. Eng. 257
(2013) 183–202. doi:https://doi.org/10.1016/j.cma.2013.01.011.
[70] J. Schröder, A numerical two-scale homogenization scheme: the FE2-
method BT - Plasticity and Beyond: Microstructures, Crystal-Plasticity
and Phase Transitions, in: J. Schröder, K. Hackl (Eds.), Springer Vienna,
Vienna, 2014: pp. 1–64. doi:10.1007/978-3-7091-1625-8_1.
[71] B. Regener, C. Krempaszky, E. Werner, M. Stockinger, Thermo-
Mechanical FE2 Simulation Scheme for Abaqus, PAMM. 11 (2011)
547–548. doi:10.1002/pamm.201110263.
[72] I. Özdemir, W.A.M. Brekelmans, M.G.D. Geers, Computational
homogenization for heat conduction in heterogeneous solids, Int. J.
Numer. Methods Eng. 73 (2008) 185–204. doi:10.1002/nme.2068.
[73] C. Helfen, S. Diebels, Numerical Multiscale Modelling of Sandwich
Plates, 2012.
[74] F. Gruttmann, W. Wagner, A coupled two-scale shell model with
applications to layered structures, Int. J. Numer. Methods Eng. 94 (2013)
1233–1254. doi:10.1002/nme.4496.
[75] F. Otero, X. Martinez, S. Oller, O. Salomón, An efficient multi-scale
method for non-linear analysis of composite structures, Compos. Struct.
131(2015)707–719.

173
doi:https://doi.org/10.1016/j.compstruct.2015.06.006.
[76] V. Papadopoulos, M. Tavlaki, The impact of interfacial properties on the
macroscopic performance of carbon nanotube composites. A FE2-based
multiscale study, Compos. Struct. 136 (2016) 582–592.
doi:https://doi.org/10.1016/j.compstruct.2015.10.025.
[77] S. Nezamabadi, J. Yvonnet, H. Zahrouni, M. Potier-Ferry, A multilevel
computational strategy for handling microscopic and macroscopic
instabilities, Comput. Methods Appl. Mech. Eng. 198 (2009) 2099–2110.
doi:https://doi.org/10.1016/j.cma.2009.02.026.
[78] I. Temizer, T.I. Zohdi, A numerical method for homogenization in non-
linear elasticity, Comput. Mech. 40 (2007) 281–298.
[79] M. Stroeven, H. Askes, L.J. Sluys, Numerical determination of
representative volumes for granular materials, Comput. Methods Appl.
Mech. Eng. 193 (2004) 3221–3238.
[80] C. Pelissou, J. Baccou, Y. Monerie, F. Perales, Determination of the size
of the representative volume element for random quasi-brittle
composites, Int. J. Solids Struct. 46 (2009) 2842–2855.
doi:https://doi.org/10.1016/j.ijsolstr.2009.03.015.
[81] T. Kanit, S. Forest, I. Galliet, V. Mounoury, D. Jeulin, Determination of
the size of the representative volume element for random composites:
statistical and numerical approach, Int. J. Solids Struct. 40 (2003) 3647–
3679. doi:https://doi.org/10.1016/S0020-7683(03)00143-4.
[82] J. Ohser, F. Mücklich, Statistical Analysis of Microstructures in Material
Science, 2000.
[83] S. Müller, Homogenization of nonconvex integral functionals and
cellular elastic materials, Arch. Ration. Mech. Anal. 99 (1987) 189–212.
doi:10.1007/BF00284506.
[84] J.R. Willis, Variational and Related Methods for the Overall Properties
of Composites, in: C.-S.B.T.-A. in A.M. Yih (Ed.), Elsevier, 1981: pp.
1–78. doi:https://doi.org/10.1016/S0065-2156(08)70330-2.
[85] S. Nemat-Nasser, M. Hori, Micromechanics: overall properties of
heterogeneous materials, Elsevier, 1993.
[86] P.M. Suquet, Elements of homogenization theory for inelastic solid
mechanics, Homog. Tech. Compos. Media. (1987).
[87] A. Bensoussan, J.-L. Lions, G. Papanicolaou, T.K. Caughey, Asymptotic
Analysis of Periodic Structures, J. Appl. Mech. 46 (1979) 477.
[88] P. Ladevèze, O. Loiseau, D. Dureisseix, A micro–macro and parallel
computational strategy for highly heterogeneous structures, Int. J.
Numer. Methods Eng. 52 (2001) 121–138.

174
[89] J.T. Oden, K. Vemaganti, N. Moës, Hierarchical modeling of
heterogeneous solids, Comput. Methods Appl. Mech. Eng. 172 (1999) 3–
25.
[90] S. Ghosh, K. Lee, P. Raghavan, A multi-level computational model for
multi-scale damage analysis in composite and porous materials, Int. J.
Solids Struct. 38 (2001) 2335–2385. doi:https://doi.org/10.1016/S0020-
7683(00)00167-0.
[91] A. Ibrahimbegovic, D. Brancherie, Combined hardening and softening
constitutive model of plasticity: precursor to shear slip line failure,
Comput. Mech. 31 (2003) 88–100.
[92] C. Miehe, A. Koch, Computational micro-to-macro transitions of
discretized microstructures undergoing small strains, Arch. Appl. Mech.
72 (2002) 300–317. doi:10.1007/s00419-002-0212-2.
[93] Ł. Kaczmarczyk, C.J. Pearce, N. Bićanić, Scale transition and
enforcement of RVE boundary conditions in second-order computational
homogenization, Int. J. Numer. Methods Eng. 74 (2008) 506–522.
doi:10.1002/nme.2188.
[94] O. van der Sluis, P.J.G. Schreurs, W.A.M. Brekelmans, H.E.H. Meijer,
Overall behaviour of heterogeneous elastoviscoplastic materials: effect
of microstructural modelling, Mech. Mater. 32 (2000) 449–462.
doi:https://doi.org/10.1016/S0167-6636(00)00019-3.
[95] K. Terada, M. Hori, T. Kyoya, N. Kikuchi, Simulation of the multi-scale
convergence in computational homogenization approaches, Int. J. Solids
Struct. 37 (2000) 2285–2311. doi:https://doi.org/10.1016/S0020-
7683(98)00341-2.
[96] V. Kouznetsova, W.A.M. Brekelmans, F.P.T. Baaijens, An approach to
micro-macro modeling of heterogeneous materials, Comput. Mech. 27
(2001) 37–48. doi:10.1007/s004660000212.
[97] C. Miehe, C.G. Bayreuther, On multiscale FE analyses of heterogeneous
structures: from homogenization to multigrid solvers, Int. J. Numer.
Methods Eng. 71 (2007) 1135–1180. doi:10.1002/nme.1972.
[98] D. Perić, E.A. de Souza Neto, R.A. Feijóo, M. Partovi, A.J.C. Molina,
On micro-to-macro transitions for multi-scale analysis of non-linear
heterogeneous materials: unified variational basis and finite element
implementation, Int. J. Numer. Methods Eng. 87 (2011) 149–170.
doi:10.1002/nme.3014.
[99] Y.-C. Ng, Deriving Composite Lamina Properties from Laminate
Properties Using Classical Lamination Theory and Failure
Criteria, J. Compos. Mater. 39 (2005) 1295–1306.
doi:10.1177/0021998305050429.

175
[100] U.S.D. of Defense, Military Handbook–MIL‐HDBK‐17‐1F: Composite
Materials Handbook, Volume 1—Polymer Matrix Composites
Guidelines for Characterization of Structural Materials, (2002).
[101] H. Ghaemi, Z. Fawaz, Experimental evaluation of effective tensile
properties of laminated composites, Adv. Compos. Mater. 11 (2002)
223–237. doi:10.1163/156855102762506272.
[102] M.G.D. Geers, V.G. Kouznetsova, W.A.M. Brekelmans, Multi-scale
computational homogenization: Trends and challenges, J. Comput. Appl.
Math. 234 (2010) 2175–2182.
doi:https://doi.org/10.1016/j.cam.2009.08.077.
[103] I.M. Gitman, H. Askes, L.J. Sluys, Coupled-volume multi-scale
modelling of quasi-brittle material, Eur. J. Mech. - A/Solids. 27 (2008)
302–327. doi:https://doi.org/10.1016/j.euromechsol.2007.10.004.
[104] A. Ibrahimbegović, D. Markovič, Strong coupling methods in multi-
phase and multi-scale modeling of inelastic behavior of heterogeneous
structures, Comput. Methods Appl. Mech. Eng. 192 (2003) 3089–3107.
doi:https://doi.org/10.1016/S0045-7825(03)00342-6.
[105] D. Markovic, A. Ibrahimbegovic, On micro–macro interface conditions
for micro scale based FEM for inelastic behavior of heterogeneous
materials, Comput. Methods Appl. Mech. Eng. 193 (2004) 5503–5523.
doi:https://doi.org/10.1016/j.cma.2003.12.072.
[106] F. Feyel, Multiscale FE2 elastoviscoplastic analysis of composite
structures, Comput. Mater. Sci. 16 (1999) 344–354.
doi:https://doi.org/10.1016/S0927-0256(99)00077-4.
[107] J.N. REDDY, An Introduction to The Finite Element Method, (n.d.).
[108] J.F. Abel, M.S. Shephard, An algorithm for multipoint constraints in
finite element analysis, Int. J. Numer. Methods Eng. 14 (1979) 464–467.
doi:10.1002/nme.1620140312.
[109] R. Hill, The Elastic Behaviour of a Crystalline Aggregate, Proc. Phys.
Soc. Sect. A. 65 (1952) 349–354. doi:10.1088/0370-1298/65/5/307.
[110] Abaqus 6.14 Analysis User’s Guide, (n.d.).
[111] P.D. Soden, M.J. Hinton, A.S. Kaddour, Lamina properties, lay-up
configurations and loading conditions for a range of fibre reinforced
composite laminates, in: Fail. Criteria Fibre-Reinforced-Polymer
Compos., Elsevier, 2004: pp. 30–51.
[112] K.D. Mishra, R.F. El-Hajjar, Non-linear strain invariant failure approach
for fibre reinforced composite materials, Int. J. Mater. Struct. Integr. 6
(2012) 284–296. doi:10.1504/IJMSI.2012.049961.
[113] W.R. Ramberg, W. and Osgood, Description of Stress-Strain Curves by
Three Parameters. Technical Note No. 902, 1943.
http://www.apesolutions.com/spd/public/NACA-TN902.pdf.

176
[114] M.H. Datoo, Mechanics of Fibrous Composites, Springer Netherlands,
2012. 10.1007/978-94-011-3670-9.
[115] B. Fiedler, M. Hojo, S. Ochiai, K. Schulte, M. Ando, Failure behavior of
an epoxy matrix under different kinds of static loading, Compos. Sci.
Technol. 61 (2001) 1615–1624. doi:https://doi.org/10.1016/S0266-
3538(01)00057-4.
[116] Simulia., Abaqus Analysis User’s Manual, 22.7.1 Time domain
vicoelasticity, 6.14 edition, 2014, n.d.
[117] E.J. Barbero, Viscoelasticity, in: Abaqus, Finite Elem. Anal. Compos.
Mater. Using Abaqus, CRC Press Boca Raton,FL, 2013: pp. 249–280.
[118] K. Park, G.H. Paulino, Cohesive zone models: a critical review of
traction-separation relationships across fracture surfaces, Appl. Mech.
Rev. 64 (2011) 60802.
[119] A. Corigliano, Formulation, identification and use of interface models in
the numerical analysis of composite delamination, Int. J. Solids Struct.
30 (1993) 2779–2811. doi:https://doi.org/10.1016/0020-7683(93)90154-
Y.
[120] P.P. Camanho, C.G. Davila, M.F. De Moura, Numerical simulation of
mixed-mode progressive delamination in composite materials, J.
Compos. Mater. 37 (2003) 1415–1438.
[121] W.C. Cui, M.R. Wisnom, M. Jones, A comparison of failure criteria to
predict delamination of unidirectional glass/epoxy specimens waisted
through the thickness, Composites. 23 (1992) 158–166.
[122] X. Lu, M. Ridha, B.Y. Chen, V.B.C. Tan, T.E. Tay, On cohesive element
parameters and delamination modelling, Eng. Fract. Mech. 206 (2019)
278–296. doi:https://doi.org/10.1016/j.engfracmech.2018.12.009.
[123] C. González, J. LLorca, Mechanical behavior of unidirectional fiber-
reinforced polymers under transverse compression: Microscopic
mechanisms and modeling, Compos. Sci. Technol. 67 (2007) 2795–
2806. doi:https://doi.org/10.1016/j.compscitech.2007.02.001.
[124] V. Murti, S. Valliappan, Numerical inverse isoparametric mapping in
remeshing and nodal quantity contouring, Comput. Struct. 22 (1986)
1011–1021. doi:https://doi.org/10.1016/0045-7949(86)90161-6.
[125] M.L.C. Jones, D. Hull, Microscopy of failure mechanisms in filament-
wound pipe, J. Mater. Sci. 14 (1979) 165–174. doi:10.1007/bf01028340.
[126] M. Carroll, F. Ellyin, D. Kujawski, A.S. Chiu, The rate-dependent
behaviour of ± 55 ° filament-wound glass-fibre/epoxy tubes under biaxial
loading, Compos. Sci. Technol. 55 (1995) 391–403.
doi:http://dx.doi.org/10.1016/0266-3538(95)00119-0.

177
[127] J. Bai, G. Hu, P. Bompard, Mechanical behaviour of ± 55 ° filament-
wound glass-fibre/epoxy-resin tubes: II. Micromechanical model of
damage initiation and the competition between different mechanisms,
Compos. Sci. Technol. 57 (1997) 155–164.
doi:http://dx.doi.org/10.1016/S0266-3538(96)00125-X.
[128] D.W.J. and M.F. Card, Torsional Shear Strength of Filament-Wound
Glass-Epoxy Tubes, 1971.
[129] Y. Zhao, S.S. Pang, Stress-Strain and Failure Analyses of Composite
Pipe Under Torsion, J. Press. Vessel Technol. 117 (1995) 273–278.
doi:10.1115/1.2842123.
[130] J. Highton, A.B. Adeoye, P.D. Soden, Fracture stresses for ± 75 degree
filament wound grp tubes under biaxial loads, J. Strain Anal. Eng. Des.
20 (1985) 139–150. doi:10.1243/03093247V203139.
[131] J. Mistry, A.G. Gibson, Y.S. Wu, Failure of composite cylinders under
combined external pressure and axial loading, Compos. Struct. 22 (1992)
193–200. doi:http://dx.doi.org/10.1016/0263-8223(92)90055-H.
[132] P.D. Soden, R. Kitching, P.C. Tse, Y. Tsavalas, M.J. Hinton, Influence
of winding angle on the strength and deformation of filament-wound
composite tubes subjected to uniaxial and biaxial loads, Compos. Sci.
Technol. 46 (1993) 363–378. doi:http://dx.doi.org/10.1016/0266-
3538(93)90182-G.
[133] J. Bai, P. Seeleuthner, P. Bompard, Mechanical behaviour of ± 55 °
filament-wound glass-fibre/epoxy-resin tubes: I. Microstructural
analyses, mechanical behaviour and damage mechanisms of composite
tubes under pure tensile loading, pure internal pressure, and combined
loading, Compos. Sci. Technol. 57 (1997) 141–153.
doi:https://doi.org/10.1016/S0266-3538(96)00124-8.
[134] G. Hu, J. Bai, E. Demianouchko, P. Bompard, Mechanical behaviour of
±55° filament-wound glass-fibre/epoxy-resin tubes—III.
Macromechanical model of the macroscopic behaviour of tubular
structures with damage and failure envelope prediction, Compos. Sci.
Technol. 58 (1998) 19–29. doi:http://dx.doi.org/10.1016/S0266-
3538(97)00078-X.
[135] J. Rousseau, D. Perreux, N. Verdière, The influence of winding patterns
on the damage behaviour of filament-wound pipes, Compos. Sci.
Technol. 59 (1999) 1439–1449. doi:http://dx.doi.org/10.1016/S0266-
3538(98)00184-5.
[136] A. Béakou, A. Mohamed, Influence of variable scattering on the
optimum winding angle of cylindrical laminated composites, Compos.
Struct. 53 (2001) 287–293. doi:http://dx.doi.org/10.1016/S0263-
8223(01)00012-5.

178
[137] P. Mertiny, F. Ellyin, A. Hothan, An experimental investigation on the
effect of multi-angle filament winding on the strength of tubular
composite structures, Compos. Sci. Technol. 64 (2004) 1–9.
doi:http://dx.doi.org/10.1016/S0266-3538(03)00198-2.
[138] G. Meijer, F. Ellyin, A failure envelope for ±60° filament wound glass
fibre reinforced epoxy tubulars, Compos. Part A Appl. Sci. Manuf. 39
(2008) 555–564.
doi:http://dx.doi.org/10.1016/j.compositesa.2007.11.002.
[139] A.E. Antoniou, C. Kensche, T.P. Philippidis, Mechanical behavior of
glass/epoxy tubes under combined static loading. Part I: Experimental,
Compos. Sci. Technol. 69 (2009) 2241–2247.
doi:https://doi.org/10.1016/j.compscitech.2009.06.009.
[140] L.A.L. Martins, F.L. Bastian, T.A. Netto, Structural and functional
failure pressure of filament wound composite tubes, Mater. Des. 36
(2012) 779–787. doi:http://dx.doi.org/10.1016/j.matdes.2011.11.029.
[141] L.A.L. Martins, F.L. Bastian, T.A. Netto, The effect of stress ratio on the
fracture morphology of filament wound composite tubes, Mater. Des. 49
(2013) 471–484. doi:http://dx.doi.org/10.1016/j.matdes.2013.01.026.
[142] G.A. Arnold, T.G. Ingram, VI. The phenomena of rupture and flow in
solids, Philos. Trans. R. Soc. London. Ser. A, Contain. Pap. a Math. or
Phys. Character. 221 (1921) 163–198. doi:10.1098/rsta.1921.0006.
[143] R.Q. de Macedo, R.T.L. Ferreira, J.M. Guedes, M.V. Donadon, Intraply
failure criterion for unidirectional fiber reinforced composites by means
of asymptotic homogenization, Compos. Struct. 159 (2017) 335–349.
doi:https://doi.org/10.1016/j.compstruct.2016.08.027.
[144] S.K. Ha, K.K. Jin, Y. Huang, Micro-mechanics of failure (MMF) for
continuous fiber reinforced composites, J. Compos. Mater. 42 (2008)
1873–1895.
[145] P.D. Soden, M.J. Hinton, A.S. Kaddour, Lamina properties, lay-up
configurations and loading conditions for a range of fibre-reinforced
composite laminates, Compos. Sci. Technol. 58 (1998) 1011–1022.
doi:https://doi.org/10.1016/S0266-3538(98)00078-5.
[146] T.J. Vaughan, C.T. McCarthy, Micromechanical modelling of the
transverse damage behaviour in fibre reinforced composites, Compos.
Sci. Technol. 71 (2011) 388–396.
doi:https://doi.org/10.1016/j.compscitech.2010.12.006.
[147] F.P. Van der Meer, Mesolevel modeling of failure in composite
laminates: constitutive, kinematic and algorithmic aspects, Arch.
Comput. Methods Eng. 19 (2012) 381–425.

179
[148] F.P. van der Meer, L.J. Sluys, S.R. Hallett, M.R. Wisnom, Computational
modeling of complex failure mechanisms in laminates, J. Compos.
Mater. 46 (2011) 603–623. doi:10.1177/0021998311410473.
[149] D. Mollenhauer, L. Ward, E. Iarve, S. Putthanarat, K. Hoos, S. Hallett,
X. Li, Simulation of discrete damage in composite Overheight Compact
Tension specimens, Compos. Part A Appl. Sci. Manuf. 43 (2012) 1667–
1679. doi:https://doi.org/10.1016/j.compositesa.2011.10.020.
[150] F. Naya, M. Herráez, C.S. Lopes, C. González, S. Van der Veen, F. Pons,
Computational micromechanics of fiber kinking in unidirectional FRP
under different environmental conditions, Compos. Sci. Technol. 144
(2017) 26–35. doi:https://doi.org/10.1016/j.compscitech.2017.03.014.
[151] W. Toh, L.B. Tan, K.M. Tse, A. Giam, K. Raju, H.P. Lee, V.B.C. Tan,
Material characterization of filament-wound composite pipes, Compos.
Struct. 206 (2018). doi:10.1016/j.compstruct.2018.08.049.
[152] Z.P. Bazant, I.M. Daniel, Z. Li, Size effect and fracture characteristics of
composite laminates, J. Eng. Mater. Technol. ASME)(USA). 118 (1996)
317–324.
[153] M.R. Wisnom, Size effects in the testing of fibre-composite materials,
Compos. Sci. Technol. 59 (1999) 1937–1957.
doi:https://doi.org/10.1016/S0266-3538(99)00053-6.
[154] Z.-M. Huang, L.-M. Xin, In situ strengths of matrix in a composite, Acta
Mech. Sin. 33 (2017) 120–131.
[155] P.D. Soden, A.S. Kaddour, M.J. Hinton, Recommendations for designers
and researchers resulting from the world-wide failure exercise, in: Fail.
Criteria Fibre-Reinforced-Polymer Compos., Elsevier, 2004: pp. 1223–
1251.
[156] A.S. Kaddour, M.J. Hinton, P.A. Smith, S. Li, A comparison between the
predictive capability of matrix cracking, damage and failure criteria for
fibre reinforced composite laminates: Part A of the third world-wide
failure exercise, J. Compos. Mater. 47 (2013) 2749–2779.
http://jcm.sagepub.com/content/47/20-21/2749.abstract.
[157] A.S. Kaddour, M.J. Hinton, P.D. Soden, A comparison of the predictive
capabilities of current failure theories for composite laminates: additional
contributions, Compos. Sci. Technol. 64 (2004) 449–476.
doi:http://dx.doi.org/10.1016/S0266-3538(03)00226-4.
[158] K. Rohwer, Predicting fiber composite damage and failure, J. Compos.
Mater. 49 (2014) 2673–2683. doi:10.1177/0021998314553885.

180
LIST OF PUBLICATIONS

Journal Publications:

Tan, V.B.C., Raju, K and Lee, H. P., Direct FE2 for concurrent multilevel

modelling of heterogeneous structures, Computer Methods in Applied

Mechanics and Engineering (Accepted, 2019).

Toh, W., Tan, L. B., Tse, K. M., Giam, A., Raju, K., Lee, H. P., & Tan, V. B.

C. (2018). Material characterization of filament-wound composite

pipes. Composite Structures, 206, 474-483.

Conference Publications:

Raju, K., Lee, H .P., Tan, V .B.C. (2019). Direct FE2 for concurrent multilevel

modeling of composites structures. In 3rd South-East Asia – Japan Conference

on Composite Materials: SEAJCCM 2019, in Bali, Indonesia, 4th to 6th August,

2019.

Toh, W., Raju, K., Yeo, C. H., Goh, S. H., & Tan, V. B. C (2017). Experimental

and numerical analysis of fibre-reinforced composite pipes subjected to

underground blasts. In 21st International Conference on Composite Materials:

ICCM21, in Xian, China, 20-25 August 2017

Raju, K., Tan, L. B., Tse, K. M., Lee, H. P., & Tan, V. B. C. (2015).

Experimental characterization of GRE composite for failure envelope

validation. In 10th International Conference on Composite Science and

Technology: ICCST/10, in Lisbon, Portugal, 2-4 September 2015.

181

You might also like