Sunteți pe pagina 1din 170

A theory of

Strongly orthotropic continuum mechanics

D.C. Kellermann

A dissertation presented for the degree


of Doctor of Philosophy

June 2008

School of Mechanical and Manufacturing Engineering, J17,


The University of New South Wales, Sydney, NSW 2052, Australia
ABSTRACT
The principal contribution of this dissertation is a theory of Strongly Orthotropic Continuum
Mechanics that is derived entirely from an assertion of geometric strain indeterminacy. Imple-
mentable into the finite element method, it can resolve widespread kinematic misrepresentations
and offer unique and purportedly exact strain-induced energies by removing the assumptions of
strain tensor symmetry. This continuum theory births the proposal of a new class of physical
tensors described as the Intrinsic Field Tensors capable of generalising the response of most
classical mechanical metrics, a number of specialised formulations and the solutions shown to
be kinematically intermediate. A series of numerical examples demonstrate Euclidean objectiv-
ity, material frame-indifference, patch test satisfaction, and agreement between the subsequent
Material Principal Corotation and P–I–C decomposition methods that produce the intermediary
stress/strain fields. The encompassing theory has wide applicability owing to its fundamental
divergence from conventional mechanics, it offers non-trivial outcomes when applied to even
very simple problems and its use of not the Eulerian, Lagrangian but the Intrinsic Frame gener-
ates previously unreported results in strongly orthotropic continua.
Dedicated to you, dad.

[Bill Kellermann, 1942–2006]

… if only you’d have seen me finish.


ACKNOWLEDGEMENTS
Dr. Tomonari Furukawa
From whom I’ve unquestionably learnt the most.

Prof. Don Kelly


Thank you, Don, for being a great mentor to me.

CMR Group and all my colleagues


Especially to Mike and Hin, for the thousand-odd discussions and the thousand-odd cups of cof-
fee. To Amm of course, Ian, Daniel, Ben, Pan, Stephen, Ryan, Alex, Edward, Phil, Mark, Daud,
Baneen, Luke, Zoltan, Garth and whoever else might have listened while I harangued.

My family
To Eva (mum), for the endless support. To my oldest brother Adam, to Ann-Maree and Marissa,
but particularly to my brother Mike who seems astoundingly to have become quite familiar with
the fields of continuum and computational mechanics.

My good friends
To Dr. Tim O’Neill for being a role model of sorts, to Mike Forward for being my best mate for
so many years, and to Richard Urwin for a genuinely motivating—and often humorous—
interest and enthusiasm in my work. Also thanks to Ed Giles and others for their ongoing and
much-appreciated encouragement.

The Cooperative Research Centre for Advanced Composite Structures


The support given by individuals associated with the CRC-ACS is gratefully acknowledged,
including Xiaobo Yu, Damian McGuckin, Rowan Paton, Michael Bannister, Israel Herszberg
and Murray Scott.

iv
ORIGINALITY STATEMENT
‘I hereby declare that this submission is my own work and to the best of my knowledge
it contains no materials previously published or written by another person, or substantial
proportions of material which have been accepted for the award of any other degree or
diploma at UNSW or any other educational institution, except where due acknowl-
edgement is made in the thesis. Any contribution made to the research by others, with
whom I have worked at UNSW or elsewhere, is explicitly acknowledged in the thesis. I
also declare that the intellectual content of this thesis is the product of my own work,
except to the extent that assistance from others in the project's design and conception or
in style, presentation and linguistic expression is acknowledged.’

Signed ………………………………………………….. Date ………………...……

v
COPYRIGHT STATEMENT
‘I hereby grant the University of New South Wales or its agents the right to archive and
to make available my thesis or dissertation in whole or part in the University libraries in
all forms of media, now or here after known, subject to the provisions of the Copyright
Act 1968. I retain all proprietary rights, such as patent rights. I also retain the right to
use in future works (such as articles or books) all or part of this thesis or dissertation.
I also authorise University Microfilms to use the 350 word abstract of my thesis in
Dissertation Abstract International (this is applicable to doctoral theses only). I have
either used no substantial portions of copyright material in my thesis or I have obtained
permission to use copyright material; where permission has not been granted I have ap-
plied/will apply for a partial restriction of the digital copy of my thesis or dissertation.'

Signed ………………………………………………….. Date ………………...……

AUTHENTICITY STATEMENT

‘I certify that the Library deposit digital copy is a direct equivalent of the final officially
approved version of my thesis. No emendation of content has occurred and if there are
any minor variations in formatting, they are the result of the conversion to digital for-
mat.’

Signed ………………………………………………….. Date ………………...…

vi
CONTENTS
ABSTRACT ...................................................................................................................... 2
ACKNOWLEDGEMENTS .............................................................................................iv
ORIGINALITY STATEMENT ........................................................................................ v
COPYRIGHT STATEMENT ..........................................................................................vi
AUTHENTICITY STATEMENT ...................................................................................vi
CONTENTS ................................................................................................................... vii
LIST OF FIGURES........................................................................................................ xii
LIST OF TABLES .........................................................................................................xvi
ABBREVIATIONS...................................................................................................... xvii
Chapter 1 ........................................................................................................................... 1
Introduction ....................................................................................................................... 1
1.1 An introduction to strongly orthotropic continuum mechanics ............................... 1
1.1.1 A brief appraisal of continuum mechanics from its inception........................... 2
1.1.2 Evidence for change .......................................................................................... 4
1.1.3 Why strongly orthotropic continuum mechanics? ............................................. 5
1.2 Objective .................................................................................................................. 7
1.3 Approach .................................................................................................................. 7
1.4 Principal technical contributions ............................................................................. 9
1.5 Publications and presented work ............................................................................. 9
1.6 Organisation ........................................................................................................... 10
1.7 Disambiguation ...................................................................................................... 12
1.7.1 Notational categorisations ............................................................................... 12
1.7.2 Definitions ....................................................................................................... 12
1.7.3 Footnotes ......................................................................................................... 12
Chapter 2 ......................................................................................................................... 13
Literature review ............................................................................................................. 13
vii
viii

2.1 Kinematically driven models for fibre-reinforced materials ................................. 13


2.2 FE-implementable constitutive models for composites ......................................... 15
2.3 Experimental techniques for finite deformation of CFRPs ................................... 17
2.4 Specialised continuum approaches for strong orthotropy...................................... 20
2.5 Some relevant micropolar and Cosserat theory ..................................................... 23
2.6 Finite element techniques for behavioural compensation...................................... 25
2.7 Multi-scale and experimental mapping approaches............................................... 26
2.8 Chapter summary ................................................................................................... 28
Chapter 3 ......................................................................................................................... 29
Strongly orthotropic continuum mechanics .................................................................... 29
3.1 Mechanics of orthotropic materials ....................................................................... 29
3.1.1 The linear tensors of the displacement state .................................................... 29
3.1.2 Strongly orthotropic kinematics and deformation ........................................... 32
3.2 The assertion of geometric strain indeterminacy ................................................... 33
3.3 The orthotropic linear strain tensor ........................................................................ 33
3.4 The material principal rotation tensor .................................................................... 38
3.5 Physical interpretation of the OLST and its components ...................................... 40
3.6 The extended Mohr plot for strain ......................................................................... 41
3.7 The extended orthotropic compliance tensor ......................................................... 42
3.7.1 The extended orthotropic stiffness matrix for plane stress .............................. 44
3.7.2 The extended orthotropic stiffness matrix for plane strain .............................. 45
3.8 Orthotropic von Mises stress invariant .................................................................. 46
3.9 The anisotropic linear strain tensor ........................................................................ 47
3.10 Numerical examples ............................................................................................ 49
3.10.1 Example 2: Satisfying Euclidean objectivity and the principle of material
frame-indifference .................................................................................................... 49
3.11 Chapter summary ................................................................................................. 53
Chapter 4 ......................................................................................................................... 54
SOCM plane shell finite elements................................................................................... 54
4.1 Finite element analysis of plane shells .................................................................. 54
4.1.1 Finite element treatment for a Tri3 element using the ILST ........................... 54
ix

4.1.2 Finite element treatment for a Quad4 element using the ILST ....................... 55
4.2 Premise: persistent material principal orthogonality ............................................. 56
4.3 Finite element treatment for a Tri3 and Quad4 element using the OLST ............. 57
4.4 Material principal corotation ................................................................................. 58
4.4.1 Directional decomposition of multi-axial materials ........................................ 59
4.4.2 Corotational procedure .................................................................................... 60
4.4.3 Equivalence between combined and coincident continua ............................... 61
4.5 Some other considerations ..................................................................................... 62
4.5.1 Fibre pinning versus fibre slippage ................................................................. 62
4.5.2 Coupled drilling moments and micropolarity.................................................. 63
4.6 Numerical examples .............................................................................................. 64
4.6.1 Example 1: Patch test for a linear multi-axial problem ................................... 64
4.6.2 Example 3: Arbitrary element under arbitrary orientation .............................. 66
4.6.3 Example 2: In-plane shear locking/softening of a biaxial material ................. 69
4.7 Chapter summary ................................................................................................... 75
Chapter 5 ......................................................................................................................... 76
The intrinsic field tensors ................................................................................................ 76
5.1 Nonlinear field tensors in mechanics ..................................................................... 76
5.1.1 Class definition: (solely) extrinsic field tensors .............................................. 76
5.1.2 Polar decomposition and the Stretch Tensors ................................................. 77
5.1.3 Polar decomposition: numerical solution procedure ....................................... 78
5.1.4 Biot strain ........................................................................................................ 79
5.1.5 Cauchy–Green and Green–Lagrange strain..................................................... 79
5.1.6 Other extrinsic field tensors............................................................................. 80
5.2 The intrinsic frame ................................................................................................. 81
5.3 The finite intrinsic field tensors for stretch ............................................................ 82
5.3.1 The orthotropic stretch tensors and skewed polar decomposition................... 83
5.3.2 Numerical solution of two-dimensional skew decomposition ........................ 85
5.4 The finite intrinsic field tensors for strain ............................................................. 86
5.4.1 The orthotropic Biot strain .............................................................................. 86
5.4.2 The orthotropic Cauchy–Green tensors ........................................................... 88
x

5.4.3 The orthotropic Green–Lagrange strain .......................................................... 89


5.4.4 The orthotropic Almansi–Euler strain ............................................................. 90
5.4.5 The orthotropic logarithmic Hencky Strain ..................................................... 90
5.4.6 An additional note on micropolarity................................................................ 91
5.5 Finite anisotropic IFTs ........................................................................................... 92
5.6 Generalised decomposition theories ...................................................................... 93
5.6.1 The objective Cholesky decomposition........................................................... 94
5.6.2 Generalised Polar–Intermediate–Cholesky decomposition ............................. 95
5.7 Numerical examples .............................................................................................. 96
5.7.1 Example 1: Verification of the IFTs against corotational result ..................... 97
5.7.2 Example 2: Behavioural comparison between the IFTs and EFTs ................. 98
5.7.3 Example 3: Euclidean objectivity and material-frame indifference .............. 102
5.7.4 Example 4: Full-range nonlinear trellis shear................................................ 104
5.8 Chapter summary ................................................................................................. 106
Chapter 6 ....................................................................................................................... 107
A case study .................................................................................................................. 107
6.1 Mechanics and rheology of CFRPs in forming ................................................... 107
6.1.1 Mechanics of engineering pre-pregs ............................................................. 108
6.1.2 A constitutive/rheological model .................................................................. 109
6.2 Testing in trellis shear .......................................................................................... 111
6.2.1 Picture frame rig ............................................................................................ 111
6.2.2 Multi-axial testing ......................................................................................... 112
6.3 Application of a stress model............................................................................... 113
6.3.1 Picture-frame kinematics ............................................................................... 113
6.3.2 Planar strains and strain-rates ........................................................................ 113
6.3.3 Applying a stress model ................................................................................ 116
6.3.4 Modelling through-thickness strain ............................................................... 118
6.3.5 Calculation of boundary forces ..................................................................... 118
6.4 Parameter identification technique ...................................................................... 119
6.4.1 Pull-rate profiling .......................................................................................... 119
6.4.2 Viscous relaxation ......................................................................................... 121
xi

6.4.3 Cyclic inference ............................................................................................. 122


6.4.4 Rate invariant decomposition ........................................................................ 124
6.5 Chapter summary ................................................................................................. 126
Chapter 7 ....................................................................................................................... 127
Overview and conclusion .............................................................................................. 127
7.1 Overview .............................................................................................................. 127
7.1.1 Classification of work.................................................................................... 127
7.1.2 Summary of equations ................................................................................... 129
7.1.3 Reflections on notation .................................................................................. 130
7.2 Conclusion ........................................................................................................... 132
7.2.1 Concluding statements................................................................................... 132
7.2.2 Continuing investigation ............................................................................... 133
7.2.3 Closing remark .............................................................................................. 134
REFERENCES .............................................................................................................. 135
APPENDICES............................................................................................................... 153
Appendix A IJNME ................................................................................................ 153
Appendix B ACAM ................................................................................................ 153
Appendix C APCOM .............................................................................................. 153
Appendix D ICTAM ............................................................................................... 153
Appendix E USNCCM ........................................................................................... 153
Appendix F Composites Australia.......................................................................... 153
LIST OF FIGURES

Figure 3.1. A continuum body under some deformation and displacement. The motions
F and F* represent different frames of observation. ....................................................... 30
Figure 3.2. a) Trellis kinematics where the connectors can be considered as inextensible
pivoted fibres; b)–c) Material family  exhibits no rigid rotation while family  rotates

through twice the shear tensor strain ............................................................................... 32


Figure 3.3. a)–d) An infinitely orthotropic material shears along the horizontal principal
material axis freely and without rigid body rotation; e) The isotropic material undergoes
rotation equal to the shear tensor strain........................................................................... 32
Figure 3.4. Characteristic strain indicators of the extended Mohr plot ........................... 41
Figure 3.5 Example problem for satisfaction of the principle of MFI, a square element
undergoes a motion F, which is also shown from a secondary observation frame as the
motion F* ........................................................................................................................ 50
Figure 3.6. Strain and stress plots for as observed from the alternate frames.
Consistency to the extended Mohr plot for strain indicates Euclidean objectivity, while
the stress plot demonstrates the frame indifference of the constitutive model. .............. 52
Figure 4.1. A persistently orthogonal material frame compared to a locking Eulerian
frame—identical to the corotational frame in this case. ................................................. 56
Figure 4.2. Example of continuum decomposition of a plain weave composite with an
infinite orthotropic ratio. This is the basis for the material principal corotation of any
multi-axial material. ........................................................................................................ 59
Figure 4.3. Shadow nodes for the material principal corotation for the example of trellis
shear. a) The initial undeformed domain; and, b) In the first iteration, each of the nodes
has two shadow coordinates. ........................................................................................... 61

xii
xiii

Figure 4.4. a) and b) Different patch test meshes for a square plate (dashed) being
deformed by a trellis shear boundary condition. A uniform von Mises plot results for
each mesh for the material family . The contour bar indicates a uniform von Mises

stress in MPa. .................................................................................................................. 64


Figure 4.5. Orthotropic von Mises stress contour for arbitrary rotations of an arbitrary
planar element ................................................................................................................. 67
Figure 4.6. Extended Mohr plot for strain states under various element rotation ........... 69
Figure 4.7. a) Trellis shear element within a picture frame test; b) The single element
mesh used, exaggerated deformation (dashed) and the co-rotated (·)cr initial
configuration of the infinitely orthotropic material  ..................................................... 70

Figure 5.1 a) The reference configuration is undeformed and universally applicable; b)–
d) The conventional deformed frames for current configuration; e) In this example, the
proposed Intrinsic Frame is an intrinsically dependent corotational frame of observation
......................................................................................................................................... 82
Figure 5.2. The deformation condition analysed: trellis motion for a
20 
 
displacement. ................................................................................................................... 99
Figure 5.3. Infinitesimal linear strain and stress for isotropic and orthotropic
formulations. ................................................................................................................... 99
Figure 5.4 Isotropic Biot strain tensor is compared to the proposed OBST for the trellis
deformation as well as the Biot (corotational Cauchy) stress states. ............................ 100
Figure 5.5 Green–Lagrange strain and stress plots ....................................................... 101
Figure 5.6 Translational and rotational motion of the domain under trellis deformation
....................................................................................................................................... 102
Figure 5.7 Euclidean objectivity and principle of material frame-indifference for Biot
strain and stress ............................................................................................................. 103
Figure 5.8 Conventional and proposed Lagrangian measures of strain and stress. ...... 103
Figure 5.9 Increments of trellis shear motion (only 20 shown) ................................... 104
Figure 5.10 Trumpet plots of Mohr's circle for Biot stress, in MPa, against trellis
fraction with a von Mises stress contour in MPa: a) an isotropic analysis, i.e. an
orthotropic ratio of 1, where the conventional and proposed stresses are
indistinguishable from one another; b) a ratio of 10 shows noticeable variation at large
xiv

deformation; c) a ratio of 100 yields a substantial variation of invariant stress; d)


approaching infinitely orthotropic, with a ratio of 1000, the proposed stress tensor
approaches energy-free motion ..................................................................................... 105
Figure 6.1. Deformation modes within a composite pre-form ...................................... 108
Figure 6.2. Mechanical topology of material type 140 ................................................. 109
Figure 6.3. a) Bias extension test, b) Picture frame test ................................................ 112
Figure 6.4. Geometry of fibre angle; fibre strains ( and  in %); and principal system

(in (). ............................................................................................................................ 114

Figure 6.5. Strains and strain-rates based on kinematic predictions from the bi-laminar
orthotropic model .......................................................................................................... 115
Figure 6.6. Deviatoric stress predictions based on application of the power law visco-
plastic model ................................................................................................................. 116
Figure 6.7. Load response of Hexcel F593 resin with Toray T-300/3K fibre............... 116
Figure 6.8. Estimated elastic stress coupled with a simplified temporal relaxation model
....................................................................................................................................... 117
Figure 6.9. Displacement profiles and ramp rate blocks for linearisation of single-axis
response ......................................................................................................................... 119
Figure 6.10. Deviatoric stress variation for constant shear-rate testing schedule ......... 120
Figure 6.11. Strain and strain rate variation with time for constant shear strain rate
testing schedule ............................................................................................................. 120
Figure 6.12. Deviatoric stress variation for constant shear-rate schedule..................... 121
Figure 6.13. Strain and strain rate variation with time for testing of constant fibre-
normal strain rate........................................................................................................... 121
Figure 6.14. Parametric plot showing force response and displacement profile (dashed)
over time........................................................................................................................ 122
Figure 6.15. Precise fit between exponential decay model and recorded relaxation stress
response. ........................................................................................................................ 122
Figure 6.16. Resultant force from first and second cycle plotted with displacement
profile. ........................................................................................................................... 123
Figure 6.17. Previous figure results for constant fibre-normal strain rate plotted against
actuator displacement .................................................................................................... 124
xv

Figure 6.18. Rate-invariant ratios of force contribution ............................................... 125


Figure 6.19. Reaction force components from the kinematic model compared to the
analytically decomposed experimental force contributions. ......................................... 125
Figure 7.1. Classification of some conventional deformation tensors against proposed
material dependent measures. Linear is considered a subset of nonlinear and isotropic a
subset of orthotropic. ..................................................................................................... 127
Figure 7.2. An illustrative summary of the contribution: strongly orthotropic continuum
mechanics is proposed to suffice the spectrum of kinematically exact finite deformation
analysis. ......................................................................................................................... 128
Figure 7.3. As they have been shown to generalise the response of the conventional
classical tensors, and on acceptance of Figure 7.2, the IFTs are capable of being exact
for all classes of problem shown. .................................................................................. 128
LIST OF TABLES

Table I. Strain categories for planar modes: columns indicate modal change, rows
indicate sign change ........................................................................................................ 40
Table II. Strain and stress values numerating the satisfaction of the principle of MFI .. 51
Table III. Material properties for dual coincident elements (horizontal and vertical) and
for equivalent biaxial elements (combined). Moduli are measured in GPa. ................... 65
Table IV. Results for spatially decomposed solution and conventional result. .............. 65
Table V. Stress and strain values between an ILST and four rotations of the OLST
solution ............................................................................................................................ 67
Table VI. Comparison of analytical and continuum solutions for isotropic picture frame
shear: only the ILST and OLST are kinematically exact, resulting in correct stress
values............................................................................................................................... 71
Table VII. Comparison of analytical and continuum solutions for an infinitely
orthotropic material undergoing energy-free picture frame shear: only the analytical
solution and the OLST are kinematically exact .............................................................. 72
Table VIII. Comparison of analytical and continuum solutions for strongly orthotropic
picture frame shear: the OLST provides the only intermediary solution and is
kinematically exact.......................................................................................................... 73
Table IX. Decomposition categorisation and the Polar–Intermediate–Cholesky
generalisation. ................................................................................................................. 96
Table X. Comparison of conventional linear and Biot strain and stress for conventional
extrinsic and proposed intrinsic formulations against previous results. ......................... 98

xvi
ABBREVIATIONS

The following abbreviations are used throughout the document:

ALST Anisotropic linear strain tensor

CFRP Continuous fibre reinforced plastic (carbon fibre reinforced plastic)

DDF Double diaphragm forming

FEA Finite element analysis

FEM Finite element method

IFRM Ideal fibre reinforced model

IFT Intrinsic field tensor

ILST Isotropic linear strain tensor

MFI Material frame-indifference

OBST Orthotropic Biot strain tensor

OGST Orthotropic Green–Lagrange strain tensor

OLST Orthotropic linear strain tensor

P–I–C Polar–intermediate–Cholesky (decomposition)

PK2 Second Piola–Kirchhoff (stress tensor)

RTP Room temperature and pressure

SOCM Strongly orthotropic continuum mechanics

TS Thermoset (plastic)

xvii
Chapter 1
Introduction
Mechanics, one of the oldest of the natural philosophies, describes both the most notice-
able and the most un-noticeable of physical phenomena. With the ever-increasing abil-
ity we attain from computing, the continuum descriptions of mechanics—lending them-
selves so well to such devices—become only more relevant. With that, so too increases
our insight and discovery into not only the media being described, but into some of the
classical theories that never really saw the light of day.
In this introductory chapter, a problem is defined from an unassuming standpoint.
We arrive there through a brief description of how and why the continuum mechanics
theories got to where they are, and then appeal to evidence and logic to assert that there
may be something missing. Whereas it is left to Chapter 2 to offer up the comprehen-
sive literature to support such a claim, this introductory approach is at least enough to
allow us to define a distinct objective of study and to outline how it is to be tackled.

1.1 An introduction to strongly orthotropic continuum mechanics

While it may at the outset seem unnecessary, we should foremost explore in the briefest
manner a chain of theoretical developments that has brought orthotropic continuum me-
chanics to where it now stands—the necessity of which will become apparent immedi-
ately afterwards. As some of these texts are not always easily found, certain parts and
notations in this section are reproduced for the benefit of the reader, and those notations
should not be kept in mind thereafter.
1
Chapter 1 Introduction 2

1.1.1 A brief appraisal of continuum mechanics from its inception


Inspired largely by his contemporaries, Newton, Bernoulli, Euler and others, it was the
scientific treatise of Lagrange [1] that set in place the analytical mechanics from which
continuum mechanics would grow. It was not until Cauchy [2-4] that the objective
physical descriptions of material distortion were written, and in his works the now fa-
mous linear strain equations were presented. What he referred to as material points are
the infinitesimal elements that define a continuum; the linearity came from his negation
of infinitely small terms of the second order and higher; and, his isotropic system was
said to be such that the coordinate axes could be specified at any orientation—what we
call objectivity or frame indifference. To achieve this objectivity, both observational
and under small rotation, he wrote

Revenons au cas où les axes coordonnés sont dirigés suivant des droites quelconques. Al-
ors, si l'on fait, pour abréger †,
w[ wK w] ½
! , " , # , °°
wx wy wz
¾ (1.1)
wK w] w] w[ w[ wK °
$  , %  , &  ,
wz wy wx wz wy wx ° ¿

where !–# are the normal strains and $–& are the shear (see [2]), and noting particu-

larly the simple averaging of the perpendicular terms of the deformation. Where his
formulations predate the notation of tensor calculus [5], what is now known as
Cauchy’s infinitesimal strain tensor H could be equivalently written in tensor form as

ª ! & %º
H « # " $» . (1.2)
« »
«¬ sym " # »¼

In this form, the tensor symmetry is quite obvious, though it was arrived at by Cauchy
through his supposition of constant uniform coefficients.
Not long after, Green [6] extended Cauchy’s equations to include terms of the sec-
ond order, his I2, and with close reference to the Mécanique Analytique his equations


Translating to: Let us return to the case where the coordinate axes are directed according to unspecified lines. Then, in order to
shorten this,…
Chapter 1 Introduction 3

came to be known as the Green–Lagrange strain. While giving a simple proof of what
we now refer to as major symmetry of the elasticity tensor, he suggests

If therefore the medium is merely symmetrical with respect to each of the three planes,
we see that I2 must remain unaltered … [and] in this way the 21 coefficients are reduced
to 9 … Probably the function just obtained may belong to those crystals which have three
axes of elasticity at right angles to each other.

and in that sense he also cites the generality of his equations to orthotropic (though, not
anisotropic) media, and the specification of a constitutive model in a frame aligned with
the material principal orientation. It is in reference to 21 and nine values that we see ex-
tension of Cauchy’s earlier supposition, in that 21 pertains to the number of terms in the
upper triangle of a six-by-six matrix (his number of equations). Were it said that the
state of a distorted material is described completely by nine equations (the number in
our shown tensor) rather than six, his 21 terms would become 45 and would reduce to
12 in place of nine. The validity of such is affirmed by evidence that Green made spe-
cific note to orthotropy, and that his second-order equations are only relevant over
Cauchy’s under a state of finite deformation.
The method of polar decomposition came later by Autonne [7]—first demonstrated
in the form we use—where he makes the lemma Toute matrice est le produit d’une uni-
taire par une hermitienne †, and subsequently demonstrates the general possibility of the
spectral decomposition A = uh. In Truesdell and Toupin’s [8] mechanics, this becomes
F = RU, meaning that any state of deformation F can be decomposed into a symmetric
right stretch component U and a unitary rotation component R. In that manner, the
stretch is symmetric and is consistent with the physical meaning of Cauchy’s measure.
It was later that Biot [9] suggested the benefits of using a strain tensor of U I, reflecting
precisely the arotational form of Cauchy’s equations, and similarly a lack of need for
any more terms than Green purports.
It is now easily seen that all these forms can be linked together, where for example
the Green–Lagrange strain can be written in terms of the right stretch, and the Cauchy
strain can be attained by negating the second-order terms of that. It is then that we con-
sider how well the continuum mechanics is described, but also how deeply it changes


Translating to: Any matrix is the product of a unitary and a Hermitian matrix
Chapter 1 Introduction 4

under a difference of conditions. It is perhaps that the mechanics have been developed
so well, owing in large part to a brilliant few, that our investigation must begin at such
an early yet critical point in this otherwise pristine domain.

1.1.2 Evidence for change


There is no doubt that a great deal of effort into investigating a great many possibilities
has been made, and we consider many of these in the literature review to come. So, it is
proper that, positive evidence to the suggestion of change or oversight should be sought.
To this end, we take particular note of comments made by Cauchy in a mémoire
[10] published some 23 years after his most regarded (and previously noted) explora-
tions into solid mechanics, whereby he discusses the physical consistency to his notion
of constancy of linear strain coefficients. We carefully consider here that his use of the
term isotropic appears to refer to objective observability rather than to material proper-
ties being equivalent in all directions:

“J’ai recherché ce que deviennent les équations des mouvements infiniment petits d’un ou
même de deux systèmes homogènes de points matériels, quand elles acquièrent la pro-
priété de ne pouvoir être altérées, tandis que l’on fait tourner les axes coordonnés autour
de l’origine, c’est-à-dire, en d’autres termes, quand les systèmes donnés deviennent iso-
tropes. Mais, dans cette recherche, les coefficients que renfermaient les équations li-
néaires données étaient supposés réduits à des quantités constantes; et, comme j’en ai fait
la remarque, cette supposition n’est pas toujours conforme à la réalité.”

In this, after suggesting that he has developed an objective set of equations to describe
the infinitesimal motion and distortion of an element of material, he essentially makes
the observation: But, in this research, the coefficients which contained the linear equa-
tions given were presumed to reduce to constant quantities; and, as I made the remark
of it, this assumption is not always in conformity with reality.
As Cauchy, in that, readily points out that his original formulations were based on a
supposition, the chain of theoretical developments given in Subsection 1.1.1 leads us to
consider that the resultant strain tensor is not always symmetric, even under his well
defined conditions of a material point being elastic, static, uniform and free from body
forces and body moments (read: not micropolar). As was shown before, it leads unto the
definition of the Biot strain, and with that in mind we think whether it might not affect
Chapter 1 Introduction 5

the known incapacity of Biot’s measure to accurately deal with orthotropic conditions:
Xiao et al. [11], having just mentioned this particular inability of Eulerian-type tensors
(Biot’s), state that

“Owing to the above fact, the Lagrangean [sic] type formulation has been popular and
seems preferable in constitutive formulation with reference to material symmetry and ma-
terial objectivity. Indeed, to establish a complete Eulerian type constitutive formulation in
terms of Eulerian quantities, one must clearly know how to treat the material objectivity
and material symmetry. It seems that generally no satisfactory treatment is available, ex-
cept for the case of isotropy which has no preferred directions.”

Neff et al. [12] point out in their work on Biot strains (for isotropic hyper-elasticity) that
they have shown that symmetric Cauchy stresses do not imply symmetric Biot strains,
and that this is contrary to claims in the literature that are valid only in the linear iso-
tropic case. In this, and most poignantly, their conclusion states

“It remains to find a sufficiently large class of isotropic free energies such that moment
equilibrium in the relaxed formulation implies automatically the symmetry of the relaxed
Biot stretch U : This question seems to be a formidable challenge.”

such that this combined seems to offer the evidence we sought for change. All things
considered, it is not presumed that any of these questions are to be answered, only that
there are enough of them and that they have enough depth to warrant the belief that
there is an area worthy of investigation.

1.1.3 Why strongly orthotropic continuum mechanics?


A material attains the description of being strongly orthotropic by possessing rigidity
high in proportion between any two of its orthogonal and privileged material directions.
In the material science, it is invariably the taking away and giving to another that offers
such esteemed structural properties. Exhibiting this directional behaviour and having
high stiffness, high strength, and low density [13], it is materials like Continuous Fibre
Reinforced Plastics (CFRPs) that provide us more capabilities for tailored design and
high performance than would regular isotropic materials [14]. Though, it is those same
characteristics that prove so troublesome in the continuum descriptions in that their in-
ternal behaviour is not always compatible.
Chapter 1 Introduction 6

These and others are the materials that have been referred to as being strongly
orthotropic and strongly anisotropic, which is a term that is adopted herein as an adjec-
tive to describe this particular kind of behaviour. It is the increasing use of these high
performance materials in transport and other such applications that has driven a major
research effort in making their use more economic [15], with attention paid to auto-
mated manufacturing that promises to reduce the large cost premium of fabrication [16,
17]. In pursuit of this, various procedures have been implemented such as resin transfer
moulding [18], thermoforming [19] and diaphragm forming [20], and so in their design
the want for effective simulation and analysis has necessitated accurate modelling and
characterisation [21].
It is with these strongly orthotropic materials in mind that we consider the various
Eulerian and Lagrangian strain tensors used to describe their deformation—tensors like
the infinitesimal Cauchy, Green–Lagrange and Biot strain. As they are necessarily
symmetric as part of their need to be insensitive to small rigid rotation [22], it is reason-
able to suggest that this symmetry relates to an unsuitable isotropic kinematic. It is
without this necessary symmetry that there exist an infinite number of indeterminate
compatible modes for any particular state of distortion, made possible because it comes
in combination with a complementary geometric rigid rotation component [8]. Con-
versely, there have been many other approaches that have successfully modelled com-
posite orthotropic media as having infinitely orthotropic kinematic response—
something which will be demonstrated further in the literature review. So, in considera-
tion of such approaches effectively modelling infinitely orthotropic behaviour, and of
the conventional treatment of orthotropic continua based on classical isotopic tensors, it
is observed that there must be an intermediary formulation. Moreover, a need is asserted
for a subsequent generalised and kinematically exact continuum basis describing all
these intermediary and inclusive classes of isotropic/orthotropic material. This is what
we shall refer to as Strongly Orthotropic Continuum Mechanics (SOCM).
Chapter 1 Introduction 7

1.2 Objective

The objective of this thesis is to establish a comprehensive continuum mechanics theory


that is kinematically exact for the deformation of strongly orthotropic continua, and
which also maintains generality over the isotropic condition.

1.3 Approach

In order to achieve the objective of this thesis, three essential approaches are introduced
to develop the contribution. Relying each on a particular successive conjecture, they are
outlined in detail in the second section of each of the Chapters 3, 4 and 5, and are re-
spectively referred to as:

1. An assertion of geometric strain indeterminacy

2. A premise that requires persistent material principal orthogonality

3. The introduction of the intrinsic frame, with respect to finite deformation

The assertion of geometric strain indeterminacy suggests that the classical tensors
have yet to be presented in a purely rational manner, and have rather relied for some
time on an assumption that services a mathematical requirement. This assumption, ten-
sor symmetry, is a basis for rotational invariance—essentially allowing first or higher-
order indifference to rotational rigid body motion. The infinitesimal strain tensor, and
later the finite stretch tensors, is separated into an indeterminate geometric strain and an
assumption of tensor symmetry that is thence discarded. By use of nothing more than
equilibrium, it is then demonstrated that we can re-derive these classical metrics for the
isotropic condition on one less assumption—an approach that benefits from the law of
parsimony † . More interesting, however, is that no initial requirement of isotropy is
served, and so the resulting parent form presents an opportunity to not only discard the
assumption of tensor symmetry, but also to discard its need. Having done so, this new
orthotropic strain tensor reflects a strong analogy to material motion; that is, what is
propounded to be the true kinematic motion of non-isotropic continua.


Entia non sunt multiplicanda praeter necessitatem
Chapter 1 Introduction 8

In the second chapter, the premise of persistent material principal orthogonality is


introduced. Given infinitesimal displacements yield no variation between the current
and reference frames, it follows that this premise is concerned with finite deformations.
Despite this, and as an intermediary step, the content presented continues usage of the
linear metrics, but does so in the sense of small strain, large rotation. Dealing with a
finite element treatment of the previously developed tensors, the persistent material
principal orthogonality is serviced by an adapted form of the well-known method of
corotation, where now the rotational measure depends on both the geometry and the
mechanical properties of the continuum. By considering a material undergoing a sig-
nificant shear deformation such that perpendicular fibres are no longer orthogonal, a
method of material principal corotation allows the maintenance of the premise of or-
thogonality in the constitutive model by convecting each continuum according to its
own material-dependent orientation.
In the third and final contributory chapter the intrinsic frame is introduced, which
encompasses the progress of the previous two via a single theoretical conduit. This sug-
gests that a consistent observational frame—or put differently, one that recognises the
true orientation of some continuum based on any arbitrary initiation—will according to
the preceding developments be intrinsically dependent. So, to the same effect of the in-
determinate linear strain tensor, the finitely deformed frame must require some intrinsic
knowledge of the continuum being observed. Now armed with a Euclid-objective equi-
librium condition and a notion of frame orthogonality and convection, the most essential
finite deformation tensors are revised. These, the stretch tensors, join the orthotropic
linear strain tensor in the proposed class of Intrinsic Field Tensors (IFTs). From there it
becomes trivial to reformulate a number of classical deformation metrics, including the
Green–Lagrange strain, but most importantly the strain tensor of Biot. With a new gen-
eralised form to his strain measure that no longer suffers from numerical locking under
orthotropic conditions, a supposition is made that it may now be preferential to the La-
grangian tensor of Green—the one conflicting with the premise of persistent principal-
material orthogonality. Moreover, it is suggested that a previously unreported solution,
one that is kinematically and energetically exact, may have been found.
Chapter 1 Introduction 9

1.4 Principal technical contributions

The respective principal technical contributions of this thesis are as follows:

Chapter 3. The basis of strongly orthotropic continuum mechanics for linear problems

o The derivation of the orthotropic linear strain tensor

o The subsequent material principal rotation tensor

o The introduction of the metric described as aspectual strain

o Application to the extended Mohr plot for strain

o The extended orthotropic compliance tensor and stiffness matrices

o The resultant anisotropic formulations based on SOCM

Chapter 4. A plane shell finite element implementation of SOCM

o Reformulation of the element system mapping matrices

o Multiple coincident continua for multi-axial materials

o The basis and procedure of material principal corotation

Chapter 5. The broad concept of intrinsic field tensors and nonlinear extension to SOCM

o Skewed polar decomposition and a numerical solution procedure

o The intrinsically dependent finite rotation and stretch tensors

o The orthotropic form of the Biot strain

o The orthotropic Green–Lagrange tensor

o Intrinsic formulations for, among others, the Cauchy–Green tensors, Almansi–

Euler tensor and logarithmic Hencky strain

o The basis for the finite anisotropic IFTs

o The generalised formulation of Polar–Intermediary–Cholesky decomposition

1.5 Publications and presented work

International peer-reviewed journal

o D.C. Kellermann, T. Furukawa and D.W. Kelly, "Strongly orthotropic continuum me-

chanics and finite element treatment," International Journal of Numerical Methods in


Chapter 1 Introduction 10

Engineering (IJNME), (29 pages, published online and in print), DOI: 10.1002/

nme.2379.

o D.C. Kellermann, T. Furukawa and D.W. Kelly, "The intrinsic field tensors of mechan-

ics," (in preparation)

International conferences

o D.C. Kellermann, T. Furukawa, J.W. Pan, "A Continuum Mechanics Solution for In-

plane Shear Locking in Plate and Shell Elements," Ninth U.S. National Congress on

Computational Mechanics (USNCCM IX), San Francisco, July 22–26, 2007 (Presented),

p. 2, 2007.

o D.C. Kellermann, T. Furukawa, and D.W. Kelly, “Finite Element Implementation and

Sub-laminar Decomposition for Strongly Orthotropic Continuum Mechanics”. The

Third Asian-Pacific Congress on Computational Mechanics (APCOM'07), Kyoto, Ja-

pan, December 3–6, 2007.

o D.C. Kellermann, T. Furukawa, D.W. Kelly, "A Theory of Strongly Orthotropic Con-

tinuum Mechanics," 5th Australasian Congress on Applied Mechanics (ACAM07),

Brisbane, Queensland, Australia, December 10–12 2007.

o D.C. Kellermann, T. Furukawa, D.W. Kelly, "Intrinsic field tensors for strongly

orthotropic continua," 22nd International congress on Theoretical and Applied Me-

chanics (ICTAM2008), Adelaide, Australia, August 24–30 2008.

Local conference

o D.C. Kellermann, T. Furukawa, D.W. Kelly, "Intrinsic field tensors for strongly

orthotropic continua," Composites Australia Conference & Exhibition, Melbourne, Aus-

tralia, March 13–14 2008.

1.6 Organisation

Having identified our area of investigation into classical continuum mechanics, the
scope of Chapter 2 is to survey the subsequent developments in the field that have a
strong relevance to the topic or that have presented some method mitigating problems
Chapter 1 Introduction 11

that may have arisen as a result. This serves both to strengthen the case that a problem
truly exists, and to demonstrate that none of the contemporary approaches are suffi-
ciently comprehensive such that they might adequately serve the objective of this thesis.
In the three chapters that follow, the theoretical contribution of this thesis is pre-
sented so as to address the problem defined in the introductory chapter at its branching
point, or most fundamental level. They begin with a section on fundamental theory, fol-
lowed by the detail of their respective approach, and then from the third section on-
wards the contributory elements are presented before reaching the associated numerical
examples. Commencing at Chapter 3, some conventional theory and interpretation on
orthotropic continuum mechanics, kinematics and stress tensors is given before the con-
cepts and techniques relating to the proposed linear continuum mechanics theory are
presented thereafter. Where that chapter is limited to linear theory, an exploration into
large rotations with linear strain is made in Chapter 4 by applying linear theory to the
finite element method and interpreting a new type of corotation. Through application of
this, the numerical examples demonstrate the validity of the method while also enumer-
ating a benchmark solution to which the nonlinear theory can refer. Chapter 5 can
thence present a decisive case for finite deformation’s use of the proposed intrinsic field
tensors.
Chapter 6, the second last, is a case study for the thematic problem of trellis shear.
Despite the sizeable gap between fully implementable techniques for the proposed the-
ory and the large deformation visco-plastic response of the experiment performed, the
results are able to be decomposed and analysed sufficiently enough to allow a rallying
case for strongly orthotropic theory. The thesis and impetus of this dissertation is ulti-
mately brought to conclusion in Chapter 7 with an over-viewing assessment, summary,
and discussion of future work in the field. A conclusive statement is given regarding
implications of the work presented and finally a closing remark is made. Note that ref-
erences, in order of appearance, and an appendix containing published or presented
works reside at the end of this document.
Chapter 1 Introduction 12

1.7 Disambiguation

1.7.1 Notational categorisations


Isotropic, orthotropic and anisotropic linear descriptions are defined according to the
denotation e, œ and æ in the upper-right corner of some symbol. In nonlinear cases,
these are respectively replaced by E, Œ and Æ. This is demonstrated as follows:

^
power
coordinate system` denotation
Eindice or denotation where
when / over

coordinate system:  ,  ,  , %, %, &


denotation: e, œ, æ, E, Œ, Æ, vm, c-r, s, pe
indice: 1, 2,3, x, y, z, [ ,K
power: 0.5, 2, 3, 4...
when/over: 0, t , current, 1%...

noting in particular that braces ^`


˜ indicate something to belong to its rightward charac-

ter and the raising of some symbol to a power is indicated by the use of parentheses
where otherwise it may be ambiguous as compared to a denotation.

1.7.2 Definitions
Whenever a term is written in Capitalised Italics it means a new label is being intro-
duced to represent a novel concept, postulate or formulation; whereas if a term is only
Capitalised it is done so to highlight a conventional or widely accepted terminology.

1.7.3 Footnotes
Footnotes appear throughout the document either to give a translation or so as to infor-
mally point out what has not been done in this dissertation. In the latter case they sug-
gest or infer some area of future study.
Chapter 2
Literature review
In this chapter an attempt is made to cluster the theories and methods of analysis for
strongly orthotropic continua according to the predominant bases of approach. Reflect-
ing also the chronology, the kinematically-driven analytical approaches to fibre-
reinforced continua are the first subject of this literature review. Choosing instead a
pure continuum description, the next category focuses primarily on the constitutive
modelling of composites—efforts that tend towards data-driven behaviour. A brief di-
gression then highlights the associated difficulties by reviewing a number of published
experimental techniques for parameter identification. By realisation of the over-
parameterised complicatedness of these constitutive descriptions, the next section sur-
veys some of the more mechanistic specialised continuum methods. Thereafter, we take
an excursion into micro-polar theory. Such theoretical specialisation is then shown to be
greatly reduced in a small category of non-continuum or subjective finite element tech-
niques; which are subsequently improved upon in the multi-scale and other subjectively
mapped methods reviewed. Finally we surmise that no work has, as yet, provided an-
swer to our objective.

2.1 Kinematically driven models for fibre-reinforced materials

The most prevalent of the strongly orthotropic (or anisotropic) class of materials are the
fibre-reinforced composites given that they provide a practical mechanism for attaining
highly directional properties. Equivalent behaviour, yet not always so strong, occurs in
13
Chapter 2 Literature review 14

biological tissue and bone, wood, rock strata and more, though the precision of the con-
stitution of the fibre-reinforced materials lends itself best to kinematic descriptions. This
is referred as being a kinematically driven approach, while a later section categorises
specialised continuum methods. While many of the methods here too are sound contin-
uum descriptions, we note that this category of works is based on ideas like kinematic
admissibility and constraint rather than constitutive response.
The most notable works on this approach were provided by Spencer—particularly in
his expansive monographs—[23, 24], whose efforts initially sought to formulate the re-
sponse of Ideal Fibre-Reinforced Models (IFRMs); and similarly those inclusive and
separate by Rogers [25]. This idealisation assumes fibres to be inextensible and usually
pin-jointed so as to create precisely a trellis-like structure that is kinematically determi-
nate, although it has been shown that even such inextensible assumptions need more in-
depth consideration [26]. In the present work, this type of treatment is denoted as being
kinematically infinitely orthotropic because it can exactly describe the motion of rigid
pivoting structural elements that provide no normal resistance. While all those proper-
ties are not persistent, the kinematic analogy remains. These equations were later util-
ised in the development of flow rules for rheological modelling [27, 28] to directly for-
mulate constitutive equations [29, 30], and later also extended to encompass evolvabil-
ity [31]. While not always directly attributed, the basis of these kinematic equations has
been widely used to characterise the behaviour, geometry and permeability of dry
woven [32-35] or braided fabrics [36], or for the purpose of thermo-plastics [37-40]
and/or pre-impregnated composites [41-43]. Also, the kinematic constraints of the
IFRM model have offered a number of additional analytical challenges [44, 45].
Of those, the approach has been particularly suitable as dry cloths, outside of small
discontinuous frictional contributions, are accurately portrayed as being infinitely
orthotropic. Those results have been used with a high degree of success in a number of
fabric forming simulations, specifically the kinematically driven type, and predomi-
nantly by those same authors interested in characterisation [37, 42, 43]. It follows that
the addition of rheological energy and rate models can be added to improve their predic-
tive ability for pre-impregnated composites as is done by Clifford, Harrison and Long
[46-48]. Similarly, there has been work towards commercially available kinematic fibre
Chapter 2 Literature review 15

forming simulation tools [49, 50], though those tools have no capacity to predict spuri-
ous wrinkling and other such mechanisms. Nonetheless, kinematically driven methods
have been highly successful albeit very specific.
Having appraised the kinematically driven approach for its successful offerings to
infinitely and near-infinitely orthotropic media, it is surmised that any intermediary
formulations would come with considerable difficulty given the strong dependence of
geometric mechanisms, and furthermore that this approach is limited in dealing with
arbitrary geometries and as a generalised description.

2.2 FE-implementable constitutive models for composites

With the increasing utilisation of finite element codes, the focus on FE-implementable
methods for composite analysis naturally followed. The advantages were, largely, due
to the high adaptability of FE particularly in cases of complex geometry, though such
interest was initially primarily directed towards structural analysis.
Where this study principally considers problems arising from large rotation and/or
finite deformation, these same constitutive approaches have expanded from structural
analysis in recent times to encounter dominantly deformation-based analysis over pre-
viously more popular kinematic approaches [51, 52]. Though despite the similarities,
the large deformation and often rheological considerations has meant the characterisa-
tion and parameter identification problems have differed significantly from those of
typical structural composites laminates [53-55]. Additionally, in being dominated by
minimum energy as determined by the constitutive model, here we regard these as data-
driven as opposed to being driven by kinematic admissibility as in the previous section;
and such approaches tend to be readily implementable as constitutive models for FEA
codes [56]. The groundbreaking efforts in this area are mostly attributed to Rogers’ [23,
57] early work, while Bhattacharyya [58] gives an authoritative account of the devel-
opments towards practical composite sheet forming analysis.
Based on finite continuum theory McGuiness and Ó Brádaigh [59, 60] proposed a
number of rheological models based on identification against the picture frame shear
experiment. Following these investigations, the predominance of the shear mode in de-
Chapter 2 Literature review 16

termining deformation energy has encouraged a focus on shear characterisation for


power-law viscosity models, although Youssef points out that the analysis scale is an
important consideration [61]. These methods have been predominantly experimentally
driven despite the fact that Morozov [62] showed a sound mechanics approach to pre-
dicting these properties.
Also, there have been achievements towards the constitutive modelling of plain dry
fabrics [63], particularly in the animation industry where efforts have been devoted to
cloth modelling—a topic covered well by Eischen and Bigliani [64]. Noteworthy also is
research utilising particle methods [65], which are well suited to that field due to speed
and a need for real-time modelling, and additionally work by Do and Herszberg [66]
who, by applying the kinematic conditions to a finite difference analysis, were able to
successfully model the draping of pre-preg fabrics—albeit with limitations on geometry
and extensibility conditions.
To a similar goal, a number of studies have focused on the development of models
that principally represent the trellising-type behaviour of woven materials, where the
initial energy contribution comes primarily from in-plane shear [41]. On that, Lin et al.
[67] were able to effectively model thermally varying viscosity by using a
rate/temperature-dependent hybrid FE model. These studies have generally focussed on
developing constitutive models for particular forming processes, including stamping of
thermoplastic parts by Lussier et al. [68-72], the slower process of press forming [73],
and double diaphragm forming [51, 74]. From this, various investigations into sensitiv-
ity [75-77] have demonstrated the importance of careful and consistent determination of
parameters. Of the many groups with attention directed towards characterisation for
process modelling, a large number participated in an extensive benchmarking effort [78]
in order to determine satisfactory cross-platform normalisation of results [79].
Critically, it is the results of Clifford, Long and others [80, 81] that depict the prob-
lems growing out of the simplified strain mechanisms regardless of the quality of the
viscous model. To a large extent, the work of Yu et al. has resulted in a practical solu-
tion to this, which we will consider with its caveats in Section 2.7.
In most cases, these constitutive models have to deal with the large change in privi-
leged material directions, what we refer to as the material principal directions. As large
Chapter 2 Literature review 17

shear results in large change of angle between initially orthogonal fibres, an orthotropic
constitutive model must refer stresses in the reference configuration to strains the refer-
ence configuration so as to avoid serious numerical locking. The simplest method to
achieve this is to utilise the Green–Lagrange strain (or similar), so as to map current
configuration geometric distortions into the context of the orthogonal constitutive model.
An alternative has been offered by Xue, Yu and Peng et al. [82-86], who used a non-
orthogonal constitutive model and respective transformation to address the issue of rela-
tively rotating material principal directions. They later extended the model to include
compaction components [87], which provides an advantage over the Green–Lagrange
strain that tends not to indicate any fibre-normal components of strain (total Lagrange
method).
Despite significant successes, this fibre-direction based approach—being kinemati-
cally improved—does not encompass the effect of a matrix component comparable in
stiffness to the fibres, i.e. it continues to model infinitely-orthotropic kinematics. In
overview, all the constitutive models suffer from the major drawback of mapping
against conventional strain measures. So, as these measures are shown to be misrepre-
sentative in other sections here, the encompassing approach of constitutive capture of
kinematic phenomena is not sufficient for the sought objective.

2.3 Experimental techniques for finite deformation of CFRPs

In this digression we examine a number of the issues and proposed solutions in associa-
tion with experimental techniques necessary for the characterisation and parameter
identification of continuous fibre reinforced materials. These experimental results are
critical in driving the mapped approaches, but particular difficulties arise in the correla-
tion and development of the many constitutive approaches just reported.
The majority of such research has been directed towards forming processes for
complex geometries that are more conducive to automation. To this end, substantial
progress has been made towards the characterisation of composites antecedent to their
design-state from both a standpoint of mathematical generality and with respect to ac-
tual commercially available materials. These utilise dry fabric pre-forming for liquid
Chapter 2 Literature review 18

moulding techniques or methods using thermoplastic impregnated composites [88].


However, pre-preg thermo-set composites ultimately offer better control of fibre/matrix
ratios [89] and higher performance under temperate conditions [90] typical to aerospace
applications. Various techniques for forming of thermo-set pre-pregs have shown prom-
ising results [91] such as Double Diaphragm Forming (DDF), although simulation re-
quires accurate material characterisation specific to their own rheology. Significant dif-
ferences in forming response exist between the three basic media: Dry fabric in pre-
forming exhibits much lower shearing resistance due to the absence of a viscous matrix
[41, 92-94], and similarly does not require any rheological consideration in its model-
ling. However, this also results in some effects becoming more significant such as intra-
laminar yarn slippage [32, 50, 95, 96] that can otherwise be reasonably negated; Ther-
moplastic impregnated fabrics require high temperatures to reach forming viscosities, so
constitutive models often must consider thermal dependency. Also, the different funda-
mental chemistry of thermoplastics means the characteristic response will differ from
that of a thermosetting polymer [58]. Hence the task of modelling thermo-set impreg-
nated fabrics escapes certain complexities of its associates, but also presents its own
unique challenges.
We note now that the most common general model is the IFRM, whereby or-
thogonally bi-directional textile reinforcement is considered as an axially inextensible
pin-jointed trellis structure embedded within an incompressible viscous fluid [57, 97,
98]. Based on the subsequent constraints of the structure, the kinematics become deter-
minate for various characterisation experiments [57] and allow for explicit calculation
of strain and strain-rate vectors. Two such experiments are generally used: the picture-
frame test [99], and the bias extension test [100]. Whilst each test applies the same trel-
lis shear deformation to the material, they have respective advantages and disadvantages
that make selection a matter of analysis needs [46]. A number of variations [29, 60, 69,
73] of the IFRF platform have been applied to thermoplastics and shown to be capable
of modelling such materials [59], as well as various approaches for application into
FEA [47, 48, 101, 102]. In implementation to certain analysis codes, application of cri-
teria such as ‘fibre locking angles’ demonstrates problems in the ability of the model to
accurately represent the full span of typical forming strains. In applying these idealised
Chapter 2 Literature review 19

models to thermo-set composites, fairly poor comparisons have been observed when
compared with experimental results [91]. In achieving satisfactory simulation results in
prediction of wrinkling deformities (compared with real formed components) an ap-
proach of function fitting was used [91]. This is indicative of a failure of the idealised
rheological models to accurately represent the stress response of fibre-reinforced ther-
mosetting polymers. Similarly, such an approach tends to breech the coordinate system
independence required of true tensors by showing a necessity towards material oriented
reference axes [103].
The geometric difficulties resulting from complex components usually involves
relatively tight double curvatures, and thus pose the much tackled problem of draping
simulation [43, 50]. In order for woven or stitched fabrics to conform to such curvatures,
high levels of shear deformation are required [63, 104]. This is axiomatic due to the
relative inextensibility [57] of the fibres in comparison to the shear and compressive
resistance of the matrix. Similarly, the effect of such a combined media causes many
problems when trying to model the material as continua [105]. Reasonable prediction of
cross ply uni-directional versus woven fabrics for bias extension tests have been shown
by Potter [106]; however, he reports that in that case use of the pin jointed net model
does not fully encompassing the complexities of the motion. Long et al. [107] assessed
the accuracy of in-plane testing and simulation approaches that use the dominant shear
energy to omit most other secondary energy mechanisms. It is additionally worth noting
the works of Lomov et al. [108, 109] in detailing a number of important measurement
considerations give results for experimental methods that test some of the varying
mechanisms with the woven textile composites. If excessive shearing is required, neces-
sary shear rates are too high or inter-laminar friction causes exorbitant resistance, then
the material will suffer sporadic wrinkling deformities [110, 111]: essentially out-of-
plane Euler buckling. Presence of such wrinkles will mean the formed component is
discarded, and such a result will probably be common to the tool. This exemplifies the
necessity for process simulation—in this case, the need to analyse tool designs for
formability before costly manufacture [58].
As it has been shown that there is difficulty in dealing with deviations between dif-
ferent experimental methods of test specifications [112], where various groups [99, 113]
Chapter 2 Literature review 20

have created testing standards to reduce the incidence, one mitigating technique is to
utilise optical strain measurement [114]. This is otherwise described as full-field strain
measurements [115], and allows the contribution of non-uniform strain to what is ideal-
ised as being uniform conditions in parameter identification, but also quantifies lost or
gained energy subsequent to experimental variations or non-conformities.
The relevance of this digression is primarily for the case study on picture frame test-
ing of trellis shear in Chapter 6, though from all this it is still pointed out that the ap-
proach to modelling is closely tied with experiment and experimental difficulties. The
benefit of continuum constitutive models is that their fundamental mechanical correct-
ness allows some experimental problems to be smoothed out in non-conformity to the
model. On the other hand, data-driven approaches, which will be examined shortly, can
be tied in with high-level data capture, such as is with optical strain measurement.

2.4 Specialised continuum approaches for strong orthotropy

The general task of modelling finite deformation for large deformation shells [116-125],
composites and laminates [123, 126-128], large rotation and corotation [124, 129-133]
and so on [22, 134-136] is a well covered topic. While we direct attention to some no-
table publications [137, 138], here we regard many such methods, particularly those that
use non-orthogonal systems such as with a Cauchy–Green deformation under shear, as
being non-Euclidean insofar as they do not reflect our observational reality. That is to
say, this is whereby a body is not distorted between relative positions in some coordi-
nate system [139]. Instead, in this section, we assess a number of specialised continuum
approaches that characteristically strive for physical analogy.
These approaches trace back to, among others, Mulhern [105] and the same authors
whom developed the kinematically driven models. The laminate theories [138] were
generally not adopted for these large deformation analyses, not because of an inability
to cope with finite deformations but rather because of their being uncomplimentary to-
wards the nonlinear constituent and rate dependent models; though, El-Abbasi and Me-
guid [140] have demonstrated an effective large deformation multilayer element for
composite analysis. Also significant early contributions were made by Pipkin [141] and
Chapter 2 Literature review 21

Conchur et al. [142], and we correspondingly regard the efforts also made towards
transversely isotropic continua [97]. Similar investigation has been directed toward rock
formation that presents slip planes akin to jointed media [143].
Much of the contemporary work falls into the approach where individual fibre fami-
lies adhere to their own material coordinate systems as shown by Spencer [23, 24, 29],
and substantial work by Zheng and Betten [144, 145] furthered the basis of Spencer by
generalising the classes of geometric tensor for most possible fibre and crystalline struc-
tures. We note that uni-directional fibre orthotropy represents just one of a large number
of possible planar material symmetries that were covered by Zheng [146]. His work was
furthered by Qui and Pence [147] to consider reinforcements embedded in neo-Hookean
type base material.
In modelling similar materials, including structurally reinforced elastomeric materi-
als, recent work by Nedjar [148] addressed these same issues by the use of structural
tensors. Building on developments by Klinkel [149, 150] and others, this formulation
allowed the projected components of deformation to be treated uniquely for any number
of fibre families that are additively decomposed. Perhaps the most comprehensive ac-
counts of this approach are given in the exceptional works by Brannon [151-154], al-
though her work differed from those mentioned in that it concerned itself with the nec-
essary rotations of constitutive models rather than with rotation or projective decompo-
sition of the strain. Effectively, this and the various aforementioned efforts have treated
material coordinates as being pinned to fibre direction—whether that be for the benefit
of the strain or the constitutive model—thus the commonality has been to model the
materials not as strongly, but infinitely orthotropic in terms of their deformation modal-
ity.
The concept of ‘preferred directions’, which we refer to as material principal direc-
tions, was further utilised by Kyriacou [155] in order to extend hyper-elastic membrane
FEA to include orthotropic behaviour by using a strain invariant based constitutive
model. While preferred directions bear utility for continuous fibre composites, a similar
fibre orientation tensor exists for less defined short fibre materials [156]. Manach et al.
[157] continued the work in presenting an analysis based on convected coordinate
frames; they specifically note that their model is devoted to fibre materials—said differ-
Chapter 2 Literature review 22

ently, they are not generalised approaches. The idea of preferred directions is also im-
plied in the work of Zheng [158], who paid particular attention to simple shearing of an
orthotropic medium with his arotational forms. This was by reference to preceding work
by Dienes [159], where it was stated that the effect of material rotation is of particular
importance in the analysis of materials with anisotropic properties. Furthering this, Die-
nes [160] anticipated that the effect of material rotation plays a crucial role for stress
rate in connection with anisotropic materials and anisotropic microstructure. Interest-
ingly, efforts by Rajagopal and Srinivasa [161] demonstrate the possibility of represent-
ing materials that indeed have micro-structure without the use of directors or Cosserat
stresses but instead by utilising the knowledge of natural configurations: the salience of
this point will become more prominent later on (see Section 5.2).
On the idea of a generalised approach, i.e. one that remains relevant for isotropic
materials, Itskov [162] presented an orthotropic hyper-elastic material model that is able
to reduce to pre-dating isotropic functions—however this offers not a smooth or inter-
mediary reduction but rather, at best, a jump from weak orthotropy with infinitely
orthotropic kinematics to isotropy with isotropic kinematics, as will any method based
on structural tensors. Christoffersen’s colourful discourse on initially isotropic elastic
materials [163] somewhat encourages investigation and promise towards generality as a
long-standing argument regarding the possible number of elastic constants of a material
are revived—an argument that echoes the remarks of Cauchy shown earlier in the intro-
duction.
Basing an analysis on a classical continuum approach, one can select almost any ar-
bitrary tensor function or even dual variable forms [164] in describing the material’s
kinematics, and extensive work by authors such as Xiao [165] has had success in de-
scribing anisotropic plasticity by combination of even partial and absolute measures. All
things considered, however, the question arises as to how we might approach the task in
a non-arbitrary manner, or maximally mechanistic sense. What is more, we seek not to
do so in a specialised manner, or one that sees a fundamental distinction between isot-
ropy or mild orthotropy and strongly orthotropy properties, but instead a generalised
and smoothly transitioning method. In any light, these major achievements in contin-
Chapter 2 Literature review 23

uum modelling serve to prove that orthotropy does require non-classical treatment, and
so it follows to assert that any level of orthotropy must thence be reconsidered.

2.5 Some relevant micropolar and Cosserat theory

The idea of describing an additional rotational degree of freedom to each orthogonal


plane, or three to the material point, goes back to early suggestion by Poisson [166]—
although it was just before his death—and later by Duhem [167] and E. Cosserat [168];
still, the formulations are generally accredited to F. and E. Cosserat in their more fa-
mous monograph [169]. In this thesis these additional micro-rotations of Cosserat the-
ory are sometimes called extraneous degrees of freedom, because they are not necessary
for the description of materials without micro-structure [170] or body moments [171]—
that is to say that those rotations are otherwise determinate and so need not be described.
A useful depiction of this local micropolar material behaviour is provided by Rovati and
Veber [172].
In its modern form, micropolar theory typically traces back to Eringen [173, 174],
Mindlin [175-178] (though it is not always clear to whom [179]) and others [8, 136,
180-182]; an excellent and concise contemporary appraisal on the subject has been writ-
ten by Ramezani and Naghdabadi [183]. As the subject of this study is the description
of materials that are strongly orthotropic, it is noted that a number of works have focus-
sed on fibre composites and micropolarity. Achenbach’s monograph [184] on micro-
structure for directionally reinforced composites was one of the earliest comprehensive
works on the topic, and some recent studies have continued to tackle the subject, includ-
ing those by Hutapea et al. [185] and Xun et al. [186]. In these investigations into direc-
tional or fibre reinforced materials, the physical microstructure is only present in the
plane normal to the fibre direction. This plane has properties that are called transversely
isotropic, where the material is orthotropic but the normal plane is uniform in all direc-
tions. Where Cosserat theory can describe physical structures such as voids or porosity
[187], we note that in cross-section the resulting circular regions of discontinuous me-
chanical properties are replicated by the fibres in the said transverse plane. Such media
have presented difficult and ongoing characterisation problems [188].
Chapter 2 Literature review 24

The formulation of this transverse isotropy into FEM, studied by authors including
Nakamura et al. [189], was given a particularly practical treatment by Yeh and Chen
[190] to also separate in-plane shear strains by using drilling DOFs for shell elements
and micro-polar theory but without reference to orthotropy or micro-structure: the result
of this was a non-symmetric yet kinematically isotropic strain tensor, in that the differ-
ential in shear pertained to drilling moment as opposed to a more conventional treat-
ment [191].
In the present investigation, our interest is directed towards structural materials in
terms of their manufacture, subsequent design state and generalised mechanical re-
sponse. With strongly orthotropic materials in mind, such as the CFRPs introduced,
there is little interest directed towards fibre-transverse behaviour due to both its lack of
stiffness contribution and its widely executed reduction to a planar or shell description.
Furthermore it is apparent that in any plane other than the rarely considered transverse
one, the directionally reinforced materials will have no true micro-structure or even
analogy to a Cosserat continuum. These aforementioned works will show themselves to
be apparent and relevant as the proposed formulations are found—in that certain
mathematical similarities become obvious—thus the ill physical analogies have been
clearly defined hitherto.
The final consideration made with reference to micropolarity regards strongly
orthotropic media in the typical plane that have been represented with some form or us-
age of Cosserat theory. A number of metaphors have been made to portray the behav-
iour of stratified media, such as the one developed by Adhikary et al. [192] and simi-
larly the ‘bookshelf mechanism’ studied by Figueiredo et al. [193]. We recognise that in
these cases the mechanisms are macro-rotational, though they benefit from a micro-
rotational description primarily because the mathematical formulations suit the task.
These particular representations have been made so with noted success, but it is the al-
ternative and truly analogous physical description that is sought in our objective.
Chapter 2 Literature review 25

2.6 Finite element techniques for behavioural compensation

In this section we refer to techniques for behavioural compensation to help class a num-
ber of significant investigations that have shown effective modelling approaches by
means of procedure or limitation when utilising finite element codes. A classic example
of such a technique is ‘reduced integration’ [118], which is widely recognised for its
ability to dispense with problems of numerical locking, though mostly bears no physical
analogy or added physicality to the representations of real phenomena.
Fernlund et al. [194], for example, gave a method for analysis of process simulation
for composite pre-forms that breaks the analysis down to two-dimensional cross-section
analysis and three dimensional-shell analysis in order to resolve problems with out-of-
plane shear response without the large computational expense. Crookston [195], also,
outlined a program-specific analysis procedure to modelling the forming of CFRPs. In
another approach, Cherouat and Billoet [196] built truss structures with one-
dimensional elements around aligned membrane elements to physically create the ex-
tensibility constraint of the fibres.
Complementary to other endeavours, Yu et al. [103, 197] mitigated an issue to do
with the numerical locking of in-plane shear. This is achieved by asserting mesh align-
ment with fibre directions as a remedy for the over-estimation of intra-ply shear stiff-
ness. This alignment, however, is limited to uni- and bi-directional materials. Continued
and comprehensive investigation by ten Thije and Akkerman [198] assessed three pos-
sible remedies for the shear locking phenomenon. The first of these two, aligned meshes
and reduced integration, have been studied in depth by aforementioned authors and oth-
ers, though the third method presents and interesting approach to the problem. By using
a multi-field element, reasonable success in mitigating the locking problems was pre-
sented. However, it would appear that there is a widespread error in the terminology of
numerical locking, considering that Belytschko et al. [22] clearly defined locking as be-
ing a problem that is irreducible by mesh refinement, and therefore any form of locking
not adherent to this rule is only displaying the incapacity of an element to exactly repre-
sent a non-uniform strain field—which is ultimately one of the underlying principles of
the finite element method. The de-coupling of strain from displacement, by means of
Chapter 2 Literature review 26

linking it to the velocity gradient instead, undoubtedly causes problems due to incom-
patibility and energy conservation, and the authors of those works note the occasional
poor accuracy of the method.
While the efforts of researchers in this category have undoubtedly provided a num-
ber of practical and effective techniques for mitigating the problems associated with the
simulation of strongly orthotropic continua, it prevails that the objective of this study is
not for numerical solutions. Instead, it seeks mechanistic and analogous representations
of deformation to which the accuracy of the solutions and absence of many numerical
problems should follow. Moreover it does not limit itself to the finite element method,
and in place seeks to be a general continuum method.

2.7 Multi-scale and experimental mapping approaches

Significant recent attention has been directed towards an approach that does not explic-
itly model the internal or meso-scale kinematics, but rather maps the response as a
multi-scale analysis or directly to experimental results. These methods can generally be
categorised as being data-driven, as the necessity for complicated constitutive models—
or model-driven approaches—is eliminated; in support of these works, it has been
shown [199] that the need for any constitutive model whatsoever can be removed, a
method which has also been directly applied to visco-plastic composites [200]. Where
CFRPs are typically comprised of fibre bundles (or tows), any woven, knitted or
stitched laminate tends to be representable as repeating unit cells of meso-scale geome-
try. These minor scale units can be simulated for a response that can then be mapped
directly to a fixed grid finite element model. In the same manner, the mapped response
can also be mapped directly to experimental results [91]. As a note to this: it has re-
cently been suggested [201] that validation of data-driven models is not needed; but if
we accept the multi-scale unit-cell mapping as being a data-driven approach then it
stands to reason that the validation of the global model based on mapping against ex-
perimental results is entirely applicable. This has been exemplified in work by Peng and
Cao [101].
Chapter 2 Literature review 27

The early investigations by Boisse et al. [202] and McBride and Chen [203] of unit-
cell geometry produced both simple numerical descriptions of the geometries and re-
sponse of the unit-cells to deformations. This multi-scale method, also referred to as a
global–local numerical approach [204] or as hierarchical modelling [102], has been
used largely for the analysis of dry fabric weaves where bending, shear and compressive
stiffness are near zero [93]. Furthermore, Boisse et al. [94, 205-207], in eliminating any
such continuum assumptions, used this multi-scale method to accurately model sub-
structural response of woven fibre composites in forming and successfully mapped the
response to various draping analyses; although due to the non-objective nature of the
approach, their meshes were limited to aligned and fixed grids correlating directly to the
unit-cell orientations. This means that no such approach can pass the patch test [208].
The potential efficacy of the technique for modelling dry woven fabrics was further
demonstrated in the correlations by Badel [209] and Hivet [210] in association with
previous authors, and highly successful non-continuum models for non-crimp fabrics
were created by Creech and Picket [211, 212]. While the model construction of such
detailed geometries may seem involved [213-215], Robitaille and others [216] have pre-
sented an algorithm allowing automatic generation based on minimal parametric inputs.
In light of this, though, practicable determination of such parameters for complex lami-
nates can provide a significant challenge [217].
A simpler representation was considered by Sidhu, Averill and Riaz [218] in a unit
cell model using truss elements to directly model fibre segments; this was extended by
King et al. [219] by separating the truss elements by springs so as to represent yarn
crossovers and slight geometric fibre extensibility. These representations, however, add
to the disadvantages of the unit cell method in that they are also discontinuous on the
major scale. In simulating this particular forming problem, perhaps the most practical
method was developed by Yu et al. [91, 103], whose investigative efforts concluded
that a simple remedy to certain in-plane shear locking problems was to strictly align
meshes with material orientation. Whilst mitigating the numerical locking, this method
also allowed the material response to be mapped directly against experimental results,
thus eliminating the need for specifically derived tensors or complex sub-structural
models.
Chapter 2 Literature review 28

This mapping method, and in fact all the multi-scale methods highlighted, cannot be
considered a continuum mechanics solution, and requires significant empirical data.
Along with this, under finite deformation all these methods suffer from the restrictions
of fixed-grid analysis and a lack of simple parametric adaptability. Nonetheless, the ef-
ficacy of this general approach is noted, and to that tune tends to strengthen the argu-
ment towards the inability of the current continuum mechanics realm.

2.8 Chapter summary

On examination of a wide range of published works and extensive investigative efforts


it remains that there is at least some error associated with any continuum description of
an orthotropic or anisotropic material under finite and unaligned distortive strain. Fur-
thermore, there is evidence present throughout this thesis that positively suggests the
error relates to a mechanics issue that has not, as yet, been addressed. While none of the
continuum methods are able to kinematically model the intermediary case of strongly
orthotropic behaviour in an exact manner, there has been substantial achievement in the
non-continuum approaches, such as the multi-scale methods—yet their success has
mostly been in the simulation of discontinuous media. Whereupon one might suggest
that these multi-scale approaches will eventually satisfactorily encompass the general
orthotropic kinematic, the problem can be shown to reduce to a simple continuous prob-
lem having no sub-structure. It is with this case that we commence a discussion in the
next chapter, because there is no reason to think that only discontinuous descriptions
exist for a medium that soundly resides in the continuum assumption.
Chapter 3
Strongly orthotropic continuum mechanics
In this chapter, the various components of the proposed linear continuum mechanics
basis are presented, including primarily the derivation of the orthotropic linear strain
tensor. Here, the approach and formulations are demonstrated for a two-dimensional
case, which is most common for orthotropic analysis as applied to shell structures. In
selected cases, the extension of equations into three-dimensional form is included.

3.1 Mechanics of orthotropic materials

3.1.1 The linear tensors of the displacement state


We first consider various denotations and formulations used in the continuum mechan-
ics of orthotropic materials necessary to describe the proposed approach. To clearly
convey the observational frame of each formulation, let us denote four classes of coor-
dinate system for planar analysis:

 x, y : Arbitrary global Cartesian


 1, 2 ,  1, 2 Ž % 1, 2 : Fiber family subsets of the material-principal
( 1, 2 : Principal strain orientation
& [ ,K : Element natural

where the axis indices are shown in parentheses. Note that the material system is always
aligned with the material principal properties.

29
Chapter 3 Strongly orthotropic continuum mechanics 30

y, Y F X,t
u :
: 0 : t0 0
:0 u* :: t t
X :* : t* t  a
x :*
x* F * X, t *
x, X

Figure 3.1. A continuum body under some deformation and displacement. The motions F and F* repre-
sent different frames of observation.

As illustrated in Figure 3.1, the coordinates of some undeformed material point are

^` ^` ^%` ^%`


> X1, X 2 @
T T
X ª¬ X x , X y º¼ or X , (3.1)

where (·)T indicates the transpose, Xx becomes X and Xy becomes Y. Under the condition
of linear infinitesimal deformations, any variation between the current (deformed) coor-
dinate frame using x and the reference (undeformed) frame using X vanishes. Similar to
these, the deformation vectors, denoted as u, are

^` ^` ^%` ^%`


>u1 , u2 @
T T
u ª¬u x , u y º¼ or u , (3.2)

where ux becomes u, and uy becomes v as is conventional.


Using these coordinate systems, the infinitesimal strain tensore, or ILST, can be de-
fined in Voigt notation by

^%`
^`
>e11 , e22 , e12 @
T T
e ª¬exx , eyy , exy º¼ or e , (3.3)

where for the rest of this thesis the addition of an overline bar indicates the same deno-
tation but specifically under conventional treatment; for example, in the cases of princi-
pal ( and material % systems. The conventional stress Ve, denoted as conjugate to
the ILST by (·)e, and the strain tensor are related by the rank-four tensor D, known as
the material stiffness matrix:

^%` ^%`
D ^ ` e or
%
Ve (3.4)
^`
Ve ^`

D ^ ` D, T
%
^`
e, (3.5)
Chapter 3 Strongly orthotropic continuum mechanics 31

noting that the global stiffness matrix is given as a function of the local material stiff-
ness matrix and projected angles in the material principal orientation matrix T. Note that
Tij Tji, and that the three-dimensional form is an anti-symmetric matrix as will be
shown in Equation (3.18). Using the displacement gradient ’u, the isotropic linear
strain tensor field can be expressed in tensor form as

1 K L
e
2
’u  ’u , or as (3.6)

1 § wui wu j ·
eij ¨  ¸¸ (3.7)
2 ¨© wX j wX i ¹

with components shown in index notation. Note that the term tensor field is sometimes
abbreviated as tensor; and that e12 and exy are the shear tensor strains equal to half the
engineering shear strain J.
By considering that the ILST is the symmetric part of the displacement gradient, the
infinitesimal rotation tensor Z is given by the residual skew symmetric component such
that
K
’u e  Z. (3.8)

This additive decomposition yields a rank-two tensor representing isotropic rigid body
rotation

1 K L
Z
2
’u  ’u , (3.9)

 without loss of informa-


which can be expressed as components of a rank-one tensor Z
tion using the unusual tensor contraction as follows:

1 § wui wu j ·
Z iu j ¨  ¸¸ (3.10)
2 ¨© wX j wX i ¹

noting the use of the cross product in the indices such that Z iu j Z k and so on.
Chapter 3 Strongly orthotropic continuum mechanics 32

3.1.2 Strongly orthotropic kinematics and deformation

 


a) Undeformed b) Freely shearing c) Deformed

Figure 3.2. a) Trellis kinematics where the connectors can be considered as inextensible pivoted fibres;
b)–c) Material family  exhibits no rigid rotation while family  rotates through twice the shear tensor

strain

In the case of a biaxial and infinitely orthotropic material, as kinematically represented


in Figure 3.2, an energy-free shearing mode described as trellis shear exists. This en-
ergy-free motion is possible because of the lack of shear resistance and the absence of
fibre-normal resistance. The material principal directions are considered as unique over-
laid continua each rotating a local material coordinate system  and . Addition of a

relatively soft isotropic medium to the interstitial regions thence results in a softly shear-
ing strongly orthotropic material that has a fibre-normal resistance. The fibre-normal
component of strain is zero under infinitesimal deformation, second-order under small
deformation and first-order under subsequent deformation of updated configurations if
the kinematic restriction is maintained.

%0 , %0 % %0

a) Undeformed b) Macro scale c) Meso scale d) Continua e) Isotropic continua

Figure 3.3. a)–d) An infinitely orthotropic material shears along the horizontal principal material axis
freely and without rigid body rotation; e) The isotropic material undergoes rotation equal to the shear ten-
sor strain

Shown in Figure 3.3 is an infinitely orthotropic response resulting in an entirely unbal-


anced shear mode, or simple shear. The persistently orthogonal material coordinate sys-
tem demonstrates that the body has undergone zero rigid body rotation; which is in con-
Chapter 3 Strongly orthotropic continuum mechanics 33

trast to the rotation tensor of Equation (3.9) and as illustrated in Figure 3.3 e). Despite
the differing kinematic descriptions of deformation leading to d) and e), these modes are
geometrically compatible. Note that this infinitely orthotropic concept has been well
illustrated by Brannon [154].

3.2 The assertion of geometric strain indeterminacy

In the introduction to this thesis a historical appraisal of the development of the most
widely used strain tensors was given. Having shown that the symmetry of these tensors
was based on mathematical assumption, a rigorously derived linear strain tensor is
sought. In doing so, the first step is to assert that using only the initial and deformed ge-
ometries and no such assumption of symmetry leads to a strain state that is indetermi-
nate. This is shown as follows.
The widely used linear strain tensor expressed in Equation (3.7) can be considered
as the indeterminate strain state

wui wu j
eij  e ji  . (3.11)
wX j wX i

and an assumption of tensor symmetry, whereby

e ji eij . (3.12)

This also infers the in-plane shear moduli are equal, that is

Dijij D jiji , (3.13)

If we discard the symmetry assumption on the reasoning that it has no derivable basis,
then an Assertion of Geometric Strain Indeterminacy follows. This becomes an integral
basis for the developments hereafter.

3.3 The orthotropic linear strain tensor

For the determination of the orthotropic deformation state, we require that the strain
tensor is no longer symmetric, and so it is extended outside of Voigt notation and in-
Chapter 3 Strongly orthotropic continuum mechanics 34

stead vectorised accordingly. This, the so called OLST, is defined as follows for mate-
rial and global coordinate frames:

^%`
>œ11 , œ22 , œ12 , œ21 @
T
œ (3.14)

^` T
œ ¬ªœxx , œ yy , œxy , œ yx ¼º (3.15)

where, because a linear strain tensor with values unique to the ILST is defined, the de-
notation œ with components œij is used rather than e (such that ethel, the ligature oe,
abbreviates orthotropic strain). Considering there is now the mathematical possibility of
a non-symmetric stress tensor, we must differentiate the Orthotropic Cauchy Stress
Tensor Vœ from the isotropic Cauchy stress tensor as a metric being energy conjugate to
the OLST rather than the ILST. Concurrently, we redefine the constitutive equation as

^%` ^%`
Vœ D ^%` œ or (3.16)
^`
Vœ ^`
D ^%`
D, T ^ ` œ.

(3.17)

The generalised 3D material principal orientation matrix used is simplified to an anti-


symmetric matrix:

ª 0 T xy T xz º
« »
T « T xy 0 T yz » (3.18)
« T xz T yz 0 »¼
¬

By the proposed basis, the strain—available in any arbitrary coordinate frame—


should be additionally dependent on the material property set, which includes the con-
stitutive behaviour and orientation angle of the material principal direction. Thus it is
expected that

^`
œ f ^ ` ’u , ^
 %`
D, ^` T . (3.19)

Here, the strain state analogous to Equation (3.11) is given by

wui wu j
œij  œ ji  , (3.20)
wX j wX i
Chapter 3 Strongly orthotropic continuum mechanics 35

and consequently the assumption of Equation (3.13) is discarded; hence the shear terms
^%` ^`
in the local and global material stiffness matrix D and D from Equations (3.16)
and (3.17) respectively must be treated as unique:

ª D1111 D1122 º ª Dxxxx Dxxyy Dxxxy Dxxyx º


" 0" «
«D » Dyyyy Dyyxy Dyyyx »»
^%`
D « 1122 D2222 », ^`
D « (3.21)
« D1212 0 » « # Dxyxy Dxyyx »
« " 0" » « »
¬ 0 D2121 ¼ ¬« sym " Dyxyx ¼»

The derivation benefits from noting Quintanilla and Straughan [220] on uniqueness of
tensor functions in thermo-elasticity, with remarks regarding isotropy and anisotropy:

“Uniqueness is proved under the weak assumption of requiring only major symmetry of
the elasticity tensor; no definiteness whatsoever is postulated.”

As material properties are specified in the former system, we require the relationship
of the global material stiffness matrix to the local properties. To obtain this, we first re-
quire the transformations T from the material coordinate frame to the global system for
strain and stress:

^`
œ T ^ ` œ and
%
(3.22)
^`
Vœ T ^%` V œ , (3.23)

where the corresponding extended but typically conventional stress tensor quantities are

^%` T
Vœ ª¬V 11œ , V 22
œ
, V 12œ , V 21
œ
º¼ and (3.24)

^` T
Vœ ª¬V xxœ , V yyœ , V xyœ , V yxœ º¼ . (3.25)

Using the transformation law for Cartesian tensors [221], the components are derived
using

^`
œij lip l jq ^%`œ pq and ^`
V ijœ lip l jq ^%`V œpq , (3.26)

noting the summation convention for free indices, and the direction cosines lij and ab-

breviations:
Chapter 3 Strongly orthotropic continuum mechanics 36

ª c sº
l « s c» , c cos T xy , s sin T xy (3.27)
¬ ¼

Thus by use of Equation (3.26) the transformation matrix is obtained as follows:

ª c2 s2 sc sc º
« 2 »
«s c2 sc sc »
T (3.28)
« sc sc c 2 s 2 »
« »
¬ sc sc s 2 c 2 ¼

Note that by the proposed formulations, there is no longer a need for a different trans-
formation for global tensor strain and global engineering strain. Using this transforma-
^` ^%`
tion matrix, D can be related to the elements of D by the procedure

^`
D T1 ^%` D T, (3.29)

where the global material matrix required by Equation (3.17) adopts potential shear
coupling terms as shown in Equation (3.21). The subsequent material stiffness matrix is

«

ª f1 ^%` D, T f9 ^ %`

D, T ^ ` D, T
f8 %
f7 ^ %`

D, T º
»
« f ^ `
D, T f ^ ` D, T f ^
D, T »»
% % %`
^` « 2 10 6
D « », (3.30)
f ^ ` D, T f ^
% %`
« 3 5 D, T »
« »
f ^
%`
« " sym" D, T »
¬ 4
¼

with ten unique functions. Following the operation of Equation (3.29), the direct diago-
nal terms are

f1 c 4 D1111  2 cs
2 2
D1122  s 4 D2222  c 2s 2 D1212  c 2s 2 D2121
f2 s 4 D1111  2 cs
2 2
D1122  c 4 D2222  c 2s 2 D1212  c 2s 2 D2121
(3.31)
f3 c 2s 2 D1111  2 cs
2 2
D1122  c 2s 2 D2222  c 4 D1212  s 4 D2121
f4 c 2s 2 D1111  2 cs
2 2
D1122  c 2s 2 D2222  s 4 D1212  c 4Q2121

and the various coupling functions result as follows:


Chapter 3 Strongly orthotropic continuum mechanics 37

f5 c 2s 2 ( D1111  2 D1122  D2222  D1212  D2121 )


f6 cs(s 2 D1111  s 2 D1122  c 2 D1122  c 2 D2222  s 2 D1212  c 2 D2121 )
f7 cs(c 2 D1111  c 2 D1122  s 2 D1122  s 2 D2222  s 2 D1212  c 2 D2121 )
(3.32)
f8 cs(c 2 D1111  c 2 D1122  s 2 D1122  s 2 D2222  c 2 D1212  s 2 D2121 )
f9 c 2s 2 D1111  c 4 D1122  s 4 D1122  c 2s 2 D2222  c 2s 2 D1212  c 2s 2 D2121
f10 cs(s 2 D1111  s 2 D1122  c 2 D1122  c 2 D2222  c 2 D1212  s 2 D2121 )

Notably, these equations cancel down to conventional transformation equations when


the omitted condition of Equation (3.13) is included.
We utilise these equations in determining the shear terms of the OLST (œij: i z j),
and it is necessary to use the constraint of static in-plane moment equilibrium implied
by the ILST, whereby

V xyœ V yxœ (3.33)

This is taken for granted in accordance with Green and Naghdi [222] who, on thermo-
viscous fluid theory (noting T is the Cauchy stress tensor), make the commentary:

“In the usual development of the classical field theory, the symmetry of the stress tensor,
namely T = TT (2.20), is deduced from the balance of moment of momentum. However,
without appeal to this balance principle, as indicated in Ref. [223], the result (2.20) may
be regarded as a consequence of invariance under superposed rigid body motions.”

By Equation (3.30), the equilibrium becomes

f8œxx  f1 œ yy  03
f œxy  f5œ yx f 7 œxx  f 6 œ yy  f5œxy  f 4 œ yx (3.34)

and by Equation (3.20) three strain components can be eliminated using

wu wv wu wv
œxx , œ yy and œ yx   œxy . (3.35)
wX wY wY wX

With minor simplification, an explicit description for one shear strain term is achieved:

ª wu wv wu wv º
f8  f 7  f1  f06  f5  f 4 §¨  ·¸ »
1
œxy « (3.36)
2 f5  f3  f 4 ¬ wX wY © wY wX ¹ ¼

By substitution of equation sets (3.31) and (3.32), the shear strain is represented in
terms of the elements of the local material stiffness matrix
Chapter 3 Strongly orthotropic continuum mechanics 38

§ wu wv ·§ D1212 D2121 ·
œxy ¨  ¸¨ s2  c2 ¸
© wY wX ¹ © D1212  D2121 D1212  D2121 ¹
(3.37)
§ wu wv · D1212  D2121
¨  ¸ sc,
© wX wY ¹ D1212  D2121

which satisfies the expectation of Equation (3.19) and is determinate despite the re-
duced-assumption strain relations of Equation (3.20). As a check for validity, we see
that under the hitherto omitted relation of Equation (3.13), this reduces to:

1 § wu wv ·
œxy ¨  ¸ , given D1212 D2121 (3.38)
2 © wY wX ¹

which is identical to the ILST, as shown in Equation (3.7). The second shear strain term
in the planar tensor is found by the same procedure, and results in the same form as
Equation (3.37) with cycled indices. Given this and Equation (3.35), the proposed
OLST can be compactly defined as follows:

§ wu wu j · § wu wu j · dij  d ji
¨¨ i  ¸¸ dij sin Tij  d ji cos Tij  ¨¨ i  ¸¸ sin 2Tij
2 2
œij (3.39)
© wX j wX i ¹ © wX i wX j ¹ 2

where, for simplification, we define a term d as the balance ratio with components

^%`
^` Dijij
dij ^%`
: % Ž , (3.40)
Dijij  ^%` D jiji

such that it is always defined by the properties in the local material frame.

3.4 The material principal rotation tensor

In order to thence describe the equivalent rotation tensor, we first express the OLST
in a similar manner to the tensor form of the ILST as in Equation (3.6):
K L
œ ’u  ’u <r a
 D<r b , where (3.41)

ra sin 2
T d  d cos 2 T  , I s and r b sin 2T< d  d T , (3.42)
Chapter 3 Strongly orthotropic continuum mechanics 39

noting that a dot < between matrices indicates the Hadamard product, Is is the skew di-
agonal identity matrix and D is the Aspectual Strain with components

1 § wui wu j ·
D ij ¨  ¸¸ . (3.43)
2 ¨© wX i wX j ¹

This allows us to use the same additive decomposition as Equation (3.8) such that
K
’u œ  \ (3.44)

so as to derive the Material Principal Rotation Tensor \ as follows:


K L
\ ’u<r a T  ’u<r a  D<r b (3.45)

This generalisation can replace the isotropic rigid body rotation tensor Z. Using the
same contraction as Equation (3.10), this reduces to a rank-one tensor \ that has com-
ponents as follows:

wui 2 wu
\ iu j
wX j
s T ji d ji  c 2 T ji dij  j s 2 Tij dij  c 2 Tij d ji
wX i
(3.46)
§ wu wu · dij  d ji
¨ i  j
¨ wX wX ¸¸ s 2Tij
© i j ¹ 2

If applied to fibre reinforced composites, this represents the change in fibre direction
rather than geometric rigid rotation—as is illustrated in Figure 3.3. As with Equation
(3.39), we note that \ reduces to Z when isotropic properties are specified (i.e., dij = ½).
Note that all generalised tensor and indicial forms are applicable to the three-
dimensional case * provided the material coordinate system is only rotated in one plane.
For compound rotations, the same procedure can be used for derivation, though these
are rarely necessary due to the dominant use of plates and shells for strongly orthotropic
applications. For the case of a transversely isotropic material, the two-dimensional
novel formulations would reduce to convention. Nonetheless, consideration of any out-
of-plane bending, where assumed shear profiles or bending coefficients are used, leads
to the conclusion that updated formulations would be necessary.

*
That a three dimensional formulation can be derived is a mathematical certainty, though it is omitted for its lack of contribution to
the impetus of this theory. It will be presented in future works.
Chapter 3 Strongly orthotropic continuum mechanics 40

3.5 Physical interpretation of the OLST and its components

The OLST presented in Equation (3.41) introduces some unconventional strain compo-
nents including the separation of shear strain into two distinct skew shear components,
and the metric denoted as aspectual strain: these each have a meaningful physical inter-
pretation. Aspectual strain produces a positive value when the aspect ratio increases, a
negative value for a decrease, and no change for pure planar dilatation strain (area
change only). Thus pure aspectual strain represents deviatoric change only as is done by
the shear components of the ILST: though this is at 45° to the maximum shear orienta-
tion. This, the principal orientation, is where the shear component conventionally di-
minishes to zero, and is where aspectual strain provides a method for identifying the
latent shear.

Table I. Strain categories for planar modes: columns indicate modal change, rows indicate sign change

Dilatation strain Deviatoric strain

Pure Normal Pure shear Aspectual Skew shear

+ + + + + + +

The aspectual strain and other characteristic changes are summarised in Table I. Note
that pure Normal strain (eii, œii) is a combination of a planar dilatation strain (ed, œd) and
aspectual strain (Dij: i z j); and, that Pure shear (eij: i z j) can be considered as a vertical
and horizontal Skew shear in combination (œij: i z j). The separation of shear into skew
components is necessary to identify non-geometric or non-isotropic rotation, and the
asymmetry of the strain state.
Chapter 3 Strongly orthotropic continuum mechanics 41

3.6 The extended Mohr plot for strain

œxy xy orthotropy: E1 ! E 2


D: Aspectual strain (œyy,œxy) 
U: Deviatoric radius
y

2I
^%,`

œyy œd : Dilatation strain

œxy U
D
Compressive strain O1 O2 O1 O2 Tensile strain
^%,`
œ œ
x (œxx,œyx) 
Trellis shear œd œ yx yx orthotropy: E 2 ! E1

Figure 3.4. Characteristic strain indicators of the extended Mohr plot

The utility of the differentiated components described in the previous section are best
illustrated by utilising an extended Mohr plot for strain; which, by the proposed contin-
uum mechanics basis, is no longer bound to the horizontal axis. In Figure 3.4 an exam-
ple of the extended Mohr plot is shown for the trellis motion of an infinitely orthotropic
material with a fibre angle of Txy = 0, where dxy = 1 such that œyx = 0. As with a conven-
tional Mohr plot for strain, I represents the angle between the measured global coordi-
nate system and the horizontal orthonormal principal axis. The strain plot, no longer
bound to the horizontal axis except when isotropic, illustrates the necessity for aspectual
strain. Here, vertical deviation from the tensile–compressive axis conveys the level of
material orthotropy and in the same manner that the horizontal position of the circle
centre indicates the state of planar dilatation strain, the vertical position is indicative of
the maximum aspectual strain. Note that the horizontal skew plane increases with the
upwards shear axis xy, and that vertical skew increases with the downwards shear axis
yx.
Another significant feature of the extended Mohr plot is that there no longer exists
an orthogonal principal plane for any deviatoric stress state of non-isotropic materials.
A decomposition of the strain tensor

>v O @ eig œ , O  ( , (3.47)

where, for brevity, ‘eig’ represents the Eigen decomposition as a function. This yields
eigenvalues O = diag(O1,O2) in two dimensions, representing the strain values for each
principal orientation, and eigenvectors v = [v1 v2] indicating the non-orthonormal prin-
Chapter 3 Strongly orthotropic continuum mechanics 42

cipal axis. The two principal orientations are conventionally horizontally opposed coor-
dinates on the Mohr plot such that 2I = 180°, but in the extended Mohr plot, as in Fig-
ure 3.4, this condition only holds for the isotropic case. If the motion of Figure 3.2 is
considered, the absence of an orthonormal principal coordinate system appears logically
correct. If necessary, the conventional principal strains can be recovered using

ª1 º
ª¬ v O º¼ eig « œ  œT » , O  ( , (3.48)
¬2 ¼

noting that the OLST can be degenerated at any time via the operation e œ  œ
T
2.

Having presented the application of the unbounded Mohr plot to our measures of
strain, we draw attention to an informal document concerned with Mohr’s circle appli-
cations: Brannon [224] makes an interesting remark after an attempt to illustrate what
three-dimensional Mohr plots * might look like for non-symmetric stress tensors:

“Clearly there is structure here, but its mathematical description is not obvious. Perhaps
working with a different shear measure would prove useful?”

While the general form and interpretation of an unbounded Mohr’s circle has not been
reached, noting previous application to the non-symmetric stresses of micropolar con-
tinua, it can be supposed that the OLST and other orthotropic measures yet to be shown
may positively answer Brannon’s question.

3.7 The extended orthotropic compliance tensor

Because the OLST is derived on the basis that in-plane shear moduli need not be equal,
an extended compliance and material stiffness tensor that relates to the unequal shear
strains must be defined. For the strongly orthotropic formulation, a linear constitutive
equation can be defined as:

œij SijklV klœ , (3.49)

where the Extended Orthotropic Compliance Tensor S, necessary to the proposed con-
tinuum mechanics basis, is defined with components as follows:
*
The realisation of an extended Mohr plot for three-dimensional strains is not attempted here.
Chapter 3 Strongly orthotropic continuum mechanics 43

^%` G ik G jl 1 Q ji  G ijG klQ lj


Sijkl (3.50)
E l

Here Eij is the Kronecker delta, and E l denotes the components of the elastic moduli

vector as in Equation (3.53). The Poisson ratio, in the usual way, is defined as the ratio
of

œii
Q ij  (3.51)
œ jj

following uni-axial loading. The expanded matrix form of Poisson ratio, following, is
given with example values shown for an incompressible material with an orthotropic
ratio of 100 to 1:

ªQ 11 Q 12 Q 13 º ª 1 0.5 0.5 º
«Q » «0.005 1 0.995»»
Q « 21 Q 22 Q 23 » «
(3.52)
«¬Q 31 Q 32 Q 33 »¼ «¬0.005 0.995 1 »¼

Here, due to incompressibility, the sum of all rows is zero. Furthermore, the elastic
moduli vector requiring only the diagonal terms for general orthotropy is given by


E ª¬ E1 , E 2 , E 3 º¼ .
T
(3.53)

Utilising these, the un-contracted three-dimensional constitutive model written as

>œ11 , œ22 , œ33 , œ23 , œ31 , œ12 , œ32 , œ13 , œ21 @


T T
S ª¬V 11œ , V 22
œ
, V 33œ , V 23
œ
, V 31œ , V 12œ , V 32œ , V 13œ , V 21
œ
º¼ (3.54)

will have the rank-four extended orthotropic compliance tensor unfolded in matrix form
as follows:
Chapter 3 Strongly orthotropic continuum mechanics 44

ª 1 Q 21 Q 31 º
« E E E »
« 1 2 3
»
« Q 12 1 Q 32 »
« E " 0" " 0" »
« 1 E 2 E 3 »
« Q 13 Q 23 1 »
«  »
« E1 E 2 E 3 »
« 1 Q 32 »
« 0 0 »
« E 3 »
{%}
« 1 Q 13 »
S « " 0" 0 0 " 0" »
« E1 »
« 1 Q 21 »
« 0 0 »
« E 2 »
« 1 Q 23 »
« 0 0 »
« E 2 »
« 1 Q 31 »
« " 0" " 0" 0 0 »
« E 3 »
« »
« 1 Q 12 » (3.55)
0 0
«¬ E1 »¼

{%} { }
This material system matrix S can be transformed into a global system S by use
of a transformation matrix derived from the tensor transformation law written in Equa-
tion (3.26).

3.7.1 The extended orthotropic stiffness matrix for plane stress


Furthermore, an extended orthotropic stiffness matrix D for the plane stress condition
can be derived by truncation and inversion of the nine-by-nine matrix form of the com-
pliance tensor—as shown in appendix Equation (3.55). This becomes

ª E1 v12 E 2 º
« »
«1  v12 v21 1  v12 v21 " 0" »
« v E E 2 »
« 21 1 »
^%` «1  v12 v21 1  v12 v21 »
D « E 2 », (3.56)
« 0 »
« 1 Q 21 »
« " 0" »
« E1 »
0
«¬ 1 Q 12 »¼
Chapter 3 Strongly orthotropic continuum mechanics 45

which, unconventionally, indicates that the shear moduli are linearly dependent on the
elastic moduli and Poisson ratio for the orthotropic case. Considering the in-plane Pois-
son ratios are related, three rather than four [225] elastic constants are required as a
minimum. For comparison with a conventional three-by-three material matrix, a single
term that relates to average shear as in the ILST is required. Given that for equivalence
between the ILST and OLST, V 12e must equal V 12œ and V 21
œ
(in their respective systems)
and using this equality the conventional shear stiffness can be found such that

D1212 e12 D1212 œ12 D2121œ21. (3.57)

Because the material system %(1,2) relates to a principal material angle of zero, and by

substitution of the strain terms from Equations (3.7) and (3.39), the equation reduces to
the degeneration

2 D1212 D2121
D1212 . (3.58)
D1212  D2121

Although derived in an alternate manner, this formula is identical to the equation used
to find equivalent properties for composite materials; that is, the well known rule of
mixtures [225].

3.7.2 The extended orthotropic stiffness matrix for plane strain


The general form for plane strain elasticity can be derived by inverting Eq (3.55) and
taking the required rows/columns:

ª 1 Q 23Q 32 E1 Q 21 Q 23Q 31 E1 º


« »
« Q pe Q pe " 0" »
« Q 12 Q 13Q 32 E 2 1 Q 13Q 31 E 2 »
« »
^%` « Q pe Q pe »
D (3.59)
« E 2 »
« 0 »
« 1 Q 21 »
« " 0"
E1 »

« 0 »
«¬ 1 Q 12 »¼
where Q pe is defined as
Chapter 3 Strongly orthotropic continuum mechanics 46

Q pe 1 Q 12Q 21 Q 23Q 32 Q 31Q 13 Q 12Q 23Q 31 Q 21Q 32Q 13 (3.60)

Note that while Q12 can be related to Q21 using its proportionality to E1 E 2 , the less

simplified form of the equations are preferential for the sake of clarity.

3.8 Orthotropic von Mises stress invariant

Despite the stress tensor being symmetric in most cases, where V xyœ V yxœ , its potential
asymmetry means the often used von Mises stress can no longer be simplified by the
assumption of equal in-plane shear stresses. Using the invariants for the rank-two stress
tensor

I1 trace V œ , (3.61)

I2
1ª 2
2 ¬«
I1  trace V ¼» and
œ 2º
(3.62)

I3 det V œ , (3.63)

the Orthotropic von Mises Stress Invariant can be determined for the non-symmetric
case using

V vm
œ
I12  3I 2 . (3.64)

Thus the equation necessary to the proposed continuum mechanics basis becomes
0.5
V vm
œ 1 ª œ
2 ¬«
xx yy yy zz zz xx xy yx yz zy zx xz ¼» . (3.65)
V  V œ 2
 V œ
 V œ 2
 V œ
 V œ 2
 6 V V
œ œ
 V V
œ œ
 V œ œ º
V

In the case where drilling degrees of freedom, dynamic body forces or micro-polar
[190] stress couplings are added, the formulation will differ in value from the conven-
tional von Mises stress.
Chapter 3 Strongly orthotropic continuum mechanics 47

3.9 The anisotropic linear strain tensor

The proposition of an intrinsically dependent strain state is by no means limited to prin-


cipally orthotropic materials. In this work, only non-isotropic materials that have no
orthotropic principality are referred to as being anisotropic. The salient difference there-
fore being that an Anisotropic Linear Strain Tensor (ALST) can be ascertained from the
intrinsic material properties at any arbitrary orientation as opposed to specifically the
material principal orientation as with the OLST. Considering this will represent strain
values unique to the ILST e and the OLST œ, we denote the ALST as æ with compo-
nents æij (such that ash, the ligature ae, abbreviates anisotropic strain).
Based on this, the key difference to the previous derivation is the replacement of the
stiffness matrix of Equation (3.21), from which the tensor is derived, with an anisot-
ropic local material stiffness matrix:

ª D1111 D1122 D1112 D1121 º ª Dxxxx Dxxyy Dxxxy Dxxyx º


« « Dyyyx »»
^%` « D2222 D2212 D2221 »» ^` « Dyyyy Dyyxy
D , D (3.66)
« # D1212 D1221 » « # Dxyxy Dxyyx »
« » « »
¬ sym " D2121 ¼ «¬ sym " Dyxyx »¼

Due to the additional coupling terms, the transformation functions in Equation (3.30)
become more complicated. Following the same derivation from Equation (3.29), the
transformation functions can be determined that include additional dependency on the
anisotropic coupling terms in the material system seen in Equation (3.66). The primary
functions are

f1 c 4 D1111  s 4 D2222  2cs 3 D2221  D2212


c 2 s 2 2 D1122  2 D1221  D2121  D1212  2c 3 s D1112  D1121
f2 c 4 D2222  s 4 D1111  2cs 3 D1112  D1121
c 2 s 2 2 D1122  2 D1221  D2121  D1212  2c 3 s D2212  D2221
(3.67)
f3 c 4 D1212  s 4 D2121  2cs 3 D2221  D1121
c 2 s 2 D1111  D2222  2 D1122  2 D1221  2c 3 s D1112  D2212
f4 c 4 D2121  s 4 D1212  2cs 3 D2212  D1112
c 2 s 2 D1111  D2222  2 D1122  2 D1221  2c 3 s D1121  D2221 ,
Chapter 3 Strongly orthotropic continuum mechanics 48

while, additionally, the coupling terms are ascertained using the functions

f5 c 4 D1221  s 4 D1221  cs 3 D2221  D2212  D1112  D1121


c 2 s 2 D1111  D2222  2 D1122  D2121  D1212  c 3 s D1112  D1121  D2221  D2212
f6 c 4 D2221  s 4 D1112  cs 3 D1111  D1122  D1212  D1221
c 2 s 2 2 D1121  2 D2212  D1112  D2221  c 3 s D2121  D2222  D1221  D1122
f7 c 4 D1121  s 4 D2212  cs 3 D1122  D2222  D1221  D1212
 c 2 s 2 D2212  D1121  2 D1112  2 D2221  c 3 s D1111  D1122  D1221  D2121
(3.68)
f8 c 4 D1112  s 4 D2221  cs 3 D1221  D1122  D2121  D2222
c 2 s 2 2 D2212  2 D1121  D2221  D1112  c3 s D1111  D1122  D1221  D1212
f9 c 4 D1122  s 4 D1122  cs 3 D2221  D2212  D1112  D1121
c 2 s 2 D2222  D2121  D1111  D1212  2 D1221  c 3 s D1112  D1121  D2221  D2212
f10 c 4 D2212  s 4 D1121  cs 3 D1111  D1122  D1221  D2121
c 2 s 2 2 D1112  2 D2221  D1121  D2212  c 3 s D1122  D2222  D1221  D1212 .

Given these anisotropic transformation functions, substitution into Equation (3.36)


yields a determinate form for the xy shear term of the ALST. Repetition of this proce-
dure yields the remaining terms, and so the anisotropic linear strain tensor can be de-
rived. It is represented compactly by the indicial formulation

§ wu wu j · a wu b wu j c
æij ¨¨ i  ¸¸ rij  i rij  rij , (3.69)
© wX j wX i ¹ wX i wX j

where the distribution ratio components are defined as

^` s 2 Dijij  c 2 D jiji  Dijji  sc Diiij  Diiji  D jjij  D jjji ½


r a
°
Dijij  2 Dijji  D jiji
ij
°
°
^` scDijij  scD jiji  c Diiji  c Diiij  s D jjij  s D jjji °
2 2 2 2

rijb ¾ D  ^%` . (3.70)


Dijij  2 Dijji  D jiji °
^` scDijij  scD jiji  s Diiji  s Diiij  c D jjij  c D jjji °°
2 2 2 2

rijc
Dijij  2 Dijji  D jiji °
¿

Furthermore, the ALST can be defined without reference to properties in the mate-
rial principal frame, noting that the simplification from working with a material princi-
Chapter 3 Strongly orthotropic continuum mechanics 49

^%`
pal orientation is lost when there are no non-zero terms in D . As the orientation an-
gles in T therefore go to zero, Equations (3.69) and (3.69) can be reduced to

§ wu wu j · wu wu j ½
¨¨ i  ¸¸ D jiji  Dijji  i Diiji  Diiij  D jjij  D jjji °
© wX j wX i ¹ wX i wX j °
æij ¾ D  ^` , (3.71)
Dijij  2 Dijji  D jiji °
°
¿

where the various terms of the stiffness matrix D are all taken in the global system.

3.10 Numerical examples

In this section, two numerical examples are presented: the first of these is aimed at
demonstrating the basic capability of the proposed continuum mechanics to deal with
arbitrary orientations within a global coordinate system. Thus its capacity to produce
conventional results is verified, but also the application of the extended Mohr plot is
shown as well as the additional information resulting from the orthotropic tensors. In
the second example, a case study is presented that is known to cause numerical prob-
lems including a type of in-plane shear locking. This seeks to demonstrate the capability
of the strongly orthotropic approach in achieving the most adaptive and accurate solu-
tion when compared to the ILST and an analytical formulation for an isotropic, infi-
nitely orthotropic and strongly orthotropic problem. The latter of these is also evaluated
using a number of representative commercial solvers.

3.10.1 Example 2: Satisfying Euclidean objectivity and the principle of ma-


terial frame-indifference
The principle of Material Frame-Indifference (or MFI), proposed by Truesdell and Noll
[136], states that “Constitutive equations must be invariant under changes of frame of
reference” of a body defined by

x F X ,t , V V X ,t (3.72)

such that it satisfies the motion


Chapter 3 Strongly orthotropic continuum mechanics 50

x* F * X , t * c t  Q t F X , t , ½
T °
°
V* V * X , t * Q t V X , t Q t , ¾ (3.73)
t* t  a, °
°¿

where Q(I) is a rotation matrix changing angle at I(t), c(t) is the relative position of the
origin of frame  X , Y from the origin of frame  * X *, Y * at time t, and a is some

arbitrary offset between time scales. The suggestion of an new constitutive law, as per
Section 3.7, demands that special care be taken in assuring MFI is maintained if we
heed the remarks of Murdoch [226]:

“Finally, to obtain the consequences of frame-indifference for a specific theory of mate-


rial behaviour it is necessary to identify intrinsic quantities. Care may be required in this
respect, as indicated by the atomistic considerations of Müller [227] and Murdoch [228].”

Y F X,t
:
, t *
Y* F * X
:*
X*
x
:0
X x* I :0 at t t0 0
X
: at t
c(t0)
:* at t* t  a

Figure 3.5 Example problem for satisfaction of the principle of MFI, a square element undergoes a mo-
tion F, which is also shown from a secondary observation frame as the motion F*

For this example, a square element, with an initial domain :0 as shown in Figure
3.5, undergoes a trellis deformation and a translation. This motion, described as F, is
also observed as F* from an alternate frame that is rotated and in relative motion. A
unidirectional material was analysed with a fibre modulus of 300
   
modulus of 3 GPa and Poisson ratios of 0.3 for both. Using the Rule of Mixtures [229]
the extended orthotropic stiffness matrix in Equation (3.56) and a rotation to the global
coordinate system using Equation (3.30) for an angle of  


anisotropic stiffness matrix was obtained:
Chapter 3 Strongly orthotropic continuum mechanics 51

ª 1.4901 0.0232 0.1861 0.0429 º


« 0.0232 0.0791 0.0029 0.1461»»
^`
D0 « u 105
« 0.1861 0.0029 0.0829 0.0054 »
« »
¬ 0.0429 0.1461 0.0054 1.1519 ¼

With only the knowledge of this anisotropic stiffness matrix, the strain condition from
:0 to : was calculated using the ALST in Equation (3.69). The stress V was found us-
^`
ing the stiffness matrix D0 . An alternate domain :* was then created to represent the

same deformation state as viewed from an alternative frame of reference whereby a


relative translation function of c occurred over the period t and relative rotation between
the frames was I. This is summarised by

c 'x, 'y 200, 50


ªcos I  sin I º (3.74)
Q I « sin I , I t 15q,
¬ cos I »¼

Subsequently, the strain condition for :0 to :* was determined using the ALST allow-
^`
ing the calculation of the stress V* using D* , which was obtained via the anisotropic
transformation functions in Equations (3.67) and (3.68). Finally, the transformed motion
and stress were ascertained according to the requirements of the MFI as set out in Equa-
tion (3.73), this is referred to as the equivalent-domain.

Table II. Strain and stress values numerating the satisfaction of the principle of MFI

x1 V (MPa)
Domain ---
x y x y xy von Mises
:0 210.0 30.0 0 0 0 0
: 275.0 190.0 2.144E+1 2.960E+1 1.095E+2 1.948E+2
:* 514.8 62.35 7.276E+1 8.092E+1 8.206E+1 1.948E +2

Table II shows the values for the point vector to the basis element 1, denoted by the
symbol ‘ in Figure 3.5, for each of the three domains. Additionally, the stress tensor
and von Mises is shown for each domain as calculated directly. To then show satisfac-
tion of MFI, the results for the domain : were applied in combination with the frame
Chapter 3 Strongly orthotropic continuum mechanics 52

relations in Equation (3.74) to the required-to-satisfy equations set out in Equation


(3.73), as follows:

­ 200 ½ ªcos 15  sin 15 º ­275½


F * X , t * ® ¾ « »® ¾
¯50 ¿ ¬ sin 15 cos 15 ¼ ¯190 ¿
>514.8, 62.35@
T

ªcos 15  sin 15 º ª 21.44 109.5 º ª cos 15 sin 15 º
V * X , t * « »« »« »
¬ sin 15 cos 15 ¼ ¬109.5 29.60 ¼ ¬  sin 15 cos 15 ¼
ª 72.76 82.06 º
« 82.06 80.92 »
¬ ¼

These equivalent-domain values for the 1 point vector and stress tensor precisely match
the values calculated for the alternate domain :* as shown in Table II.

4
(%)

x 10
strain, œNxyxy,,œNyxyx(%)

Shear stress, V xy, V yx

1
1 0.5
-7.5°, ILST
0 -7.5°, OLST
Shear strain,

0
-22.5°, "
-0.5

-1 -1
-1 0 1 -1 0 1
Axial strain, Nxx
strain, œ xx, N
œ (%)
yy (%)
yy Axial stress, V ii 4
x 10

Figure 3.6. Strain and stress plots for as observed from the alternate frames. Consistency to the extended
Mohr plot for strain indicates Euclidean objectivity, while the stress plot demonstrates the frame indiffer-
ence of the constitutive model.

Furthermore, Figure 3.6 shows the consistency of the ALST to the modified Mohr plot
for strain, and the corresponding stress plot generated using the extended anisotropic
stiffness matrix. The result indicates the objectivity of the formulations where the mate-
rial constitutive model is indifferent to the frame in which it is observed.
Chapter 3 Strongly orthotropic continuum mechanics 53

3.11 Chapter summary

By asserting that multiple geometrically compatible deformation modes can exist, and
that these deformation modes have strain fields made determinate by a dependency on
material properties, a novel linear strain tensor has been proposed. This tensor offers the
only general capability for describing exact kinematic behaviour on acceptance of the
previous assertion. Additional to this, the necessary formulations for practical analysis
and implementation have been developed so as to be sufficiently comprehensive, and
thus the proposed linear continuum mechanics basis is offered. Having only addressed
infinitesimal deformations, the consequence of finite motion must next be considered.
Chapter 4
SOCM plane shell finite elements
In this chapter, the application of the finite element method in two-dimensional elastic-
ity using the proposed strongly orthotropic continuum mechanics basis is presented.
Formulation for general plate and shell elements is not explored here; however, the
plane shell formulations given are considered a necessary primer for the development of
both. These elements reflect persistent material principal orthogonality, and so are used
to introduce the method of material principal corotation.

4.1 Finite element analysis of plane shells

4.1.1 Finite element treatment for a Tri3 element using the ILST
To implement the ILST into the finite element method, the strain of a master element in
the natural coordinate system & is defined: for a Tri3 element, this becomes

T
^& ` ª wu wu wv wv º
e « w[ , wK , w[ , wK » . (4.1)
¬ ¼

The strain is directly related to the nodal displacement vector q

> q1 , q2 , q3 , q4 , q5 , q6 @
T
q (4.2)

by a shape function matrix denoted as G:

^& `
e Gq (4.3)
54
Chapter 4 SOCM plane shell finite elements 55

The natural domain strain vector can then be mapped to the global domain state vector
using the conventional mapping matrix A :

^`
A u ^ `e
T &
ª¬exx , eyy , exy , Z z º¼ (4.4)

where, given the second-order susceptibility of normal strain to rotation, the rotation
tensor is appended such that the strain state becomes unique. Using the Jacobian matrix

ª wx wy º
« w[ w[ » 1 ª J 22  J12 º
J « » where J 1 , (4.5)
« wx wy » det J «¬  J 21 J11 »¼
« wK wK »¼
¬

and by considering the ILST and rotation tensor as in Equations (3.7) and (3.10) respec-
tively, the mapping becomes

ª 2 J 22 2 J12 0 0 º ª1 0 0 0 1 0 º
« 2 J 21 2 J11 »» «0 0 1 0 »»
1 « 0 0
« 0 1
A , where G (4.6)
2 det J «  J 21 J11 J 22  J12 » «0 1 0 0 0 1»
« » « »
¬ J 21  J11 J 22  J12 ¼ ¬0 0 0 1 0 1¼

such that

B AG , and V e DBq. (4.7)

Here, the definition of an A and G matrix is extraneous; however, their utility becomes
apparent when the alternative formulation for A is given in Section 4.3.

4.1.2 Finite element treatment for a Quad4 element using the ILST
Implementation of the ILST for a Quad4 element requires only few changes to the for-
mulation outlined in the previous subsection. The displacement vector q, rather than
Equation (4.2), becomes

> q1 , q2 , q3 , q4 , q5 , q6 , q7 , q8 @
T
q , (4.8)

while the shape function matrix in Equation (4.6) is replaced by


Chapter 4 SOCM plane shell finite elements 56

ªK  1 0 1 K 0 1 K 0 1  K 0 º
«[  1 0 1  [ 0 1 [ 0 1 [ 0 »»

G , (4.9)
4 « 0 K 1 0 1 K 0 1 K 0 1  K »
« »
¬ 0 [ 1 0 1  [ 0 1 [ 0 1 [ ¼

where K and [ are as usually defined [230]. This allows use of the linear state tensors in
the finite element method for linear static analysis.

4.2 Premise: persistent material principal orthogonality

Where Chapter 3 presented a continuum theory based entirely on an assertion of geo-


metric strain indeterminacy, the contributory component to this chapter relies on an ad-
ditional premise: the property of persistent material principal orthogonality. Figure 3.3
formerly demonstrated the asserted association between material properties and de-
formed material orientation, the concept is now furthered to suggest that even under a
finite deformation the principal material properties will remain orthogonal to the new
orientation—and in a persistent manner.

œ22 œ22
Y

y
^ ,%`

^%`

^%`

^`

^`

G
œ12 \z

^ ,%` ^%` ^%` ^` ^`


X 1 1 x x
Global reference Fibre-normal Orthogonal Locking Eulerian/
Skew shear
frame compaction material frame C-R frame

Figure 4.1. A persistently orthogonal material frame compared to a locking Eulerian frame—identical to
the corotational frame in this case.

This is in contrast to the predominant three conventional frames, all of which see the
case depicted in Figure 4.1 as being a symmetric and rotation-free motion. This includes
the Eulerian, Lagrangian and corotational frame. In such an example, the orthogonal
measures will numerically lock as the unseen rotation of the horizontal fibre will result
in a necessary fibre axial strain for compatibility. The Lagrangian frame avoids this by
suggesting obtuse material principality, and therefore does not satisfy the premise of
orthogonality. Thus for large rotations we see that this approach is consistent with the
Chapter 4 SOCM plane shell finite elements 57

observation of Figure 3.3. In the subsequent sections, the formulations and method are
presented that allow solution of the problems thus far in concert with the strongly
orthotropic linear theory of the previous chapter.

4.3 Finite element treatment for a Tri3 and Quad4 element using
the OLST

In order to formulate an element using the global domain state vector required by the
proposed continuum mechanics basis we must arrive at the same mapping as Equation
(4.6). A concatenation of the OLST and the material principal rotation tensor of Equa-
tions (3.39) and (3.46) respectively, Equation (4.4) is replaced by

^`
A u ^ ` e.
T &
ª¬œxx , œ yy , œxy , œ yx ,\ z º¼ (4.10)

By defining the distribution components of the said tensors as

rija d ij sin 2 Tij  d ji cos 2 Tij and (4.11)

dij  d ji
rijb sin 2Tij , (4.12)
2

the strain and rotation tensors can be expressed in the simplified form

§ wu wu j · a § wu wu j · b
N ij ¨¨ i  ¸¸ rij  ¨¨ i  ¸¸ rij and (4.13)
© wX j wX i ¹ © wX i wX j ¹

wui a wu j a § wu wu j · b
\ iu j rji  rij ¨ i  ¸¸ rij . (4.14)
wX j wX i ¨ wX wX
© i j ¹

By use of this notation, the new mapping matrix A, replacing A in Equation (4.6), is
applied in a linear fashion so as to yield Equation (4.10). In the planar case, it is written
as follows:
Chapter 4 SOCM plane shell finite elements 58

ª 2rxxa J 22 2rxxa J12 0 0 º


« »
0 0 2ryya J 21 2ryya J11
1 «« a »
A rxy J 21  rxyb J 22 rxya J11  rxyb J12 rxya J 22  rxyb J 21 rxya J12  rxyb J11 » (4.15)
det J « a »
« ryx J 21  ryx J 22 ryxa J11  ryxb J12 ryxa J 22  ryxb J 21 ryxa J12  ryxb J11 »
b

« ra J  rb J ryxa J11  ryxb J12 rxya J 22  rxyb J 21 rxya J12  rxyb J11 »¼
¬ yx 21 yx 22

Considering the intermediary natural state (Equation (4.3)) need not be determined, and
using

B AG , (4.16)

the required strain/rotation tensor quantities can be related directly to the displacement
field:

^`
B u ^ `q
T 
ª¬œxx , œ yy , œxy , œ yx ,\ z º¼ (4.17)

For Tri3 elements, the shape function matrix from Equation (4.6) and the nodal dis-
placement vector from Equation (4.2) are used; whereas, for a Quad4 element, Equa-
tions (4.8) and (4.9) are used respectively. The element stiffness matrix K can then be
determined by integration over the spatial domain of the master element in the usual
manner;

B T ^ ` DB det Jd [ dK ,
1 1
tn ³ ³

K (4.18)
1 1

where tn denotes the plate or shell thickness through the normal.


Few additional changes are required for practical implementation into the well es-
tablished finite element method, though the additional stress and strain terms must be
carried throughout the computational procedure and the orthotropic von Mises formula-
tion of Equation (3.65) should be used. The finite element formulation outlined here
was used to generate the numerical results in the section following.

4.4 Material principal corotation

The proposed continuum mechanics basis requires that directional properties of


orthotropic media are treated as being persistently orthogonal and Euclidean. This sec-
Chapter 4 SOCM plane shell finite elements 59

tion presents the corotational procedure required to ensure this response from continua
that model multi-axial materials.

4.4.1 Directional decomposition of multi-axial materials


To maintain an orthogonal coordinate system, interwoven materials and the like must
have each individual fibre family spatially decomposed into distinct coincident continua
due to the angular change between principal fibre directions. The term coincident is
considered more appropriate than coexistent as the latter refers to the temporal rather
than spatial domain. Due to the determinacy added by the OLST, each unique contin-
uum gains its own principal rotation.

2
2
Y

^`
^`

Undeformed
multiaxial +
material

deformation
X ^` ^`
1 1

Local
^`
œ22
2

2
^`

^`

^`
œ11 ^`
œ12
+

Body rotation
^` ^` ^`
œ21 1 1
y

^`
Effective \z
deformed state +

x ^` x x
\z

Figure 4.2. Example of continuum decomposition of a plain weave composite with an infinite orthotropic
ratio. This is the basis for the material principal corotation of any multi-axial material.

An example of this decomposition is shown in Figure 4.2 for a biaxial woven com-
posite, where the fibre families are each considered in their own local material coordi-
nate system,  and  respectively, and are treated as being infinitely orthotropic for

conceptual simplicity. A bidirectional interwoven material is normally represented by a


single orthotropic continuum, or by a single anisotropic continuum for any oblique ori-
entation. Because woven lamina comprise of fibre families existing at the same thick-
ness interval, laminated composite elements are typically not used at this intra-ply level.
Chapter 4 SOCM plane shell finite elements 60

This is because such elements integrate through the thickness of the materials, and
would result in an erroneous bending response.
Additionally, a weakly coupled torsion coefficient can be concatenated to Equation
(3.21) adding energy due to the interaction between coincident continua. A strong cou-
pling is also possible and would represent a micro-polar moment. This approach has
been described as a fictitious spring constant for a drilling DOF [231], although it does
have real physical meaning. For example, the resistance between relatively rotating fi-
bre bundles in a woven composite, and furthermore the interaction as they reach what is
termed the shear locking angle [51], which is a type of physical rather than numerical
locking.

4.4.2 Corotational procedure


The method of element corotation is widely used and can be reasonably adapted to co-
incident continua providing that the material dependent implied rotations are dealt with
correctly. The corotational procedure has been comprehensively reported [124, 133,
232], and we do not attempt to reproduce the theory here. There are, however, a number
of special considerations in the algorithmic implementation of what is described here as
material principal corotation that we shall briefly regard.
In the simplest case, material principal corotation can be represented by

Xcr QXT , (4.19)

where the rotation matrix is based on the material principal rotation tensor of Equation
(3.46) rather than the isotropic (geometric only) rotation of Equation (3.10) in order to
instead reflect the proposed internal kinematics. A set of nodal transformations associ-
ated to a number of coincident elements may represent continua of varying principal
orientations, an example of which is shown in Figure 4.2. This infers that multiple coin-
cident continua will require multiple simultaneous coordinates for each node based on
their element association. We refer to these as shadow nodes with coordinates deter-
mined according to their respective shadow transformations. Thus for any given mate-
rial principal rotation tensor, \ z , the shadow transformation will be Q \ z according

to the rotation matrix given in Equation (3.74).


Chapter 4 SOCM plane shell finite elements 61

^` :
0-cr
3
4 3 4

^ ,`
:0
^`
:0-cr
2
1 2 1

a) Domains at initial solution b) Pre-counter-rotated shadow nodes

Figure 4.3. Shadow nodes for the material principal corotation for the example of trellis shear. a) The
initial undeformed domain; and, b) In the first iteration, each of the nodes has two shadow coordinates.

An example of the pre-counter-rotation of nodes associated with differently oriented


fibre families is shown in Figure 4.3. A similar process to the pre-emptive coordinate
change in Equation (4.19) can be repeated to recalculate the boundary conditions. If the
boundary displacements are fixed, as in the trellis shear example, exact convergence
will be achieved in the first iteration. Particular care must be taken in the algorithmic
implementation as nodal properties and conditions must be stored according to each
node–element association as opposed to storing only unique nodal data and addressing
it via connectivity lists. This can present significant difficulty in application to existing
solver codes depending on the architecture of the integration and other routines. Due to
certain simplicities afforded by the determinacy of the trellis condition, a moderate form
of the procedure has been used to solve the later-demonstrated numerical example in
Subsection 4.6.3.

4.4.3 Equivalence between combined and coincident continua


For the necessity of comparison, we require equivalent properties to a conventional sys-
tem. This can be found by equating the required single biaxial element with two coinci-
dent elements with half the thickness each and of perpendicularly assigned properties.
The resulting functions for an equivalent combined material matrix Dc are

2Q 12Q 21
Qc , (4.20)
Q 12 Q 21

E1 1 Q 21  E 2 1 Q 12
E c 1 Q c and (4.21)
2 1 Q 12Q 21
Chapter 4 SOCM plane shell finite elements 62

2 E1 E 2
G c , (4.22)
E1 1 Q 21  E 2 1 Q 12

which make the necessary components as follows:

ª E Q c E c 0 º
1 « c »
Dc «Q c E c E c 0 » (4.23)
1 Q c 2 « »
¬« 0 0 G c 1 Q c 2 ¼»

A similar approach can be taken to derive the combination equations for lamina com-
prised of unequal thickness or unequal properties in each fibre family.

4.5 Some other considerations

4.5.1 Fibre pinning versus fibre slippage


In the literature review we surveyed a number of investigations to do with fibre slippage
in woven and stitched-through biaxial composites. In these, the significance of inci-
dence of relative motion between tow crossover points and how that could positively
affect shape compatibility of pre-forms were assessed. While it is not a necessary con-
sideration to structural strongly orthotropic materials due to typical design compliance,
in the instances of forming or highly flexible part analysis the admissibility of the slip-
page condition is a necessary consideration.
We regard this issue in particular because of the earlier suggestion of node sharing
between coincident elements, whereby nodal points that, in a Lagrangian mesh, are
strictly associated with material points such that they would effectively ‘pin’ the con-
tinua together. This is representative of one of the constraints of the IFRM. It is only to
be said that caution must be exercised in applications where admissibility of the slipping
condition is required, though suggestion can be made *. However, that said, we have ear-
lier observed that the literature supports use of multiple shell elements with contact
conditions for modelling multiple plies in preforms required to undergo large deforma-
tion. The extension to that technique would suggest that the intra-ply fibre-slippage
*
The Arbitrary Eulerian–Lagrangian (ALE) method provides the ability to model shell structure without the pervasive spatial con-
nection between material points and nodal coordinates.
Chapter 4 SOCM plane shell finite elements 63

condition can be provided by use of multiple coincident shells with discrete nodes and
symmetric outwardly penalising penetration conditions.

4.5.2 Coupled drilling moments and micropolarity


The earlier reviewed proposal of Yeh and Chen [190] utilised extraneous degrees of
freedom in shell elements in order to allow a micropolar drilling moment and rotation to
shell elements. This was shown in particular because it affords us the possibility of ap-
plying a micropolar or weak rotational coupling between elements.
The latter case can be used to apply a condition analogous to physical phenomena
such as angular fibre locking functions. This can be achieved by concatenating an intra-
laminar rotation to the transformation matrix:

ªT 0 º
Tc « 0 1» (4.24)
¬ ¼

where T is given by Equation (3.28); and

^%` ª ^%` D 0 º ^` ª ^` D 0 º


Dc « », Dc « » (4.25)
¬ 0 Dr 3r 3 ¼ ¬ 0 Drzrz ¼

^%` ^`
where D and D follow on from Equation (3.21). Last, the new material matrix
transformation functions are as given in Equation (3.30) with the additions

^` ^`
Dc 1: 4,1: 4 D,
^` ^`
Dc 1: 4,5 Dc 5,1: 4 0, and (4.26)
^`
Dc 5,5 1.

For the former case, the possibility of sustaining a micropolar moment exists as the
vectorised strain carries the additional terms. This offers the function of a differential
between in-plane strains and stresses wherein descriptors such as the wryness tensor in
Cosserat theory are not required.
Chapter 4 SOCM plane shell finite elements 64

4.6 Numerical examples

4.6.1 Example 1: Patch test for a linear multi-axial problem


In this numerical example, we address simple but necessary results to demonstrate the
validity of the proposed approach. The problem models the linear solution to a two-
dimensional square region that is subjected to a trellis boundary displacement. The key
features to be considered are:

o Implementation into the finite element method;

o Use of multiple elements;

o Patch test satisfaction between uniform and arbitrary meshes;

o Ability of the strongly orthotropic formulation to agree with accepted solution; and,

o Agreement between solutions of single biaxial elements and dual coincident elements.

We consider the features of two mesh layouts for each of the two continuum mechanics
approaches; that is, conventional biaxial ILST form and proposed OLST coincident
continua.

a) b)

9 7 11.035
7 8 8
Boundary: 9
11.035
*(dashed)o*, 4
subject to a trellis 4 5 6 5 11.035
motion F & &,t 6
11.035
1
1 2 3 2
3 11.035

Figure 4.4. a) and b) Different patch test meshes for a square plate (dashed) being deformed by a trellis
shear boundary condition. A uniform von Mises plot results for each mesh for the material family . The

contour bar indicates a uniform von Mises stress in MPa.

The isoparametric and arbitrary meshes can be seen in Figure 4.4 as a) and b) respec-
tively, each shown in their deformed state with boundary *. Mesh 2 is essentially the
same as Mesh 1 but with the centre element rotated. This allows the need for material
aligned meshing to be investigated for each computational mechanics approach, which
Chapter 4 SOCM plane shell finite elements 65

is a common problem for strongly orthotropic shear deformation. The initial geometry,
*0, can be seen by the dotted line, which is prior to the trellis motion F, whereas the de-
formed geometry, *, is given by a solid line.

Table III. Material properties for dual coincident elements (horizontal and vertical) and for equivalent
biaxial elements (combined). Moduli are measured in GPa.

Layer Direction Elastic Modulus Plane Poisson Ratio Shear Modulus

1 1.800E  12 4.999E  1.200E+2


Vertical fiber family: 
2 1.800E+2 21 4.999E  1.799E-1

1 1.800E+2 12 4.999E  1.799E 


Horizontal fiber family: 
2 1.800E  21 4.999E  1.200E+2

Biaxial model Combined 9.011E+1 Combined 9.988E  0.3593E-1

Table III lists the material properties used in the example, where Equation (3.56) de-
termines the response matrix for the proposed solution. The properties were substituted
into Equations (4.20) to (4.22) and constructed the elastic response matrix of Equation
(4.23). Following the case of uncured composite materials, properties are assigned that
reflect this material, where near incompressibility is used, and an orthotropic ratio of
1000 to 1.

Table IV. Results for spatially decomposed solution and conventional result.

Normal Shear von Mises


Layer
Dir. Strain Stress Plane Strain Stress Sub-lam. Equiv.

1 5.051E  9.099E+0 12 2.007E  3.611


Horizontal 11.04
2 5.051E  1.364E  21 3.009E  3.611
7.738
1 5.051E  1.364E  12 2.007E  3.611
Vertical 11.04
2 5.051E  9.099E+0 21 3.009E  3.611

Biaxial Comb. 5.051E  4.556E+0 Comb. 2.010E  3.611 NA 7.738

The output results are shown in Table IV; as the results for Mesh 1 and 2 were iden-
tical, only results between the two continuum models are given. Subsequent to Equation
(4.21), the average of the normal stresses between the horizontal and vertical layers is
equal to the value for the biaxial model in each direction. Importantly, the shear strains,
Chapter 4 SOCM plane shell finite elements 66

œ12 and œ21, for each layer are distributed according to the orthotropic balance ratio of
Equation (3.40), as in this case the aspectual strain is zero. Furthermore, the sum of
these components is equal to that of the conventional biaxial model. It can also be seen
that each of the component shear stresses is identical according to the necessary condi-
tion of equilibrium, and that these components are also equal to the conventional model.
Last, we note that the von Mises stress for each layer, by the sub-laminar decomposition,
is not equal to the biaxial model; however the equivalent von Mises, where the stresses
are averaged between the layers, is as predicted by the biaxial model.
The results indicate the ability of the proposed method to be identical to a conven-
tional biaxial orthotropic model in the simple linear elastic case. In doing so—and with-
out additional information—the decomposed and strongly orthotropic method provides
a more informative analysis.

4.6.2 Example 3: Arbitrary element under arbitrary orientation


A plane stress element of arbitrary geometry was set in four arbitrary rotational orienta-
tions, each having the equivalent local boundary displacement condition. This was done
to confirm that a number of fundamental behavioural expectations are met, including
the ability of the formulations to predict linear stress results exactly equivalent to those
by the ILST. The example used material properties whereby the principal elastic
modulus was 180GPa, the secondary modulus was 180MPa according to the
orthotropic ratio 1.0E3 and the major Poisson ratio was 0.4999. The shear moduli, G12

and G 21 , and the local material stiffness matrix were found according to the extended
orthotropic stiffness matrix in Equation (3.56). The properties for a conventional
benchmark solution were calculated using the degeneration of Equation (3.58). Further
expectations included the ability of the formulations to produce consistent strain tensors
via the OLST in Equation (3.39) and the developed coordinate system transformation
functions of Equation (3.30). This was validated using the invariant stress measure of
von Mises, using the orthotropic formulation of Equation (3.65), and the extended Mohr
plot according to Figure 3.4. Similarly, consistency to the conventional method was ex-
pected as per the Eigen solutions of Equations (3.47) and (3.48).
Chapter 4 SOCM plane shell finite elements 67

The analysis used the Quad4 element outlined in Section 4.3 with four arbitrarily se-
lected nodal coordinates. For nodes 1 through 4 these were [0.358, 0.483]T, [ !#
0.633]T, [ !$ !&!#< T and [0.581, !#< T respectively. All degrees of freedom
were constrained, with displacements of [0.001, 0.005]T applied to node 1. For the sub-
sequent rotated systems, all coordinates and displacements were transformed accord-
ingly.
Chapter 4 SOCM plane shell finite elements 68

strain for 0° is the average of the unique 0° OLST values; although each produces the
same shear stress verifying the relationship between the OLST and the extended
orthotropic compliance tensor. When Equation (3.46) was applied as in Equation (4.10),
the material principal rotation \ z was determined to be 5.610E #
>    a-
tional orientations at the first integration point: this is representative of the intermediary
solution between geometric rigid rotation and the infinitely orthotropic case represented
by Figure 3.2. In comparison, the conventional Equation (3.9) generates a rotation of
2.397E #
>@>\^ _ @ >  \@` _ ation. The
latter result is regarded as meaningless according to the assertion of strongly orthotropic
behaviour, to this we draw analogy to observations made by Day and Potts [233] on iso-
tropic Mindlin beams and shear locking in a similar issue:

“… the parameter chosen as the rotation degree of freedom is dependent on the geometry
and orientation of the elements, and its physical meaning is unclear.”

Nonetheless, the value can be obtained by the proposed continuum mechanics basis if
G12 G 21 is specified.
The added information gained through the non-orthonormal Eigen analysis and re-
coverability of the conventional analysis is demonstrated by first solving the principal
and pseudo principal strains: respectively equivalent to

ª 1.124E  1 6.435E  1º
O eig œ eig « » diag 1.103E  1, 4.147E  1 and
¬9.647E  4 4.126E  1¼

ª1 º ª 1.124E  1 3.222E  1º
O eig « œ  œT » eig « » diag 9.298E  2, 6.180E  1 .
¬2 ¼ ¬3.222E  4 4.126E  1¼

Using the extended Mohr plot, it is shown that the pseudo or conventional principal
strains, calculated from the degenerated tensor as per Equation (3.48), lie on the con-
ventional Mohr’s circle. At the same time, the principal strains solved from the OLST
as per Equation (3.47) add meaningful information to the analysis without any addi-
tional input. This is demonstrated in Figure 4.6, where the equivalence of results from
the OLST and Table V becomes apparent.
Chapter 4 SOCM plane shell finite elements 69

Shear strain œxy, œyx (%)


0.6 0°, ILST
0.4 0°, OLST
72°, "
0.2
188°, "
0 273°, "
0.2 Principal ILST
0.4 Principal OLST
0.5 0 0.5 1
Axial strain œxx, œyy (%)

Figure 4.6. Extended Mohr plot for strain states under various element rotation

In summary, it has been shown that the proposed continuum mechanics produces the
same solutions as the conventional approach for a linear problem. It is clear that addi-
tional meaningful information regarding the strain state and rotation of the continuum
can be obtained without any increase above the minimum initial problem specification.

4.6.3 Example 2: In-plane shear locking/softening of a biaxial material


Control of a picture frame rig, which exerts trellis shear in the specimen, is displace-
ment between diagonal hinges. The material, based on the previous numerical example,
now uses two perpendicular layers that are coincident as in Figure 4.2. Three example
problems were solved to demonstrate the adaptive performance of the ILST, an analyti-
cal solution and the proposed OLST for three different orthotropic ratios. The first ma-
terial was isotropic using an elastic modulus of 180GPa; the second example used an
infinite orthotropic ratio, and hence the secondary principal elastic modulus was 0GPa;
and a final example used a strongly orthotropic material, where the elastic moduli were
identical to the previous numerical example: 180GPa and 180MPa. In the last example,
the results from three commercial FEA programs were also evaluated.
Chapter 4 SOCM plane shell finite elements 70

a) b)
G 3
4 3 4
: ^`
:0 ^` :
0-cr
:0
y x
y x 2
1 2 1
F(X,t) Conventional Material principal
or first run co-rotated

Figure 4.7. a) Trellis shear element within a picture frame test; b) The single element mesh used, exag-
gerated deformation (dashed) and the co-rotated (·)cr initial configuration of the infinitely orthotropic ma-
terial 

By assuming a uniform strain distribution, use of an analytical formulation that is ex-


plicit to the displacement control is allowed. Figure 4.7 depicts the basic mechanics of a
picture frame shear test where, in this case, a trellis shear element is oriented parallel to
the frame sides; it also depicts the mesh details for the domain in the initial configura-
tion :0 and an example material principal corotation for the infinitely orthotropic mate-

rial with horizontal alignment (layer ). Note that in each case the initial coordinates for

node 1 were [0, 0]T and for node 3 were [100, 100]T. Given that displacement G is a con-
trolled input depending on time t and resulting in a trellis motion F(X,t), then the fibre
angle M can be directly related to displacement:

ªG º
M cos 1 «  cos M0 » , (4.27)
¬ 2L ¼

where M0 is the undeformed fibre angle—in this case 2M0 = 90°. The displacement ap-

plied in all cases was a 1% diagonal extension, such that G 2 ; according to the pic-
ture frame motion, all boundary constraints are determined by this single displacement
value.
For comparison, a solution using the ILST with combined laminate properties as per
Equation (4.23) was found in each case. Additionally, an analytical model was used
with the effective modality of Figure 3.2. For this solution, layers  and  were each

considered in their own local axis and each represented one of the perpendicular fibre
Chapter 4 SOCM plane shell finite elements 71

families with material principal directions of y and x respectively. According to work by


Spencer [29], the internal strains are kinematically determinate under trellis shear: as
part of an infinitely orthotropic or idealised fibre-reinforced model, local axial fibre
strain, or material principal strain, is zero; and the local fibre-normal strain is related to
fibre angle. Thus

^` ^` ^` ^`


e11 e22 0, e11 e22 sin 2M  1, (4.28)

with the shear strains being

^` ^` 1
e12 e12 sin 2 E 0  2 E . (4.29)
2

Given the small displacements being applied, any difference between finite and infini-
tesimal shear strains is negated here. Finally, the proposed OLST was used for an
evaluation solution.

Table VI. Comparison of analytical and continuum solutions for isotropic picture frame shear: only the
ILST and OLST are kinematically exact, resulting in correct stress values

Strain (%) Stress (MPa)


Model
x y xy yx x y xy

ILST 5.051E 3 5.051E 3 1.005 — 1.818E+1 1.818E+1 1.206E+3

 2.020E 2 0 1.005 — 4.848E+1 2.423E+1 1.206E+3


Analytical
 |0 2.020E 2 1.005 — 2.423E+1 4.848E+2 1.206E+3

 5.051E 3 5.051E 3 1.005 1.005 1.818E+1 1.818E+1 1.206E+3


Proposed OLST
 5.051E 3 5.051E 3 1.005 1.005 1.818E+1 1.818E+1 1.206E+3

Table VI shows the results using the isotropic material properties. The ILST pro-
duces the exact solution given that there is no rigid body rotation for the trellis motion
as asserted in Figure 3.3. Similarly the validity of the OLST is shown as it returned
identical results for each layer, which were also identical to the exact solution—in this
case the ILST. In contrast, the analytical solution demonstrated its inability to behave
symmetrically given that it is based on a fibre-reinforced model. Note that for compari-
son of stress invariants, the average stress tensor can be used, where
Chapter 4 SOCM plane shell finite elements 72

{avg}
V
2

1 { }
V  { } V . (4.30)

This is equivalent to a single biaxial element as represented by Equation (4.23).


For the second example—the infinitely orthotropic case—the analytical method was
used to generate the exact solution given that zero fibre-extension is assured under infi-
nitely orthotropic trellis motion (see Figure 3.2). A corotational procedure was used for
the OLST to eliminate the rigid body rotation according to the material principal rota-
tion tensor. The conventional linear rotation tensor, however, indicated no rigid body
rotation and so the ILST procedure did not require corotation. Note that were the motion
instead to have been specified according to Figure 3.2 then the ILST would have re-
quired corotation where the OLST would not. As with the other iterative solvers pre-
sented in the final example, the linear first run for the OLST returned an overly stiff so-
lution, but also returned the distinct material principal rotation tensor. A subsequent run
used a preliminary implementation of material principal corotation as described in Sec-
tion 4.4, where essentially each fibre family was pre-counter-rotated by the returned ro-
tation values.

Table VII. Comparison of analytical and continuum solutions for an infinitely orthotropic material under-
going energy-free picture frame shear: only the analytical solution and the OLST are kinematically exact

Strain (%) Stress (MPa)


Model
x y xy yx x y xy

ILST 5.051E 3 5.051E 3 1.005 — 4.546 4.546 0

 2.020E 2 0 1.005 — |0 |0~||| 0


Analytical
 |0 2.020E 2 1.005 — |0 |0~||| 0

 2.020E 2 0 0~||| 2.010 |0 |0~||| 0


Proposed OLST
 |0 2.020E 2 2.010 0~||| |0 |0~||| 0

Table VII presents the results for the pure trellis motion resulting from an infinite
orthotropic ratio. This is an energy-free motion, and hence the ILST demonstrates its
invalidity due to non-zero axial stresses. The OLST, however, was able to adapt to the
trellis kinematics based on material properties alone as is evident when compared to the
exact solution provided by the analytical method.
Chapter 4 SOCM plane shell finite elements 73

In the final numerical example, where strongly orthotropic material properties were
used for each lamina, three additional solutions were found using commercial FEA pro-
grams. The NASTRAN code SOL101 was used to ascertain the fourth result, and al-
though susceptible to second-order rigid rotation, the example problem had no such
geometric rotation. Here, the equivalent combined continuum properties were as for the
ILST. Using nonlinear geometry in the ANSYS code, the fifth solution was indifferent
to rigid body rotation, including second-order components, by way of Polar Decomposi-
tion [136]: where the deformation gradient tensor F is decomposed into a rotation tensor
K
R and the left stretch tensor U such that it satisfies F = RU, where F ’u  I . The
nonlinear solver Marc was used to perform the sixth solution using the small strain and
total Lagrangian formulation.

Table VIII. Comparison of analytical and continuum solutions for strongly orthotropic picture frame
shear: the OLST provides the only intermediary solution and is kinematically exact

Strain (%) Stress (MPa)


Model
x y xy yx x y xy

ILST 5.051E 3 5.051E 3 1.005 — 4.556 4.556 3.611

 2.020E 2 0 1.005 — 1.818E 2 3.637E 2 3.611


Analytical
 |0 2.020E 2 1.005 — 3.637E 2 1.818E 2 3.611

 2.014E 2 4.543E 8 3.009E 3 2.007 3.626E 2 1.821E 2 3.611


Proposed OLST
 4.543E 8 2.014E 2 2.007 3.009E 3 1.821E 2 3.626E 2 3.611

NASTRAN 101 5.051E 3 5.051E 3 1.005 — 4.556 4.556 3.611

ANSYS NLGeom 0 2.020E 2 1.005 — 1.812E 2 1.821E+1 3.611

Marc Total Lgrg. 0 0 1.005 — 0 0 3.611

A summary of the stress and strain results is presented in Table VIII showing the so-
lutions from six possible procedures for the strongly orthotropic problem. Here, the
proposed solution using the OLST and material principal rotation tensor was capable of
producing a kinematically intermediate solution as is shown by the fibre-axial and fibre-
normal strains. In contrast to this, the solution of the ILST is clearly incorrect as the re-
sult disregards non-isotropic material kinematics and hence does not model the defor-
mation mode required for strong orthotropy and fibre near-inextensibility; as with the
Chapter 4 SOCM plane shell finite elements 74

second example where it over estimates the stiffness. This erroneous response is known
as in-plane shear locking. Meanwhile, in normal strain, the infinitely orthotropic ana-
lytical model further demonstrates its failure in adapting to the intermediary case: the
entire planar compaction was directed to the fibre-normal axis. As there is currently no
other method providing an intermediary solution, and as the OLST is soundly derived,
we propound these values to be the exact solution.
Additionally, the NASTRAN SOL101 result replicates that of the ILST. The
ANSYS solution, whilst rotationally insensitive, has a rotational basis measured against
an arbitrary element edge or coordinate; it is evidentially mesh-dependent, whilst also
suffering from the locking problem of the ILST. In the Marc solution, by using material
principal stretches—which after shearing are no longer orthogonal—fibre inextensibil-
ity is reflected in the normal stresses producing a soft shearing result. Critically though,
the second-order fibre-normal compaction can no longer be seen in normal strain. Over-
all none of these conventional methods can reproduce the performance of the imple-
mentation of the OLST with material principal corotation. This result, considered exact,
identifies the fibre near-inextensibility, fibre-normal compaction, and the balance be-
tween these resulting from the finite orthotropic ratio—or intermediary case.
In shear strain, the results show all procedures produced the same total shear strain;
however, the presented continuum mechanics is more informative in identifying the
balance between planar shear values. The distribution between these is approximately
representative of the orthotropic ratio. In the results for shear stress, all values are equal:
this suggests the distribution of strains by the OLST is correct when used with the ex-
tended orthotropic compliance tensor in comparison to the conventional result using the
shear modulus determined using Equation (4.22).
In this example we applied a small yet finite distortion, which according to the
newly defined material-principle rotation also underwent a finite rotation that was cor-
respondingly co-rotated. The necessity for such a corotation is reaffirmed by the work
of Manach and Rio [157], who wrote:

“It has been observed that the use of convected material coordinates is essential to take
into account large elastic deformations with regards to a classical linear orthotropic ap-
proach.”
Chapter 4 SOCM plane shell finite elements 75

The results for the three example problems—using isotropic, infinitely orthotropic and
strongly orthotropic material properties—can be summarised by noting that kinematic
adaptivity and intermediacy was provided only by the OLST. Moreover, the only exact
solution for both the isotropic and infinitely orthotropic case resulted from use of the
proposed continuum mechanics basis.

4.7 Chapter summary

Based on the premise that the privileged directions of a material’s constitution remain
persistently orthogonal, and by use of the linear theory developed in Chapter 3, a treat-
ment of plane shell finite elements has been given. Under finite rotations, these ele-
ments were shown to benefit from the proposed method of material principal corotation
so as to mitigate existing instances of numerical locking. A linear and an iterative prob-
lem in the numerical examples were able to illustrate the ability of the method to arrive
at a consistently accurate solution—as compared to the specialised analytical and con-
ventional approaches—and thus a insusceptibility to the said instances of erroneous re-
sponse. Furthermore, the proposed continuum mechanics provided the only solution
with both a soft shearing response and the second-order strain component due to fibre-
normal compaction for the strongly orthotropic case; and thus demonstrated its capacity
to generalise the elastic behaviour of isotropic media, infinitely orthotropic media and
all intermediary cases.
Chapter 5
The intrinsic field tensors
By asserting that mechanistic strain metrics must be dependent on intrinsic material
properties and satisfy equilibrium, a novel orthotropic linear strain tensor was earlier
proposed. It is used in this chapter as a basis for a general orthotropic equivalent to the
nonlinear field tensors for deformation. This also leads to an interesting treatment of a
number of classical decompositions. A general case is thus made for the usage of the so-
called intrinsic field tensors, including demonstrative numerical examples.

5.1 Nonlinear field tensors in mechanics

5.1.1 Class definition: (solely) extrinsic field tensors


To define the class described as the Extrinsic Field Tensors we observe that Equation
(3.6) describes some tensor field, and is therefore itself a field tensor. The field over
which it refers is, as with all mechanical field tensors, a solely state or extrinsic vari-
able; it is independent of any of the material’s intrinsic mechanical properties, such as
those in the stiffness matrix D. To exemplify this we consider that some geometry un-
dergoing a fully constrained field motion F will, under a conventional strain metric,
have a strain field that is totally invariant to the type of material to which the geometry
is comprised.

76
Chapter 5 The intrinsic field tensors 77

5.1.2 Polar decomposition and the Stretch Tensors


The ILST, while having shear terms that are insensitive, has susceptibility in its normal
strains to small rotation. While this can be resolved using an iterative corotational pro-
cedure, as demonstrated in Subsection 4.6.3, it is classically dealt with using the method
of Polar Decomposition [8]. By this method the deformation gradient is separated into
its symmetric and anti-symmetric parts, though it is a multiplicative rather than additive
decomposition—as is used in the ILST and small rotation tensor—and thus accounts for
second-order effects. For this reason it directly resolves the said sensitivity in the nor-
mal components and is exact in shear even for large rotation, though as with the ILST
relies on an assumption of tensor symmetry to do so.
The symmetric part of the deformation gradient F can be isolated in a Lagrangian
frame, and is called the Right Stretch Tensor UE (denoted here as the isotropic right
stretch tensor. It is related to the isotropic finite rotation tensor RE as follows.

F R E UE (5.1)

where the notation of (·)E indicates it to be specifically the conventional nonlinear iso-
tropic form. Additionally, the Left Stretch Tensor VE—the geometric stretch as ob-
served from an Eulerian frame—is defined in

F VERE. (5.2)

It is notable that the left and right stretch tensors become identical under infinitesimal
deformation.
To solve these equations, the right stretch tensor is mathematically forced to have
symmetry by defining it to be the matrix square root of the always-symmetric FTF:

FTF U E 2
(5.3)

This is simply verified by substitution of Equation (5.1) into the previous equation.
Recognising that R is an orthogonal tensor, and hence that RT = R 1, then the equation
reduces as follows [22].

R UE R EUE U R E 1
U
T E T E T
FTF E
REUE UE (5.4)
Chapter 5 The intrinsic field tensors 78

By observing that Equation (5.3) forces U to be symmetric and thus that UT = U—or
similarly that we asserted the stretch tensors to be the symmetric part—then the previ-
ous equation becomes identical to Equation (5.3). The practical numerical solution of
this formulation is the subject of the subsection that follows.

5.1.3 Polar decomposition: numerical solution procedure


The numerical solution of a polar decomposition is regarded as a difficult and computa-
tionally expensive problem [131]. A possible numerical procedure to undergo a polar
decomposition requires the solution of a matrix square root problem. In the case of the
right stretch tensor, the symmetric U must be found from

UE = FTF . (5.5)

An efficient approach to this is to utilise a standard Eigen decomposition function first


to obtain the eigenvectors and eigenvalues. This can be represented as

>v O @ eig F T F , (5.6)

where in a two-dimensional operation yields a two-by-two matrix of eigenvalues O. Be-


ing diagonal, the square root of O can be easily computed. Four possible solutions result

>O @
0.5
s , where
ª s11 0º ª  s11 0º ª s11 0 º ª  s11 0 º (5.7)
s1 «0 ; s2 ; s3 ; s4
¬ s22 »¼ « 0
¬ s22 »¼ «0
¬  s22 »¼ « 0
¬  s22 »¼

of which only one solution is valid, as determined by a simple objective minimisation.


Firstly, the four possible solutions to Equation (5.5) are computed:

Ui vsi v 1 (5.8)

This is used in the arbitrary objective function

x
i j
^
min ¦¦ ª Fij  U ij º < ª Fij  U ij º
¬ x¼ ¬ x¼ ` (5.9)

to ascertain which of the four possible matrices Ux from (5.7) is the correct physical
stretch of the deformation gradient, following the possible surrender of signs via FTF.
Chapter 5 The intrinsic field tensors 79

Using whichever value x is found to be the optima, the resulting unique value for the
right stretch tensor

UE Ux (5.10)

becomes the solution for Equation (5.5).

5.1.4 Biot strain


We next examine the Biot strain tensor H, which is equivalent to an ILST that has not
undergone any rigid body rotation, or similarly is equivalent to a converged corotation
of the ILST according to the linear rotation tensor. This is

HE U E  I, (5.11)

with components

H ijE U ijE  G ij . (5.12)

The Biot strain relies on the use of the right stretch tensor UE shown in Subsection 5.1.2
obtained via the polar decomposition procedure of the deformation gradient F as shown
previously. It is a linear measure of strain that is accurate for large nonlinear rotation
and large strain assuming the linearity of the metric is deemed adequate—note that for
hyper-elasticity in particular a nonlinear strain measure is preferred. The conventional
Biot strain bears its principal utility from being rotationally insensitive. Its disadvantage,
however, arises from this frame suffering the same linear locking problem as the ILST.

5.1.5 Cauchy–Green and Green–Lagrange strain


Having established the validity of the stretch tensors UE and VE, we can show a defor-
mation tensor to be ascertainable without having to perform the decomposition in Equa-
tion (5.1). Given the assumption of symmetry for the stretch tensors and the operation
shown in Equation (5.4), the square of the right stretch tensor can be found explicitly
using the deformation gradient:

FTF U E T
UE U E 2
(5.13)

The right and left Cauchy–Green tensors make use of this relation, and are defined as
Chapter 5 The intrinsic field tensors 80

CE U E 2
and B E V E 2
, (5.14)

which can, by use of Equation (5.13), be written in terms of the deformation gradient—
thus taking advantage from the assumption of tensor symmetry utilised therein. These
can also be written explicitly in terms of the geometry as

CijE xm ,i xm , j , BijE xi ,m x j ,m (5.15)

Furthermore, the inverse relationships can be defined such that the inverse Cauchy–
Green tensors are

cE C E 1
F 1F  T ½
°
¾ (5.16)
b E
B E 1 T
F F 1
°
¿

The inverse left Cauchy–Green tensor b, in particular, is often referred to as the Finger
Tensor [136].
Using these metrics, the Green–Lagrange strain E can be shown. It is particularly
useful because of its independence to the rotational component of deformation, as can
be seen in Equation (5.4), and also because it can be ascertained without the need to un-
dergo the polar decomposition procedure. It can also be defined with respect to the
Cauchy–Green or by use of the right stretch tensor; these are shown respectively as

(
2
F F  I
1 T
2
C  I
1 E
U  I »¼
1ª E 2 º
2 «¬
(5.17)

which can be expressed as having components depending on terms only from the dis-
placement gradient ’u:

( ij
1
2
ui, j  u j ,i  um,ium, j . (5.18)

Note that the terms from um,ium,j are of the second order and if these are negated the
equation becomes identical to that of the ILST.

5.1.6 Other extrinsic field tensors


Where the Green–Lagrange strain is widely used as a finite deformation measure in the
material or Lagrangian frame, an equivalent tensor attributed to Almansi is also recog-
Chapter 5 The intrinsic field tensors 81

nised that is instead in the Eulerian frame: this, sometimes also referred to as the Al-
mansi–Euler tensor, can be written as

dE
1
2
I  F  T F 1
1
2
I  bE

2 ¬«
I  V ¼»
E 2 º
(5.19)

where the components can be found directly as

dij
1
2
ui, j  u j ,i  um,ium, j (5.20)

In cases of hyper-elasticity the logarithmic Hencky strain is particularly useful due


to its energy conjugate response. It also has the utility of a logarithm in that the cumula-
tive sum of incremental strains is always identical to the total current–reference measure.
It can be written with respect to the left Cauchy–Green tensor as

hE log n V E , (5.21)

noting that a natural logarithm is used.

5.2 The intrinsic frame

In Chapter 3 a linear continuum mechanics basis was presented that stemmed from an
assertion of geometric strain indeterminacy. In that linear theory there is no distinguish-
able difference from an Eulerian or Lagrangian frame as the strains are infinitesimal. In
the same manner of Section 3.2, and as a nonlinear extrapolation of that postulate, the
so-called Intrinsic Frame is now necessary. Being neither Eulerian nor Lagrangian in
the classical sense, the intrinsic frame stipulates that the orientation and hence correct
relative frame is dependent on both geometry and internal properties.
Chapter 5 The intrinsic field tensors 82

a) b) c) d) e)
Y Y y y :

E
:
: : :
:0 :
x :
X X x the
Intrinsic
reference Eulerian co-rotational Lagrangian
Frame

Figure 5.1 a) The reference configuration is undeformed and universally applicable; b)–d) The conven-
tional deformed frames for current configuration; e) In this example, the proposed Intrinsic Frame is an
intrinsically dependent corotational frame of observation

In Figure 5.1 the frames for some finite trellis deformation are illustrated. Note in
particular that the Lagrangian frame is not orthogonal and additionally is non-Euclidean.
The intrinsic frame, shown for the measure of stretch, has any number of orientations,
but is made determinate according to the material’s constitutive model. In effect, the
variation of some value E allows this frame to reflect precisely the decomposition
shown in Figure 4.2.
In contrast to the material property indifference of the extrinsic field tensors, as set
out in Subsection 5.1.1, an equivalent intrinsic field tensor for infinitesimal strain can be
derived. Given its description as the OLST, the denotation œ (ethel) was used to differ-
entiate it from the ILST e; in the following section the intrinsic field tensors are denoted
as unique from their solely extrinsic counterparts in a like manner. This is done using
the upper-case notation (·)E, (·)Œ and (·)Æ for isotropic, orthotropic and anisotropic for-
mulations respectively when referring to items in the nonlinear domain.

5.3 The finite intrinsic field tensors for stretch

The intrinsic frame previously shown suggests that measures of deformation will be af-
fected not only in shear, but in all terms under a finite change. The metric nearest to de-
formation gradient is material stretch, and so the intrinsically dependent stretch tensors
are developed here in reliance to a proposed skewed decomposition. A numerical solu-
tion procedure for this decomposition is also offered.
Chapter 5 The intrinsic field tensors 83

5.3.1 The orthotropic stretch tensors and skewed polar decomposition


In order to derive an intrinsically dependent and thus subsequently, potentially, asym-
metric left and right stretch tensor, the assertion of geometric strain indeterminacy is
invoked as was done to find the OLST in concert with the use of the intrinsic frame. In
the same manner that the ILST was shown to be a combination of an indeterminate met-
ric and an assumption of tensor symmetry (Equation (3.11) and (3.12)), the Right
Stretch Tensor can be assigned the alternate definition for determinacy:

FTF U E T
UE (5.22)

i.e. the indeterminate multiplicative decomposition, and:

U E T
= UE , (5.23)

which is an assumption of tensor symmetry, such that when combined with Equation
(5.22) these yield Equation (5.3). By asserting that the general case for multiplicative
decomposition is FTF = UTU, it follows that FTF = U2 becomes only a special case for
the condition of isotropy. As was done in Chapter 3, the propositionally ill assumption
of tensor symmetry is similarly discarded and we thence require some sound basis to
which the tensor will become both determinate and intrinsically dependent.
The now orthotropic right stretch tensor, variable over an intrinsic field, is denoted
as UΠrather than UE for the isotropic equivalent, with the partial definition

F R ΠUΠ. (5.24)

RŒ, accordingly, is the orthotropic finite rotation tensor. We first note that by removing
the condition of Equation (5.23), Equation (5.22) is still able to conform to the verifica-
tion in Equation (5.4). Therefore, this multiplicative decomposition is potentially
skewed considering only the now reduced-assumption requirement of Equation (5.22)
with UE replaced by UŒ. It is then made determinate by a necessity to satisfy shear equi-
librium. The equilibrium condition is borrowed from the derivation in Section 3.3, and
thus for a two-dimensional case the orthotropic right stretch tensor takes the proposed
general form
Chapter 5 The intrinsic field tensors 84

FTF U ΠT
U Π, where
ª U xxŒ rxya xˆ  rxyb U xxŒ  U yyŒ º (5.25)
U Œ
« »
« ryxa xˆ  ryxb U yyŒ  U xxŒ Œ
U yy »
¬ ¼

where the distribution components are as adapted from the definitions in Equation
(4.11) and (4.12). To reduce the number of coefficients so as to simplify the coming ob-
jective functions of the numerical solution procedure, we can rewrite this equation as

ª U xxŒ xˆ  rxyb U xxŒ  U yyŒ º


U Œ
« », (5.26)
« rˆ a xˆ  rxyb U xxŒ  U yyŒ U Œ
»
¬ yy ¼

noting that the choice of x̂ is totally arbitrary in terms of its prefixed ratio and is incon-
sistent between equations. The ratios of this equation are

a
d yx sin 2 T xy  d xy cos 2 T xy
rˆ and (5.27)
d xy sin 2 T xy  d yx cos 2 T xy

d xy  d yx
rxyb sin 2T xy . (5.28)
2

The rotational component can then be found by inversion of Equation (5.24):

RΠU Π-1
F (5.29)

The equivalent orthotropic left stretch tensor, denoted as VΠrather than VE, uses the
same multiplicative rotation tensor of Equation (5.24), and is of the form

F VŒRŒ (5.30)

where, following the solution of Equation (5.29), VΠcan be found using

VΠR
Π-1
UΠR Π(5.31)

if not otherwise found directly from a left polar decomposition. This left decomposition
is similarly defined as

V Π= FF T (5.32)

such that all terms satisfy the condition


Chapter 5 The intrinsic field tensors 85

ª VxxŒ rxya xˆ  rxyb VxxŒ  VyyŒ º


V Œ
« » (5.33)
« ryxa xˆ  ryxb VyyŒ  VxxŒ V Œ
»
¬ yy ¼

noting that, as with Equation (5.25), the solvable placeholder variable x̂ has no particu-
lar relation to the material coordinate system terms in % x, y .

5.3.2 Numerical solution of two-dimensional skew decomposition


In this section a practical solution procedure for the two-dimensional skew decomposi-
tion is outlined. This same procedure is utilised in the numerical examples at the end of
the chapter. Given a finite deformation measure for aspectual strain, denoted A rather
than D, with components

AijE U iiE  U Ejj (5.34)

an estimation based on the isotropic right stretch tensor in Equation (5.1) can be found.
This secondary estimation UŒs can be obtained using the procedure set out in Subsec-
tion 5.1.3 after correcting it with an approximation for aspectual strain AE based on the
isotropic equivalent:

U Œcc U E  A Erb (5.35)

Note that this function, while resembling the OLST, is not Euclid objective. By vector-
ising the secondary estimation, indicated by braces ^`
˜ , a primary estimation UŒc is ob-

tained following rotation using the tensor transformation in Equation (3.28) as a func-
tion of the linear material principal rotation explicitly found from Equation (3.46):

^U c` T^U cc`
Œ Œ
(5.36)

From this the initial values are found as


xˆinit . U xyŒc  rxyb U xxŒc  U yyŒc , U Œ
xx init . U xxŒc , U
Œ
yy init . U yyŒc (5.37)

for the numerical optimisation

min ¦ fi xˆ , U xxŒ , U yyŒ .


2
(5.38)

i
Chapter 5 The intrinsic field tensors 86

Based on the requirement set out in Equation (5.26) substituted into the reduced as-
sumption definition derived from Equation (5.22), the three objective functions in f are
found:

° xx xy xx yy xy xx yy
­ U Œ 2  rˆ a 2 xˆ 2  2rˆ a r b xˆ U Œ  U Œ  r b 2 U Œ  U Œ 2  ª F T F º ½
¬ ¼ xx °
° °
U xxŒ xˆ  U yyŒ rˆ a xˆ  rxyb U xxŒ  U yyŒ  ª¬ F T F º¼
2
f ® xy
¾ (5.39)
° °
° U yyŒ  xˆ 2  2 xrˆ xyb U xxŒ  U yyŒ  rxyb U xxŒ  U yyŒ  ª¬ F T F º¼ °
2 2 2

¯ yy ¿

Following the convergence of Equation (5.38), where in the numerical examples the
Levenburg–Marquardt algorithm has been applied, the resultant values are used in
Equation (5.26) to yield a numerical solution for the orthotropic right stretch tensor UŒ.
As is shown in Subsection 5.3.1, the subsequent tensors can then be easily ascertained.

5.4 The finite intrinsic field tensors for strain

Having acquired an intrinsic formulation for the stretch tensors, it is demonstrated that
any essential deformation-based tensor can be reformulated and hence generalised as an
intrinsic field tensor.

5.4.1 The orthotropic Biot strain


Knowing now a solution of a skew decomposition according to a skew-diagonal ratio
defined to satisfy moment equilibrium, we use the orthotropic right stretch UΠto define
the Orthotropic Biot Strain Tensor (OBST):

HΠU Π I, having components
(5.40)
H ijΠU ijΠ G ij

Bearing close similarity to the OLST, this can be written as having linear components in
the material principal frame that will, in the first order only, satisfy the formula

HŒ U Œ
 I <r a  A Π<r b
dij  d ji (5.41)
H Œ
ij U E
ij  G ij dij sin Tij  d ji cos Tij  U  U
2 2 E
ii
E
jj 2
sin 2Tij
Chapter 5 The intrinsic field tensors 87

which tends to function best when oriented closer to the material frame ij  1,2,3. Note
that the orthotropic finite aspectual strain can be written as

AΠU <I 1  1 U <I
Œ Œ
(5.42)

Furthermore, generality is held in that Equation (5.40) reduces precisely to the conven-
tional formulation of Equation (5.11) for the isotropic condition. For exactitude in the
second order, the decomposition procedure set out in Subsection 5.3.1 must be adhered
to. Therefore in the case of finite deformation, the components of the OBST can only be
defined as in Equation (5.40).
Now having observed both the conventional and orthotropic Biot strain, we note that
both procedures require the solution of a polar decomposition. Furthermore, as will be
demonstrated in the numerical examples, the OBST does not suffer from the locking
phenomena of the isotropic Biot strain as the frame is convected to a purportedly exact
orientation. This is known because both Biot tensors are equivalent to their respective
co-rotated linear infinitesimal strains, where the aforementioned case of locking was
tested with the ILST and OLST in numerical example 4.6.3.
The supposition to this new non-locking behaviour of the proposed orthotropic Biot
tensor is as follows:

In conventional cases, use is resorted to the Green–Lagrange strain and energy-conjugate


second Piola–Kirchoff (PK2) stress due to their non-locking performance for orthotropic
materials [11] as compared to the locking Biot strain and Cauchy (true) stress. As the
proposed OBST will not suffer from such instances of numerical locking, its use will po-
tentially be preferential to the Green–Lagrange/PK2 tensors where use of those metrics is
disadvantageous with relation to their well-known [137] lack of physical meaning.

This, as compared to the earlier-presented linear theory, becomes prevalent because


these incidents of numerical locking only occur under finite deformation, i.e. nonlinear
deformation, the subject of this chapter. Considering the Biot strain is equivalent to a
co-rotated linear strain, we take note of the following excerpt from Munn [234] on the
validity of H, “The so-called ‘infinitesimal’ strain tensor is shown to be a valid finite
strain tensor for all symmetric deformations (which involve no rotation)”. What Munn
did not comment on was the fact that the infinitesimal strain tensor is not a valid finite
Chapter 5 The intrinsic field tensors 88

strain tensor when describing orthotropic or anisotropic media if the definition of rota-
tion is taken to be that of Z. However, his original statement holds true if rotation is de-
fined as instead the material principal one—the intrinsic equivalent. Adding to this,
comments by Moija and Bufler [137], in their discussion on updated Lagrangian polar
decompositions, reinforce the notion that:

“It is believed, according to the evaluations of Biot [9], Bu‚ [235], Sansour and Bu‚
[236, 237], and Simo and Taylor [238], that the Biot strain tensor as the strain measure
has more advantages than the Green strain tensor since the geometry meaning of the
Green strain tensor is not so clear as that of the Biot strain tensor.”

Noting lastly Wriggers and Grutmann [239], who make point on the utility of Biot’s
tensor as a strain metric, these comments in combination bolster the previous supposi-
tion by the notion that it is: valid in general; valid uniquely because of its rotation; and,
advantageous in physical meaning and utility.

5.4.2 The orthotropic Cauchy–Green tensors


Following the derivation of the essential orthotropic stretch tensors, we can now simply
derive the Cauchy–Green tensors as IFTs. Based on the earlier shown general case for
the right Cauchy–Green tensor in Equation (5.14), which does not use one of the direct
geometric forms, it is proposed that in its orthotropic form, CΠbecomes

U H
Π2
 I .
2
CŒ Œ
(5.43)

This can be written as having components

CijŒ U imŒU mjŒ H imŒ H mjŒ  G ij (5.44)

and similarly the orthotropic left Cauchy–Green tensor is

BΠV Π2
, where BijŒ VimŒVmjŒ . (5.45)

In these three definitions, the direct relation of Equation (5.15) is unseen. Physically, the
Cauchy–Green tensors can be interpreted as the square of the change in material point
distances. Thus the effect of the proposed orthotropic Cauchy–Green tensors is to meas-
Chapter 5 The intrinsic field tensors 89

ure those distance changes according to the current orientation that is material principal
rotation invariant rather than to that which is geometric rigid rotation invariant.
Utilising these intrinsically dependent forms we can similarly show the equivalent
inverse tensors, where the second of these can be described as the orthotropic Finger
tensor

cΠC Π1
and b ΠB Π1
, (5.46)

having surrendered the explicit geometric definition to F.

5.4.3 The orthotropic Green–Lagrange strain


Based on the assertion that FTF = U2 is valid only as a special case, it is thence proposed
that the general form—encompassing both isotropic and orthotropic media—is reduced
to only the latter tensor formulation in Equation (5.17), and thus the Orthotropic Green–
Lagrange Strain Tensor (OGST) Πreplaces the isotropic equivalent E as follows:

Œ=
2 ¬«
U  I ¼» , having components
1ª Œ 2 º
(5.47)
Œij
2
U imU mj  G ij
1 Œ Œ

This can similarly be written as

1 Œ
Œ= ªC  I º¼ , (5.48)

noting that it is currently * is not expressible as an explicit function of the deformation


gradient F, as F has no material-intrinsic dependencies. These formulation will be ex-
tended to anisotropic formulations for Biot and Green–Lagrange strain as HÆ and Æ in
the subsequent Section 5.5
The physical interpretation of the tensor Πas compared to the conventional metric
E is less influential in terms of strain-induced energy as compared to the variation be-
tween the Biot strains. The reason for this being that the use of Lagrangian material
avoids any such numerical locking resulting from the relatively rigid material principal
direction(s). From the standpoint of energy alone, a variation occurs in total shear as a

*
This is not to say it is inexpressible, rather that a formulation has not, as yet, been found. This will be the subject of future study.
Chapter 5 The intrinsic field tensors 90

result of the difference between the shear functions from symmetric shear to simple
shear. To convey this, let us consider the example of an infinitely orthotropic continuum
under some shear angle M (equivalent to J for small strain only): the total shear under an
isotropic and hence symmetric measure will be J 2sin M 2 rather than according to

Πwhere it will be treated as J sin M  sin 0 . From the standpoint of internal numeri-
cal values the tensor Πmay vary significantly due to its asserted material principal ro-
tation versus geometric rigid-body rotational only. These features add perspective to the
supposition of Subsection 5.4.1.

5.4.4 The orthotropic Almansi–Euler strain


In Subsection 5.1.6, the Eulerian equivalent to the Green–Lagrange strain was previ-
ously shown; by utilising the now developed orthotropic left stretch and the Finger ten-
sor, the equivalent intrinsic formulation can be derived. This takes the form

I  V Œ º» I  bŒ .
1ª 2 1
dŒ « (5.49)
2¬ ¼ 2

Notably, Equation (5.20) is inherited now as only a special case for isotropy. In the
physical sense, as with the orthotropic Green–Lagrange strain, this tensor varies from
the conventional Almansi–Euler strain in total shear and rotation. The strain-induced
energy may be slightly less as compared with a conventional formulation due to the
previously mentioned rotational dependency of the shear functions under finite defor-
mation. Despite being closely similar from an invariant standpoint, the actual values
within the tensor may vary significantly given that it invokes a rigid body rotation fol-
lowing the assertion of geometric strain indeterminacy.

5.4.5 The orthotropic logarithmic Hencky Strain


Noting the previously described utility of the logarithmic strain measure of Hencky, an
intrinsic field formulation follows making use of the orthotropic left stretch, where
again the use of a direct geometric relation is prohibited without the condition of isot-
ropy. This, the orthotropic logarithmic Hencky strain, is defined here as

hΠlog n V Π, (5.50)
Chapter 5 The intrinsic field tensors 91

considering again that it is the natural logarithmic measure for strain. The physical dif-
ferences are inherited directly from those described for the new stretch and Biot tensors.
We regard here that each of these formulae, Equations (5.40) through (5.50) will be
unique to its conventional counterpart given it is under at least some principally shear-
ing deformation and describes a non-isotropic continuum. Furthermore, it is claimed in
a reserved manner that any essential deformation metric can be correspondingly refor-
mulated as an IFT, to which in justification we appeal to comments made by Haupt and
Tsakmakis [164] in their disposition on mixed or duel variable tensor functions in me-
chanics. Note that their epsilon H is the Piola strain tensor, which is closely related to the
Almansi–Euler strain.

“Moreover, it is quite clear, that each arbitrary isotropic tensor function of E or H could be
taken as a strain measure, if certain general properties are guaranteed.”

As any of the conventional essential deformation tensors can be written in close relation
to these, it stands to reason that having replaced those same tensors with new and inde-
pendent ones we can suggest the same conclusion for IFTs.

5.4.6 An additional note on micropolarity


Pursuant to the suggestion of coupled moments in Subsection 4.5.2, where it was sug-
gested that micropolarity could be described within the mathematical infrastructure
shown here without having had to invoke the wryness tensor, we make a further suppo-
sition:

It is quite possible that micropolar behaviour can be treated in a manner so as to be effec-


tively the same as defined by F. and E. Cosserat [169] and later Eringen [174] by the sim-
ple addition of moment terms to the extraneous degrees of freedom of the system in ques-
tion. In that sense, and as shown by reference [190], the local measures of micro-rotation
would no longer be necessary due to the pervasive ability of the proposed orthotropic
Biot strain to macroscopically represent these. *

This idea is well exemplified by the erroneous portrayal of twisted-cord rope extension
kinematics by the micro-polar approach [240] and similarly in the bookshelf mechanism

*
Though it stands to be shown, there is reasonable evidence provided here and in the literature to suggest that micropolar behaviour
could be represented within the proposed continuum mechanics infrastructure with only little additional formulation.
Chapter 5 The intrinsic field tensors 92

[193], noting that such a system can be described by SOCM without needing to intro-
duce a fictitious micropolar moment and micro-rotation. We again assert that only a
macro-rotational description is analogous.
Above all, it should be noted that the careful mathematical treatment of these cases
has precise physical meaning, whereby the proposed theory has no such necessity for
micro-structure, micro-rotation or micropolar moments in order to cause the prevailing
tensor asymmetries. This is demonstrated by the ability to give strain asymmetry with-
out stress asymmetry. Instead, the theory puts forward a notion that the kinematics of
homogeneous orthotropic continua are intermediate to widely accepted constitutive ex-
tremes [10, 148] while relying on intrinsic material properties.

5.5 Finite anisotropic IFTs

In the same manner that the anisotropic linear strain tensor æ was previously shown to
follow the derivation of the orthotropic linear strain tensor œ, it can be shown that the
equivalent intrinsic field tensors for finite anisotropic deformation can be found. Where
Equation (5.24) was able to encompass the conventional polar decomposition of Equa-
tion (5.1), it too can be generalised by the anisotropic form

F R ÆUÆ , (5.51)

where UÆ is the anisotropic right stretch tensor and RÆ is the anisotropic finite rotation
tensor. Borrowing the earlier derived planar equilibrium condition in Equation (3.69),
the respective indeterminate decomposition and requirement for anisotropic moment
balance are as follows:

FTF U Æ T
U Æ , where
ª U xxÆ rxya xˆ  rxyb U xxÆ  rxyc U yyÆ º (5.52)
Æ
U « a »
¬« ryx xˆ  ryxU yy  ryxU xx
b Æ c Æ
U yyÆ ¼»

thus utilising the previous equations from Section 3.9 as well as the counterparts with
cycled indices, such that
Chapter 5 The intrinsic field tensors 93

a
s 2 Dxyxy  c 2 Dyxyx  Dxyyx  sc Dxxxy  Dxxyx  Dyyxy  Dyyyx
r
Dxyxy  2 Dxyyx  Dyxyx
xy

b
scDxyxy  scDyxyx  c 2 Dxxyx  c 2 Dxxxy  s 2 Dyyxy  s 2 Dyyyx
r . (5.53)
Dxyxy  2 Dxyyx  Dyxyx
xy

c
scDxyxy  scDyxyx  s 2 Dxxyx  s 2 Dxxxy  c 2 Dyyxy  c 2 Dyyyx
r
Dxyxy  2 Dxyyx  Dyxyx
xy

By the same reasoning of Section 5.4, the Cauchy–Green and Biot tensors—among
others— can be found with direct relation to the anisotropic right stretch:

CÆ U Æ 2
, HÆ UÆ  I (5.54)

where a linearisation of the anisotropic Biot strain, built from the isotropic stretch, is

HÆ U E
 I <r a  U E <I 1 b r 1 U E <I r c
(5.55)
H ijÆ U E
ij  G ij rija  U iiE rijb  U Ejj rijc

The physical kinematic condition of a continuum described be these tensors is one


that represents a material principal orientation satisfying not only the asymmetric shear
relation but also the requirement of the normal strains such that coupling terms are taken
into account for the equilibrium. Moreover, Green’s Lagrangian tensor—shown in its
conventional form E in Equation (5.17)—is now expressible as the anisotropic Green–
Lagrange strain tensor Æ such that

Æ=
2 «¬
U  I »¼ .
1ª Æ 2 º
(5.56)

In addition to this, equivalent anisotropic IFTs can easily be shown for the left and in-
verse tensors as well as an anisotropic logarithmic Hencky strains. These formulations
are omitted here for lack of necessity.

5.6 Generalised decomposition theories

By considering that the conventional stretch tensors rely on polar decomposition and
that an infinitely orthotropic tensor separation bears similarity to the well known de-
Chapter 5 The intrinsic field tensors 94

composition of Cholesky, the objective infinite and intermediary decompositions hith-


erto proposed can be treated as purely mathematical operations; these are presented re-
spectively as follows. Note that in this section, notations used elsewhere may be as-
signed alternate definitions in order to remain consistent with well establish mathemati-
cal denotations.

5.6.1 The objective Cholesky decomposition


Where typical polar decomposition takes the form F = RU such that R is orthogonal and
U is symmetric, if we denote FTF as the symmetric matrix A the solution of this opera-
tion is defined by

A U2. (5.57)

This, in two-dimensions, can be seen as a requirement to satisfy

ªU xx U xy º
U «U U ». (5.58)
¬ xy yy ¼

Under the condition of infinite horizontal orthotropy, the proposed form of matrix U
would be an upper triangle matrix, and thus the decomposition F = RU could, under the
material frame of observation, be considered as a QR factorisation. Following the pro-
cedure of Equation (5.4), this becomes a Cholesky decomposition that generally takes
the form

A U T U. (5.59)

Note that the use of U for an Upper triangle matrix is coincidental in this case, as the
intention is to keep notation similar to those used in mechanics. In solution it must
therefore satisfy

ªU xx U xy º
U « 0 U ». (5.60)
¬ yy ¼

It was shown in Subsection 5.3.2 that Equation (5.57) can be solved using an Eigen
decomposition, and given that A is symmetric so too must be U. This is a critical feature
if we consider that, as is shown by a conventional Mohr’s circle, Eigen decomposition
can be performed from any frame of observation while remaining objective. This prop-
Chapter 5 The intrinsic field tensors 95

erty is known as satisfaction of Euclidean Objectivity. When U is rotated the off-


diagonal terms are affected equally and so the condition of Equation (5.58) is main-
tained. Alternatively, following some rotation, diagonal terms in the Cholesky decom-
position equally added-to would no longer satisfy the condition set out in Equation
(5.60). For this reason it can be seen that a conventional Cholesky decomposition can-
not satisfy the condition of Euclidean Objectivity, and thus cannot be used as an opera-
tion in continuum mechanics.
In the same fashion that the polar decomposition can be solved as an Eigen problem,
the various factorisations QR, QL, RQ and LQ are all reducible to a standard Cholesky
form. Where Q represents an orthogonal (rotation) matrix, and R and L represent Right
and Left triangle matrices, these are respectively: right infinitely horizontal orthotropic;
right infinitely vertical orthotropic; left infinitely horizontal orthotropic; left infinitely
vertical orthotropic. Note that right and left refer to upper and lower triangularity, not
right and left observational frames.
In order to ensure that such decompositions are instead objective, the general Objec-
tive Cholesky requirement is here proposed to replace the form of Equation (5.60). Re-
quiring the numerical solution of x̂ , it is

ª sin 2T º
« U xx xˆ cos 2 T  U xx  U yy
2 »,
U « » (5.61)
« xˆ sin 2 T  U  U sin 2T U yy »
«¬ xx yy
2 »¼

where T is the angle of rotation from the reference frame of decomposition. This, in the
meaningful mechanical analogue, is the angle from the material frame.

5.6.2 Generalised Polar–Intermediate–Cholesky decomposition


Having developed an objective form of the Cholesky decomposition, it can now be
shown that both the Polar and objective Cholesky decompositions can be generalised as
boundary cases of the variable-skew polar decomposition presented in Subsection 5.3.1
By introducing an arbitrary skew ratio E, which under the mechanical analogy is com-
prised of the balance ratio, the objective decompositions are made general by the fol-
lowing objective form; it is described here as the generalised Polar–Intermediate–
Cholesky Decomposition (P–I–C decomposition).
Chapter 5 The intrinsic field tensors 96

ª xˆ ª¬ E 2sin 2 T  1  cos 2 T º¼ º
« U xx »
«  U xx  U yy E  1 2 sin 2T »
U « » (5.62)
« ¬
« xˆ ª E 2 cos 2 T  1  sin 2 T º »
¼ U yy »
« U  U »
¬ xx yy E  1 2 sin 2T ¼

By this requirement, the additional feature is added to Equation (5.61) of full asym-
metric variability. This offers the novel intermediary cases that were necessary to ad-
dress the theory of Strongly Orthotropic Continuum Mechanics previously. As a result,
we can summarise the various decomposition/factorisation methods according to their
conventional sorts, orthotropy and direction using only the skew ratio E:

Table IX. Decomposition categorisation and the Polar–Intermediate–Cholesky generalisation.

Decomposition Type Relative Orthotropy Direction E Generalisation

Eigen right 0.5


Polar isotropic none
Eigen left 0.5

QR horizontal 1
right
QL vertical 0
Obj. Cholesky infinite P–I–C
RQ horizontal 1
left
LQ vertical 0

right 0<E<1
intermediary intermediary orthotropic any
left 0<E<1

Note that in Table IX each form can be considered in a right or left relative frame based
on whether it is equated to A = FTF or FFT.

5.7 Numerical examples

In this section four numerical example are presented to demonstrate the behaviour of the
intrinsic field tensors. First, the proposed orthotropic Biot strain and stress are validated
against the previous corotational example in Chapter 4. Next, the general behaviour of
the proposed metrics is analysed for a large strain—and hence exemplifying—problem.
A further example is shown for a series of large rotations to demonstrate the efficacy of
Chapter 5 The intrinsic field tensors 97

the solution procedure and that all proposed tensors can satisfy Euclidean objectivity
and material frame-indifference. The final example shows the response over the full
range of trellis shear and illustrates the ability of the orthotropic Biot strain to replicate
energy-free trellising.

5.7.1 Example 1: Verification of the IFTs against corotational result


Having partially based the validation of the IFTs by reasoning of the close relationship
between the corotational method and the Biot strain [130], the existence of the same
close relationship between the material principal corotational method and orthotropic
Biot strain is verified here. In order to achieve this, the same deformation gradient (a
1% trellis motion) and material properties as in numerical example 4.6.3 were assigned
to the proposed formulation of Equation (5.25). Using the deformation gradient

^%,%` ª0.9999495 0.0100505º


F «0.0100505 0.9999495»
1% ¬ ¼

and adhering to the exact numerical procedure set out in Subsection 5.3.2, the
orthotropic right stretch tensor UΠand finite material principal rotation RΠ(Equation
(5.29)) were ascertained from the skewed polar decomposition as

^%` ª0.9999999 0.0200699 º


UŒ « 0.0000301 0.9997986 » and
1%
¬ ¼
^%` ª 0.9999498 0.0100204 º
RŒ «0.0100204 0.9999498 » ,
1%
¬ ¼

such that UΠsatisfies the equilibrium ratios in consideration of Equations (5.27) and
(5.28). This compares to the conventional isotropic right stretch tensor and finite rota-
tion that are identical to the deformation gradient due to its symmetry in this example:

^%` ª0.9999495 0.0100505º ^%` ª1 0 º


UE «0.0100505 0.9999495» and RE «0 1 »
1% ¬ ¼ 1% ¬ ¼

While the symmetry of UE is expected, we observe in particular that it is also symmetric


with respect to the secondary diagonal causing the same numerical locking as if the
ILST had been used for the finite problem. Furthermore the rotation tensor RE indicates
Chapter 5 The intrinsic field tensors 98

no rigid rotation of the frame % and therefore a conventional corotational procedure


will not improve this result.
Following this, the intrinsic and extrinsic forms of the Biot and Green–Lagrange
strains were solved as per Equations (5.40) and (5.47) in succession.

Table X. Comparison of conventional linear and Biot strain and stress for conventional extrinsic and pro-
posed intrinsic formulations against previous results.

Strain (%) Stress (MPa)


Model
x y xy yx x y xy

(co-rotated) ILST 5.051E 3 5.051E 3 1.005 — 4.556 4.556 3.6110

Biot 5.051E 3 5.051E 3 1.005 — 4.556 4.556 3.6110

 2.014E 2 4.543E 8 3.009E 3 2.007 3.626E 2 1.821E 2 3.6108


Co-rotated OLST
 4.543E 8 2.014E 2 2.007 3.009E 3 1.821E 2 3.626E 2 3.6108

 2.014E 2 4.543E 8 3.009E 3 2.007 3.626E 2 1.821E 2 3.6108


Orthotropic Biot
 4.543E 8 2.014E 2 2.007 3.009E 3 1.821E 2 3.626E 2 3.6108

Green–Lagrange |0 0 1.005 — |0 0 3.6108

Marc Total Lgrg. 0 0 1.005 — 0 0 3.6108

Ortho. Grn–Lgrg.  |3.015E 5 2.011E 2 2.007 3.008E 3 |3.618E 2 3.618E 2 3.6104

Table X presents the results of the numerical example in Subsection 4.6.3 alongside the
conventional Biot strain and the now-developed orthotropic Biot strain. An additional
significant figure is given for shear stress in order to convey the differences between
even a very small yet finite trellis deformation. Note that while the orthotropic PK2
stress appears significantly different from the conventional PK2 stress, the von Mises
stress for each of these is 6.2537 and 6.2541 MPa respectively. The small reduction in
overall stress energy from the proposed PK2 stress occurs due to the difference between
measurements of shear in the Lagrangian material frame as compared to the proposed
intrinsic (material) frame that was described in Subsection 5.4.3.

5.7.2 Example 2: Behavioural comparison between the IFTs and EFTs


The physical interpretation of the proposed IFTs is aided by illustrating the strain and
stress response for the conventional tensors against the orthotropic linear, Biot and
Chapter 5 The intrinsic field tensors 99

Green–Lagrange strains under finite deformation. In this numerical example we use a


composite material with a fibre elastic modulus of 300GPa, a matrix modulus of 3GPa,
a primary Poisson ratio of 0.3 (for each) and a fibre volume ratio of 0.5. Using the stan-
dard rule of mixtures [229] the primary and secondary material principal moduli and
Poisson ratios were determined: E1 , E 2 , Q 12 , and Q 21 respectively. Using a 100 ×

100 >„^  nt with a 20 


  >\ @
 @@ 
     >
condition, the deformation in Figure 5.2 was achieved.
Global Y (mm)

100

50

0
0 50 100
Global X (mm)

Figure 5.2. The deformation condition analysed: trellis motion for a 20
 >\ @ement.

Three energy conjugate tensor pairs were analysed for the ascribed deformation. The
first of these was the infinitesimal linear strain tensor and Cauchy stress in both its iso-
tropic form and intrinsic orthotropic form.
Shear strain, œ , œ (%)

yx
œ

40
Shear stress, V , V

200
yx

xy
œ

0°, Linear
20
xy

0°, OLST 0
Prnpl. e
0
Prnpl. œ
-200
-20
-40 -20 0 20 40 -600 -400 -200 0
Axial strain, œxx, œyy (%) Axial stress, Vœ , V
xy yx
œ

Figure 5.3. Infinitesimal linear strain and stress for isotropic and orthotropic formulations.

In Figure 5.3 the resulting values are shown, where the 0° fibre orientation was used.
The resulting plot is closely similar to the previously shown small (1%) strain plots of
numerical example 4.6.3, despite the now very large strain magnitudes. Moreover, even
though the OLST indicates the asymmetric shear distribution and has a non-zero rota-
Chapter 5 The intrinsic field tensors 100

tion tensor, it does not indicate a rotation in orientation on the extended Mohr plot, i.e.
an angular difference between the co-ordinate joining chords of the conventional
(dashed line) and proposed (continuous line) states. This attribute is inherent to a linear
formulation, and leads to the Mohr plot for Cauchy stress having identical stress states.
yx

yx
H
Shear Biot stress, V , V
Shear Biot strain, H , H

40

xy
200

H
xy

0°, Biot
20 0°, OBST
0
Prnpl. H E
0
Prnpl. H Œ
-200
-20
-40 -20 0 20 40 -400 -200 0 200
H H
Axial Biot strain, Hxx, Hyy Axial Biot stress, Vxx, Vyy

Figure 5.4 Isotropic Biot strain tensor is compared to the proposed OBST for the trellis deformation as
well as the Biot (corotational Cauchy) stress states.

In the second case, results are plotted for both the Biot strain tensor and the pro-
posed OBST (Figure 5.4). Where in Figure 5.3 there was no indication of relative rota-
tion between the strain tensors by the co-ordinate chords, there now is clearly a varia-
tion. The angle of the OBST chord, 2I from the horizontal, shows that the original
frame would be less than 45° from the conventional strain-principal orientation for rota-
^`
tion-free motion—precisely the pre-counter-rotated frame :0-cr indicated in the hori-
zontal fibre corotation of Figure 4.7. It should also be observed that the strain state for
the OBST has moved slightly to the left, this is indicative of the increase in fibre-normal
compression (the left coordinate) and the decrease in axial fibre compression. The direct
result of this is clear in the Biot stress plot (also the corotational Cauchy stress), where
the proposed strain state (solid line) has clearly reduced in diameter showing a reduction
in deviatoric strain-induced energy. Its centre has also moved significantly toward the
vertical axis, conveying a marked reduction in dilatational strain-induced energy—this,
in essence, is demonstration of the solution to the numerical locking problem of corota-
tional Cauchy methods under orthotropy. This result, as a corollary to the suggestion in
numerical example 4.6.3, is purported to be the exact kinematic and energy condition
for such a problem.
Chapter 5 The intrinsic field tensors 101

yx

Œ
yx
40

Shear PK2 stress, S , S


Grn–Lgrg strain, Œ , Œ

200

Œ
xy
xy

20 0°, Grn–Lgrg 100


0°, OGST 0
Prnpl. E
0 Prnpl. Π-100
-200
-20
-40 -20 0 20 -200 0 200
Green–Lagrange strain, Œ xx, Œ yy Axial PK2 stress, SŒ , SŒ
xx yy

Figure 5.5 Green–Lagrange strain and stress plots

The last tensor pairs considered in this example are the isotropic and orthotropic
forms of the Green–Lagrange strain and 2nd Piola–Kirchoff stress tensors. Notably, the
orthotropic strain measure proposed illustrates the presence of a dilatational strain, spe-
cifically a compaction, that is absent in the conventional plot given its centrality to the
vertical axis. Given that we know this problem to force a compaction on the continua,
the superiority of the proposed measure becomes apparent.
The physical meaning to these tensors was described in Section 5.4.3 earlier, where
the absence of a fibre normal component to strain—typical of Total Lagrange solu-
tions—eliminates said incidents of numerical locking from the use Biot’s methods. For
this reason we regard the comparative stress states of the proposed Figure 5.4 and the
conventional state (dashed) in Figure 5.5. This is best illuminated by a direct compari-
son of the von Mises stresses for each case:

V vm
E
5.845 MPa V vm
Œ
4.308 MPa
E Œ
S vm 4.260 MPa S vm 4.169 MPa

The first of these stress invariants, the Biot von Mises V vm


E
, is a numerically locking so-
lution (or more precisely, numerically hardening), the effect of this will be more con-
clusively shown in the full-range strain example 5.7.4. The next of these, the orthotropic
Biot von Mises, is substantially lower due to its kinematic exactitude, noting also that it
accounts for fibre-normal energy. The PK2 energies that follow are both lower again
due to their account of the fibre-normal component, though the proposed form, the
Œ
orthotropically determined PK2 von Mises S vm , is smaller in value because of its
Chapter 5 The intrinsic field tensors 102

asymmetric measure of shear—noting that this decrease in energy overshadows the in-
crease resulting from the account of compaction. We note here that this shear energy
component is similarly lower between the two Biot stresses as well. While the conven-
E
tional PK2 von Mises S vm is actually closer to the purportedly exact result than the pro-

posed intrinsic PK2 von Mises, it is pointed out that this could be described proverbially
as being two wrongs making a right.

5.7.3 Example 3: Euclidean objectivity and material-frame indifference


In Chapter 3 the OLST was presented and had its Euclidean objectivity and material-
frame indifference verified in numerical example 3.10.1, in this example the same ex-
pectations are verified for the Biot strain/co-rotated Cauchy stress and Green–Lagrange
strain/PK2 stress. Here, a more systematic approach is taken in that a 90° incremental
range is examined such that the extended Mohr plot for strain demonstrates objectivity
and the Mohr plot for stress shows us frame indifference.

800
Global Y (mm)

600

400

200

0
0 500 1000
Global X (mm)

Figure 5.6 Translational and rotational motion of the domain under trellis deformation

Utilising the same element, large 20 >


>

 \ \>>
the previous example, and a series of ! 
†‡!!!!< T
> 

@e-
ments, the motion of Figure 5.6 was generated. This motion shows that the observa-
tional frame can be in motion and at any rotational orientation leaving the invariant
form of the proposed tensors unaffected. Given that each of these strain tensors is used
to calculate a stress tensor according to the proposed extended stiffness matrix, which
must also be transformed in rotation, then the invariance of the stress tensor is indica-
tion that the constitutive material law is indifferent to the frame to which it is observed.
Chapter 5 The intrinsic field tensors 103

0°, Biot
yx

yx
H
0°, OBST
Shear Biot strain, H , H

Shear Biot stress, V , V


40

xy
-10°, "
xy

200

H
-20°, "
20
-30°, " 0
-40°, "
0
-50°, "
-60°, " -200
-20
-70°, "
-40 -20 0 20 40 -80°, " -400 -200 0 200
H H
Axial Biot strain, Hxx, Hyy Axial Biot stress, Vxx, Vyy

Figure 5.7 Euclidean objectivity and principle of material frame-indifference for Biot strain and stress

These properties, respectively, are demonstrations of Euclidean objectivity and sat-


isfaction of MFI, and are shown in Figure 5.7 for the proposed orthotropic Biot strain
and its energy conjugate material principal corotational Cauchy stress. In each figure,
the dashed line shows the orientation of the invariant Mohr’s circle for the equivalent
conventional strain and stress.

0°, Grn–Lgrg
Œ
yx

yx

40
Shear PK2 stress, S , S

0°, OGST
Grn–Lgrg strain, Œ , Œ

200
Œ
xy

-10°, "
xy

20 -20°, " 100


-30°, " 0
-40°, "
0 -50°, " -100
-60°, " -200
-20 -70°, "
-40 -20 0 20 -80°, " -200 0 200
Green–Lagrange strain, Œ xx, Œ yy Axial PK2 stress, SŒ , SŒ
xx yy

Figure 5.8 Conventional and proposed Lagrangian measures of strain and stress.

In Figure 5.8, the same performance is demonstrated for the orthotropic Green–
Lagrange strain and its energy conjugate stress. Once again, the coincident and equal-
length chords indicate them each to be consistent and indifferent to the observational
frame. As with the previous numerical example, there is less variation in the Lagrangian
measures as compared with the corotational measures. The results show that each of
these proposed IFTs satisfies the mechanical requirements for a valid continuum tensor.
Chapter 5 The intrinsic field tensors 104

5.7.4 Example 4: Full-range nonlinear trellis shear


In this, the final numerical example, the purportedly superior orthotropic Biot strain
measure is exercised over the full range of trellis shear. This is done so as to display its
ability to approach energy-free motion as the material approaches infinite orthotropy
without the use of non-Euclidean coordinates. It is also used to justify the categorisation
of conventional Biot measures, antecedent to this example, as being susceptible to nu-
merical locking given strong orthotropy.

100

50

0
0 50 100 150

Figure 5.9 Increments of trellis shear motion (only 20 shown)

In order to show a full-range response to trellis motion, 40 increments were analysed


for orthotropic ratios of 1, 10, 100 and 1,000. Figure 5.9 illustrated the kinematic
reponse for 20 incremements from zero to 99% range. While 99% range is a rarely
attained in practical situations, magnitudes of shear deformation commiserate to this are
often utilised in hand-formed woven composite objects [74]. To present the 40 Mohr
plots for each orthotropic ratio the trumpet plot is introduced here. In this, the vertical
axis represents the fraction of total possible displacement G of the diagonal. Each
horizontal slice of the trumpet then displays the Mohr plot for stress at that displacment.
The surface of this is utilised by overlaying a colour map for the von Mises invariant
stress, which is a single measure for both the deviatoric energy, or diameter of the circle
at that slice, and the dilatational energy, the shift to the left or right.
Chapter 5 The intrinsic field tensors 105

Trellis 250 80
fraction
1 200 1
60
150
0.5 0.5 40
100

10 4 20
0 50 0 2
-20 0 -8 -6 0
-10 -4 -2 -2
a) 0 -10 b) 0 -4
2

35
35
30
30
1 1 25
25
20 20
0.5 0.5 15
15
10 10
0 1 0 1 5
5
-3 0 -3 0
-2 -1 -2 -1
-1 -1
c) 0 d) 0

Figure 5.10 Trumpet plots of Mohr's circle for Biot stress, in MPa, against trellis fraction with a von
Mises stress contour in MPa: a) an isotropic analysis, i.e. an orthotropic ratio of 1, where the conventional
and proposed stresses are indistinguishable from one another; b) a ratio of 10 shows noticeable variation
at large deformation; c) a ratio of 100 yields a substantial variation of invariant stress; d) approaching
infinitely orthotropic, with a ratio of 1000, the proposed stress tensor approaches energy-free motion

Figure 5.10 displays the resulting plots for each of the isotropic, the weakly and
strongly orthotropic, and the approaching-infinitely orthotropic materials. In each case
the largest of the trumpets is the conventional Biot stress, and the equal or smaller trum-
pet is the proposed solution. As is the case with all the formulations presented in this
work, the isotropic problem—generalised as having an orthotropic ratio of 1—produces
an identical strain-induced energy, and an identical strain state, to the conventional Biot
result over the entire range of motion. We next observe the weakly orthotropic instance,
with a ratio of just 10, where the differential between trumpet shells is significant at
large strains. In the next plot a now substantial difference in energy is observable over
the full range of finite strain, with the conventional formulation adding around 300%
resistance to the proposed and purportedly exact. The last plot, with an orthotropic ratio
Chapter 5 The intrinsic field tensors 106

approaching the infinite case, demonstrates a marked difference in energy. This


orthotropic ratio of 1,000, it should be noted, is if anything a conservative value when
considering composite forming that has similar magnitude strains, and orthotropic ratios
in excess of 10,000. This last example most clearly justifies the categorisation of the
behaviour of normal Biot strain/corotational Cauchy stress as numerically locking. It
also shows the tendency of the proposed solution towards the previously stipulated con-
dition of energy-free trellis motion as the material approaches infinite orthotropy—this
is shown as a thin black line about the axes. Considering this, the objectives of this ex-
ample have been met; and the earlier-made supposition of Subsection 5.4.1 has been
further supported, whereby the proposed Biot measures were said to potentially prefer-
entially replace certain instances of use of the non-Euclidean Green–Lagrange tensor.

5.8 Chapter summary

In the previous chapters we have developed an argument for a new kinematics represen-
tation of linear strain under small or large rotation, of which its equilibrium equations
were applied to present an equivalent condition for the spectral decomposition of finite
deformation. As this readily produces the orthotropic Biot strain tensor, we have been
able to show the result to be exact in comparison to the earlier corotational examples.
By this validating factor it is that the many finite deformation tensors are rationally de-
rived. These, together with the linear strongly orthotropic measure, are referred to as the
intrinsic field tensors—a class of physical tensors that are shown to be capable of de-
scribing the general kinematic behaviour in any desired frame of observation and for
any level of isotropy, orthotropy or anisotropy in the continuum.
Chapter 6
A case study
Trellis shear has in this study, thus far, been a driving example of strongly orthotropic
kinematic behaviour. No more theoretical additions are to be made, and instead, for this
case study, experimental analysis of pre-impregnated woven composites is focused
upon to demonstrate the efficacy of persistent material principal orthogonality and infi-
nitely orthotropic kinematics. In the first section, pre-preg mechanics and a visco-plastic
constitutive model based on a commercial code are given. Following this, an experi-
mental commission utilised here is justified and described before the assumptions for
the explicit strain equations are given along with the relevant kinematics for application
into a stress model. This is used, in the section thereafter, to design a parameter identifi-
cation and characterisation technique that is closely linked to a number of custom ex-
perimental techniques—the results of which are used to strengthen the argument for
SOCM and the orthotropic Biot strain in particular as a physically analogous theory.

6.1 Mechanics and rheology of CFRPs in forming

State of the art approaches to finite element simulation have implemented some of the
most advanced rheological and constitutive models for forming analysis of polymeric
composites with continuous fibre-reinforcement. These models have been predomi-
nantly derived for thermoplastic materials; consequent to this, certain physical mecha-
nisms are not demonstrated when they are instead used to model thermoset impregnated

107
Chapter 6 Case study 108

fabrics. Furthermore, a coordinate system dependency was shown in the literature re-
view to be common to the most successful rheological models.

6.1.1 Mechanics of engineering pre-pregs


The structure of laminated pre-forms invariably results in a complicated modelling task
for even the simplest loading cases in forming analysis. This is due to nonlinearity in
instances of large deformation and the discontinuity and anisotropy of the media—as
their heterogeneity is on a scale significant to component size.

A preform stack Individual plies Fibre families Individual tows

Inter-laminar sliding Intra-laminar tor- Out-of-plane bending Axial strain


Inter-laminar rotation sional shear Intra-laminar shear Tightening
Fibre-normal compaction In-plane bending

Figure 6.1. Deformation modes within a composite pre-form

Figure 6.1 shows, for example, a plain weave laminate consisting of a stack of uncured
engineering pre-pregs with many levels of meso-structure. Note that the dominant
modes in forming analysis are in-plane shear and fibre-normal compaction.
The most efficient approaches to forming analysis consider the material on a macro-
scopic scale—though many successful investigations have been made into meso-scale
structural modelling as reported in the literature review—as the nonlinearity, anisotropy
and transience of the analysis type require high computational costs for even the sim-
plest macroscopic modelling approaches. In this approach, fibre families can be mod-
elled separately, although they can be constrained by their associated half as a result of
compatibility at shared nodes. The implication is that weave type may have an impact
on response that does not affect the model constitution, only the parameters within it.
This concept was illustrated in Section 2.2, where only the constitutive models are con-
sidered in characterisation and analysis.
Chapter 6 Case study 109

6.1.2 A constitutive/rheological model


In numerical example 4.6.3 we considered a static small strain analysis of a plane-
weave composite under trellis shear, and compared three commercial codes based on
classical mechanics with the proposed and analytical solutions. In this subsection we
study the commercial PAM-FORM platform that is based on explicit rather than implicit
FEA.

}shear coupling
G , Q

}matrix
E m P
M
E1
E2
}fibre

Figure 6.2. Mechanical topology of material type 140

Figure 6.2 depicts Material type 140, which is reported to simulate the behaviour of an
engineering pre-preg under forming conditions [74]. Its construction comprises of a
parent sheet for shear coupling, a visco-elastic matrix sheet and a nonlinear elastic biax-
ial fibre subset, which closely follows the IFRM model, though requires additional con-
straints to improve solver stability and convergence.
The shear coupling acts as a stabilising element for numerical instability and as a
method of modelling certain other unique properties such as the fibre shear locking
phenomenon

°­ Mlock G G
M® ,
°̄t Mlock G G lock

where M is the biaxial offset, and G is the shear modulus. Here, G lock is a constant that

models the post-locking shear behaviour.


The matrix, or thermo-visco-elastic material phase, acts to simulate the viscous resin
component of the pre-preg material. The initial elasticity is such that

E m max ª¬ E1 , E 2 º¼
(6.1)
Qm 0.5,
Chapter 6 Case study 110

given E1 and E 2 are the fibre moduli for axial directions. For example, in a plain weave
this represents the warp and weft respectively (in this case similar), noting that the Pois-
son ratio (Qm) for the matrix is assumed and its value of 0.5 denotes incompressibility.
The deviatoric stress is given by

sij V ij  V mG ij (6.2)

where the invariant mean (or hydrostatic) stress is

1
Vm trace V (6.3)
3

Based on the Maxwell Model [241], the shear stress can be expressed by

sij 2 P eij , (6.4)

where eij is the strain rate. The variable viscosity is given by the power law

P12 K en 1 , (6.5)

using the invariant measure of strain rate e eij eij and temperature dependence by

§ T0  T ·
K K exp ¨ [K ¸ , and
© T ¹
(6.6)
§ T T ·
n n exp ¨ [ n 0 ¸,
© T ¹
where K , [K , n and [ n are the power law coefficients, T is the temperature and T0 the

reference temperature. For parameter identification in the forming state T T0 , devia-


toric stress reduces to
n 1
sij 2K0 eij eij eij . (6.7)

For the linear elastic fibre phase, interwoven through the matrix material are the fi-
bre bundles acting as a relatively rigid and elastic trellising frame. For a biaxial cloth

^`
E1 elastic modulus of direction 1 fibres
^` 
E 2 elastic modulus of direction 2 fibres
Chapter 6 Case study 111

and, for nonlinear power law shear

G ij f eij LC1, (6.8)

where each modulus can incorporate the effects of weave straightening and other such
nonlinear effects if necessary. Due to the discontinuous nature of the fibre bundles, a
continuum approach would lead to significant overestimation of the bending and shear
stiffness of the fibres. Thus a bending factor (Factbending) and shear factor (Factyz,xz) are
incorporated into the model to account for the effect of sliding between fibres,

kic Facti u ki , i  yz , zx, bending


Typically: 0.1  Facti  1.E  6.

where ki is an uncorrected stiffness coefficient for a shell element.

6.2 Testing in trellis shear

Two dominant methods exist for testing trellis shear deformation: picture-frame and
bias extension experiments. Each method has particular benefits and disadvantages
making the choice of which reliant on the nature of the subsequent analysis.

6.2.1 Picture frame rig


Picture-frame tests apply an idealised translational boundary condition to a specimen,
which in the case of small strain is required to undergo pure shear but under large strain
will replicate the trellis motion of Figure 6.3 b). They are often favoured because of
their kinematically determinate properties and quasi-uniform specimen strain field. This
field is not considered fully uniform because there is potentially-significant lag on
strains towards the centre of the specimen—especially in consideration of viscous re-
sponse—as well as some negative boundary effects due the constraint of relative rota-
tion at the boundary and hence in-plane bending of each tow along the periphery. Also,
due to the fully-constrained peripheral boundary condition, out-of-plane buckling is sig-
nificantly restrained—although does eventually occur. Typically, this is beyond practi-
cal forming limits.
Chapter 6 Case study 112

a) b)

Figure 6.3. a) Bias extension test, b) Picture frame test

Bias extension tests apply force directly to the axis of principal extension, which
result in necessary shear strain (Figure 6.3a)). The idealised strain field is a series of
uniform regions that can be determined kinematically; although, the boundaries between
these regions necessitate in-plane bending. In actual testing, these strain regions are
more significantly non-uniform than in the picture-frame largely due to the free bound-
ary of the test. Fraying of the free boundary can be problematic. Within the central re-
gion of ideal strain, there is no tensile constraint along the fibre axis, and hence charac-
teristic buckling along the compressive bias occurs at lower comparative strains. De-
spite this, the test tends to have better physical analogy to forming methods with similar
free-boundaries, such as double diaphragm forming and manual lay-up. In consideration
of the approach taken here, kinematic determinacy is highly desirable, and similarly the
picture-frame approach provides a more idealised and singular strain region—albeit
quasi-uniform. For this study, quasi-uniformity is not considered analytically.

6.2.2 Multi-axial testing


In the traditional sense, multi-axial testing refers to methods capable of controlling and
measuring more than one displacement axis simultaneously. In a somewhat different
manner, the testing of the principal planar deformation in biaxial composites is neces-
sarily multi-axial by its mechanical nature. Whether woven or stitched-biaxial, the ef-
fective or direct binding between fibre families permits only multi-axial deformation—
Chapter 6 Case study 113

in this case a trellis-type deformation. Here the deformation usually referred to as


‘shearing’ is actually a combination of both a shear mode and a compaction mode.

6.3 Application of a stress model

6.3.1 Picture-frame kinematics


The outputs and control of a picture frame rig, used in a universal testing machine, are
load, displacement and time. In typical setup, the lower hinge is stationary and the upper
is attached to the hydraulically actuated load cell. By negating the effects of non-
uniform strain distribution, an analytical formulation is to be developed that is explicit
to experimental controls and output. Given this, and the geometric constraints of the
picture-frame, the formulation becomes a single trellis shear element.
Figure 4.7 depicts the basic mechanics of a picture frame shear test where, in this
case, a shear element is oriented parallel to the frame sides. Given that pull rate is the
common experimental input, then displacement G tG , and the fibre angle (M) can be
directly related to displacement:

ª tG º
M cos 1 «  cos M0 » , (6.9)
¬ 2L ¼

thus giving an explicit relationship between fibre angle and displacement. The rate of
change of fibre angle ( M ) is given by

0.5
G ­° ª tG º ½°
2

M ®1  «  cos M0 » ¾ (6.10)
2L ° ¬ 2L ¼ °¿
¯

6.3.2 Planar strains and strain-rates


For the most common of bi-directional reinforcements, the axes lie normal to one an-
other as either 0/90° (warp and weft) or +/- 45° (double bias).
Chapter 6 Case study 114

L ^`

Y
œ yy
exy ^`
œxx
M L
M

1
eyx X X X

Figure 6.4. Geometry of fibre angle; fibre strains ( and  in %); and principal system (in ().

Based on the assumption of fibre inextensibility, and using a fibre-oriented material co-
ordinate system %, the element layout becomes Figure 6.4, which shows the dimen-

sional relationships of the element inset in Figure 4.7 a), where axial strains remain
static under picture frame testing. The shear strains, based on Figure 4.2, are:

^%`
2 ^ `œ yx 2 ^ `œxy
 
J (6.11)

such that J is the engineering shear strain, recalling that the unique shear strains do not
directly affect strain energy. From Figure 6.4, we can write

^%`
J sin 2M0  2M . (6.12)

To derive strain rate ( Jij ), the first derivative of Equation (6.12) is used:

^%` d
J ªsin 2M0  2M º¼
dt ¬ (6.13)
2 cos 2M0  2M M ,

where M is given by Equation (4.27). This approach produces the same outputs when
compared with nonlinear static FEA such as the NASTRAN SOL600 solver. Subsequently,
there is no prediction of volumetric change or fibre compaction.
We consider the shear strains in corotational and persistently orthogonal frame for
each fibre family, showing total shear strain to be equal to the inter-tow sliding shear.
This is reflected by the infinitely orthotropic Biot strain, and requires different treatment
in solving the stress response. By separation of the warp and weft into individual layers
each with its own relative axis, these are able to use the convecting orthotropic model.
Chapter 6 Case study 115

The axial strains can be determined though a similar geometric approach, from Fig-
ure 6.4 where L is the length of the sides of the picture frame, this is

^` L sin 2M  L
œ yy
L (6.14)
sin 2M  1,

where the form is derived from displacement. Due to the symmetric nature of the ex-
perimental procedure, the strains in x and y are equal, hence:

^` ^`
œxx œ yy , and
(6.15)
^` ^`
œxx œ yy 0.

In Figure 4.2, the separation of layers  and  can be seen, where each is considered

with its own local axis. The axial strain rate can be determined using the derivative of
Equation (6.14):

^` d
œ yy ªsin 2M  1º¼
dt ¬ (6.16)
2M cos 2M

and similarly

^` ^`
œ xx œ yy . (6.17)

Thus all the strains and strain rates are explicit by boundary conditions, and are shown
against a particular constant ramp rate in Figure 6.5 as follows.

0.8 10
fibre-normal fibre-normal
Strainrate, dœ/dt (/s)

shear 8 shear
0.6
Strain, œ (–)

principal
6
0.4
4
0.2
2

0 0
0 100 200 300 0 100 200 300
Time, t (s) Time, t (s)

Figure 6.5. Strains and strain-rates based on kinematic predictions from the bi-laminar orthotropic model
Chapter 6 Case study 116

Note the initial gradient of zero indicates the pure simple shear of the experimental
method holding only for the infinitesimal limit.

6.3.3 Applying a stress model


By applying the power law stress model for a shear thinning fluid, as described by
Equation (6.7), the deviatoric stress response can be predicted over a typical experi-
mental testing schedule:

Figure 6.6. Deviatoric stress predictions based on application of the power law visco-plastic model

Figure 6.6 illustrates that both the fibre-normal and the shear response fields are irregu-
lar curved surfaces producing unsuitable data for characterisation.

0.2
0.18 60mm/min
0.16
20mm/min
0.14
Load (kN)

0.12
0.1
0.08
0.06
0.04
0.02
0
0 20 40 60 80 100
Displacement (mm)

Figure 6.7. Load response of Hexcel F593 resin with Toray T-300/3K fibre

In Figure 6.7 a typical Thermoset (TS) plain weave response for two successive cycles
at different constant ramp rates is shown. The obvious characteristic local maxima
greatly reduces in the second (slower) load cycle.
Chapter 6 Case study 117

As it is a TS pre-preg composite, we describe the previously shown local maxima by


supposing that some proportion of cross-linking has already occurred. These are elastic-
ity-forming bonds, so we model failure at strains equal to that of the material in its de-
sign state. This value is a commonly specified mechanical property: percentage elonga-
tion. For the particular matrix material being modelled, Hexcel F593 resin, the value is
2% axial strain, and the neat resin elastic modulus is 2.96GPa. Using the von Mises dis-
tortion energy criterion, the shear failure strain is estimated as

^%`
œii œ yld , and
(6.18)
^%`
œij 3œ yld , i z j ,

given a Poisson ratio, under the incompressibility condition, of 0.5. We further model
the subsequent elastic constant as directly proportional to the ratio of formed bonds
compared to the cured resin (denoted as N herein). Using distortion energy, the shear
modulus is similarly estimated:

^%`
E i N E spec. , and
1  (6.19)
^%`
G ij N Gspec. .
3

At initial simulation, elastic yield has been modelled using a simple exponential de-
cay model, though modelling of the described local maxima is not subject of the current
investigation. Elemental stress/strain response is shown based on this model:

0.25 0.1
Abs. viscous stress, (MPa)
Abs. elastic stress, (MPa)

Fibre-normal
0.2 0.08
Shear
0.15 0.06

0.1 0.04

0.05 Fibre-normal 0.02


Shear
0 0
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
Abs. strain, (œ) Abs. strain, (œ)

Figure 6.8. Estimated elastic stress coupled with a simplified temporal relaxation model
Chapter 6 Case study 118

6.3.4 Modelling through-thickness strain


The through-thickness strain, œzz, is kinematically determined based on incompressibil-
ity and hence using conservation of volume. At such low hydrostatic pressure, this is
typically valid. The formulation follows:

^%` Ainitial
œ11 1
Afinal
2 L2 sin M0 cos M0
1 (6.20)
2 L2 sin M cos M
sin 2M0 csc 2M  1,

where Ainitial and Afinal are the effective surface areas of the specimen. Because there is
no out-of-plane shear and a plane element is used,

^` ^%`
œzz œ11. (6.21)

By taking the first derivative of Equation (6.20), the through-thickness strain rate can be
calculated:

^` d
œ zz ªsin 2M0 csc 2M º¼
dt ¬ (6.22)
2sin 2M0 csc 2M cot 2M M.

6.3.5 Calculation of boundary forces


To account for the substantial increase in material thickness as a result of local x and y
material compression, the normal strain (œzz) is used to estimate thickness (tn), allowing
calculation of corrected external force (Ps) due to shear resistance:

V L tn 1  œzz
Ps , (6.23)
2

where V Sij for deviatoric contribution. Similarly, the reaction force due to fibre-

normal stresses (Pfn) is given by

§ cos 2 M ·
Pfn V L tn 1  œzz ¨  sin M ¸ , (6.24)
© sin M ¹

allowing explicit prediction of reaction forces.


Chapter 6 Case study 119

6.4 Parameter identification technique

6.4.1 Pull-rate profiling


The strain relationships developed in the previous section stipulate that a significant
contribution to internal strain energy will come from fibre-normal compression, and
hence the same contribution should constitute part of the reaction force measured as a
response to material testing. As the aim is parameter identification, the most orthogonal
experimental set is desired.
Given that shear strain and its rate of change are the most influential factors, we in-
tend to control boundary conditions such that they are linearised. In a manner similar to
Equation (6.13) and (6.16), the variation of displacement given a constant shear strain-
rate can be obtained:

­ ª 1 º ½
G 2 L ®cos «M0  sin 1 Jt »  cos M0 ¾ . (6.25)
¯ ¬ 2 ¼ ¿

This can be solved over any displacement range given a set of shear strain-rates to pro-
duce a loading profile.

100 30
Displacement,  (mm)

Ramp rate (mm/min)

80 25
0.0625 mˆ/s 20
60 0.125 mˆ/s
15
0.25 mˆ/s
40
0.5 mˆ/s 10
20 1.0 mˆ/s 5
2.0 mˆ/s
0 0
0 2000 4000 6000 0 5 10 15
Time, t (s) Block number

Figure 6.9. Displacement profiles and ramp rate blocks for linearisation of single-axis response

As shown in Figure 6.9, in implementing this testing schedule, along with a series of
varying shear strain-rates, the displacement profile must be discretised into ramp rate
blocks for input into the testing control system.
Chapter 6 Case study 120

Figure 6.10. Deviatoric stress variation for constant shear-rate testing schedule

Figure 6.10 demonstrates the resulting shear plane deviatoric stress component, which
now exhibits planar linearity. Note that a further advantage to this technique is in the
elimination of viscosity changes, particularly considering the material is modelled as a
non-Newtonian shear-thinning fluid.

0.8 20
fibre-normal fibre-normal
Strainrate, dœ/dt (/s)

0.6 shear 15 shear


Strain, œ (–)

principal
0.4 10

0.2 5

0 0
0 500 1000 1500 0 500 1000 1500
Time, t (s) Time, t (s)

Figure 6.11. Strain and strain rate variation with time for constant shear strain rate testing schedule

The linearisation of the shear strain-rate can be seen in Figure 6.11.


As the force response is composed of two primary deformation modes, fibre-normal
contributions can be given the same treatment,

­ ª1 º ½
G 2 L ®cos « sin 1 œt
  1 »  cos M0 ¾ . (6.26)
¯ ¬2 ¼ ¿

Equation (6.26) predicts a requirement of an infinite ramp rate for the first infinitesimal
increment as the picture frame has no initial component of compaction.
Chapter 6 Case study 121

Figure 6.12. Deviatoric stress variation for constant shear-rate schedule

A planar response is predicted for the fibre-normal contribution (Figure 6.12), and simi-
larly, the strain rates demonstrate that now, as fibre-normal strain-rate remains constant,
the shear rate rapidly reduces, as shown in Figure 6.13.

0.8 6
fibre-normal fibre-normal
5
Strainrate, dœ/dt (/s)

0.6 shear shear


Strain, œ (–)

principal 4

0.4 3

2
0.2
1

0 0
0 2000 4000 6000 0 2000 4000 6000
Time, t (s) Time, t (s)

Figure 6.13. Strain and strain rate variation with time for testing of constant fibre-normal strain rate

6.4.2 Viscous relaxation


In determining the transient response coefficients of a rate dependent model, an effec-
tive approach is to take relaxation data at the end of each pull test. During ‘hold’ blocks,
every variable becomes fixed outside of the decaying stress. An example result for cy-
clic testing and decay of the experimental force is shown in Figure 6.14.
Chapter 6 Case study 122

120 0.7

100 0.6

Reaction Force (kN)


Displacement (mm)
0.5
80
0.4
60
0.3
40
0.2
20 0.1

0 0
0.0 500.0 1000.0 1500.0 2000.0 Time (s)

Figure 6.14. Parametric plot showing force response and displacement profile (dashed) over time.

The results suggest exponential decay from visco-elastic behaviour. A simple model of
exponential decay can be applied to model this:

t t peak
V V peak  V ref enatural
Odecay
 V ref , (6.27)

where enatural is Euler’s constant and tpeak is the time at which decay begins. Here, stress
can be decayed against a value exceeding a baseline.

1
Reaction Force
0.8 Exponential decay
Reaction force (kN)

0.6

0.4

0.2

0 Time (s)
0.0 50.0 100.0 150.0 200.0

Figure 6.15. Precise fit between exponential decay model and recorded relaxation stress response.

By isolating the first decay block, a simple iterative approach was applied to determine
the coefficient of decay, Odecay, and Figure 6.15 shows the model fit to be sound

6.4.3 Cyclic inference


Given the proposed model involves two embedded stress mechanisms within the micro-
structure of the matrix, the experimental load constitutes superimposed components. For
effective explicit characterisation and determination of model parameters, a technique is
Chapter 6 Case study 123

suggested based on an extrapolation of the micro-structural constituents. If it were that a


low concentration of polymeric cross-links is likely to fracture internally in a mode of
progressive rupture, the second cycle response would represent purely the viscous flow
resistance. The differential load between first and second cycle would represent the
mechanism that suffered damage—although there may also be a second-cycle propor-
tionate constituent based on other modes of flow conditioning.
The inference was applied to cyclic results from a constant fibre-normal strain-rate
test to demonstrate a number of the analysis techniques, the ‘corrected’ force was in-
dexed to the thickness invariant and a corrected cyclic differential indicated first cycle
damage.
0 -0.1 -0.2 -0.3 -0.4 -0.5
30 100
1st Cycle Load profile
25 1st corrected
2nd Cycle 80
2nd corrected

Displacement (mm)
20 Cycle differential
60
Load (kN)

Differential poly.
15
40
10

5 20

0 0
0 -0.1 -0.2 -0.3 -0.4 -0.5
Control (m H) - Linear with time

Figure 6.16. Resultant force from first and second cycle plotted with displacement profile.

Figure 6.16 demonstrates the initial first and second cycle normalisation according to a
thickness indexing function related to the through-thickness strain formulation, noting
that the incompressible fluid assumption is invariant to rate and stress. The force predic-
tion formulations developed earlier indicate a linear proportionality of force with re-
spect to thickness changes. The first and second corrected cycles were used to create a
differential force component, where in the ideal case this isolates the cross-link contri-
bution to reaction force. Similarly, the second cycle response represents the pure vis-
cous fluid response.
Chapter 6 Case study 124

30
1st Cycle
25 1st corrected
2nd Cycle
2nd corrected
20 Cycle differential
Load (kN)

Differential poly.
15

10

0
0 10 20 30 40 50 60 70 80 90 100
Displacement (mm)

Figure 6.17. Previous figure results for constant fibre-normal strain rate plotted against actuator dis-
placement

Figure 6.17 shows the same result in the displacement domain, which has a signifi-
cantly different shape due to the transient load profile used. This normalised and iso-
lated alleged stress mechanism indicates good qualitative agreement with the model out-
lined previously. Remembering the varying kinematic relations, a fairly linear initial
slope is followed by what resembles characteristic yield onset, proceeding with a decay-
ing stress.

6.4.4 Rate invariant decomposition


Given the reaction data are isolated into their respective mechanisms, normalised
against thickness changes, and are produced using a testing schedule to linearise the fi-
bre-normal response, the viscous fluid component can now be decomposed into the
shear and fibre-normal contributions. This is done using the rate-invariant ratio derived
using the kinematics predictions, accepted rheological stress model and proposed treat-
ment of fibre-family orthogonality.
Chapter 6 Case study 125

0.8

Ratio (–)
0.6
Fibre-normal
0.4 Shear

0.2

0
0 20 40 60 80 100
Displacement,  (mm)

Figure 6.18. Rate-invariant ratios of force contribution

Figure 6.18 depicts the proposed rate-invariant ratios. The valuable quality of this ratio
vector lies in its independence to material parameters. Generated solely based on kine-
matic qualities, it is highly suited to decomposition of the experimental results for effec-
tive explicit parameter identification. Subsequently, the ratio is used to break the vis-
cous component into its inversely predicted shear and fibre-normal contribution—
allowing direct comparison with kinematically simulated force responses.

10 10
Visc. Compaction, P (N)
Visc. Shear, P (N)

8 8

6 6

4 4

2 2

0 0
0 50 100 0 50 100
Displacement,  (mm) Displacement,  (mm)

15 20 Analytical
Experimental
Total Visc., P (N)
Elastic, P (N)

15
10
10
5
5

0 0
0 50 100 0 50 100
Displacement,  (mm) Displacement,  (mm)

Figure 6.19. Reaction force components from the kinematic model compared to the analytically decom-
posed experimental force contributions.
Chapter 6 Case study 126

The comparison in Figure 6.19 shows good model fit in viscous shear beyond the initial
displacement. This is principally owing to the application of infinitely orthotropic
kinematics, which in our analytical functions is equivalent to the intrinsic Biot tensor.
We make some additional notes to this analysis: the initial variation is expected as
in this instance the required ramp rate cannot be practically attained. Note that only the
fibre-normal test has been processed thus far. The fibre-normal component shows good
correlation over the entire range, which is as expected, considering the particular ex-
perimental result was designed around analysis of this component of stress. There is a
significant divergence towards the end as the angle M approaches the physical fibre
locking angle. It is also worth noting that the proposed elasticity damage model shows
good initial correlation, whereby an iterative approach was used to estimate the specu-
lated ‘bond formation ratio’, N, in order to identify the elastic moduli. The resulting
value was approximately 0.25%, hence suggesting that 0.25% of the reaction had al-
ready formed in the uncured material. Most promisingly, the yield point prediction
based on manufacturer’s specifications of the neat resin in its design state correlates
precisely with the experimental peak. Lastly, it is apparent that the decay function is
inadmissible.

6.5 Chapter summary

In this case study an analytical implementation for the infinitely orthotropic kinematic
condition has been used to model an uncured pre-impregnated polymeric composite
with continuous fibre reinforcement. Using Spencer’s kinematic equations, shown to be
equal to SOCM measures under the infinitely orthotropic condition, an explicit formula-
tion equivalent to the orthotropic Biot strain was used in its description. Based on this,
the parameter identification of the viscous stress model was achieved using orthogonal-
ised experimental schedules and a series of novel experimental techniques. The results
indicate a sound comparison between model predictions and force response from the
trellis shear test, and offer validation of the intrinsic strain measure that was previously
developed.
Chapter 7
Overview and conclusion
In this, the final chapter, the conclusions of the thesis are drawn from the original con-
tributions and described in association with the objective. By offering a classification of
the presented work the general nature of the theory is apparent, though we also observe
the modesty of the acquired capabilities. A summary of key formulations is given, and
in reflecting upon our notation some underlying themes are tied together. Finally the
formal conclusions and continuing investigation are suggested before a closing remark.

7.1 Overview

7.1.1 Classification of work


If we consider a family of possible displacement state tensors, whereby there exists an
encompassing set classified as nonlinear and of orthotropic formulation, then linear and
isotropic formulations may be considered as subsets.

Œ,RŒ,UŒ e,Z: Linear isotropic strain and rotation tensors


œ,\: Linear orthotropic tensors
E,Œ: Isotropic and orthotropic Green–Lagrange strain
E,RE,UE Z œ,\
eZ
RE,UE: Isotropic polar decomposition tensors
A = B ˆC B
A
C
A,B,C D D
RŒ,UŒ: Intrinsic polar decomposition tensors

Figure 7.1. Classification of some conventional deformation tensors against proposed material dependent
measures. Linear is considered a subset of nonlinear and isotropic a subset of orthotropic.
127
Chapter 7 Summary and conclusions 128

Figure 7.1 shows the nesting of the conventional infinitesimal state tensors, e and Z, as
the union between isotropic and linear (infinitesimal or arotational). It naturally follows
that the anisotropic formulations can encompass these due to their capacity to degener-
ate to the orthotropic and isotropic forms.
As it was mentioned in the introduction, the classical mechanics are pristine in deal-
ing with almost all of the common analysis scenarios. Nonetheless, in the literature re-
view it was demonstrated that there is an area of inaccuracy that has been, at least sec-
ondarily, the subject of many investigations to do with strongly orthotropic continua.

Isotropic Cauchy, Green, etc.


Š   \_

Weak
Weakly
orthotropic

Strong SOCM

Infinite IFRM, structural tensors etc.


problem ‰
@@@^@_

Figure 7.2. An illustrative summary of the contribution: strongly orthotropic continuum mechanics is
proposed to suffice the spectrum of kinematically exact finite deformation analysis.

As has been said, and as suggested in Figure 7.2, SOCM is propounded to provide a
mechanism to fill this modest region of inexactitude, a region that tends to proliferate
only at high levels of orthotropy or anisotropy and most commonly at finite strains.

Isotropic
Š   \_

Weak

The intrinsic field tensors

Strong
Infinite
problem ‰
@@@^@_

Figure 7.3. As they have been shown to generalise the response of the conventional classical tensors, and
on acceptance of Figure 7.2, the IFTs are capable of being exact for all classes of problem shown.
Chapter 7 Summary and conclusions 129

While bringing attention to the many caveats throughout the document, it is reasonable
to say the IFTs, as is shown in Figure 7.3, offer a general and unique ability to describe
finite deformation of isotropic, orthotropic and anisotropic continua under, at the least,
the studied conditions of linear elasticity and finite and distortive plane deformations.

7.1.2 Summary of equations


To summarise the mathematical formulations that provide realisation of SOCM, the
most necessary equations are given here. To put their form in context, the ILST, OLST
and ALST developed in Chapter 3 are shown here in succession:

Linear strain :
ILST
1 K L 1 § wui wu j ·
e
2
’u  ’u , eij ¨ 
2 ¨© wX j wX i
¸¸
¹

OLST
K L
œ ’u  ’u <r a  ar b
§ wui wu j · § wu wu j · dij  d ji
¨¨  ¸¸ dij sin Tij  d ji cos Tij  ¨¨ i  ¸¸ sin 2Tij
2 2
œij
© wX j wX i ¹ © wX i wX j ¹ 2

ALST
K L K K
æ ’u  ’u <r a  ’u<I 1 rb  1 ’u<I r c
§ wu wu j · a wu b wu j c
æij ¨¨ i  ¸¸ rij  i rij  rij
© wX j wX i ¹ wX i wX j

Essential to this, the constitutive equations providing linear elasticity are summarised:

Linear elasticity :
Constitutive model
œij SijklV klœ , œ SV œ , ^%`
D f ^ `S
%

Compliance tensor

^%`
G ik G jl 1 Q ji  G ijG klQ lj
Sijkl
E l
Chapter 7 Summary and conclusions 130

In solving the relaxed stress for finite rotation problems, the corotational procedure of
Chapter 4 maintains material orthogonality by use of the intrinsic linear measure of ro-
tation. Noting internal definitions are not repeated, the enabling equation is highlighted:

Material principal corotation :


Material principal rotation tensor
wui 2 wu j 2 § wu wu j · dij  d ji
\ iu j
wX j
s T ji d ji  c 2 T ji dij 
wX i
s Tij dij  c 2 Tij d ji  ¨ i 
¨ wX wX ¸¸ s 2Tij
2
© i j ¹

Finally, the boxed equations of Chapter 5 are gathered here, noting that the omissions to
these equations—such as the general definition FTF = U2—are of equal significance to
the novelty of the IFT proposition. They are as follows:

Finite deformation :
Orthotropic right stretch tensor
FTF U ΠT
U Π, where
ª U xxŒ rxya xˆ  rxyb U xxŒ  U yyŒ º
U Œ
« »
« ryxa xˆ  ryxb U yyŒ  U xxŒ U Œ
»
¬ yy ¼

Orthotropic Biot Strain Tensor (OBST)


HΠU Π I, H ijΠU ijΠ G ij

Orthotropic Green–Lagrange strain tensor (OGST)

Œ=
2 «¬
U  I »¼ , Œij
1ª Œ 2 º
U imU mj  G ij
1 Œ Œ
2

This summary, especially in the linear formulations and its notations, should provide a
primer to the reflections on our usage and terminology hereto.

7.1.3 Reflections on notation


In the introduction to this dissertation we noted an intention to exploit the less-often
used denotation of e rather than epsilon, H, for the small strain tensor. This decision
owed partially to the fact that the use of e would be conducive and consistent to usage
of the ligatures ethel, œ, and ash, æ for the coming orthotropic and anisotropic descrip-
tions. The initial interpretation to this was something of a drawback considering that
Chapter 7 Summary and conclusions 131

epsilon in mathematics typically refers to a value that is very small, such as error or
higher-order terms. This, of course, compliments its use for small strain, as it is the in-
finitesimal or first-order approximation to the Eulerian finite strain (Almansi–Euler ten-
sor) or Lagrangian finite strain (Green–Lagrange tensor). Furthermore, there is a recog-
nisable convention in usage of lower-case graphemic Roman characters to refer to Eule-
rian metrics and of the upper-case to refer to Lagrangian metrics; and where the infini-
tesimal frame is concerned, there is no difference between these, thus it follows that we
ideally avoid use of a Roman grapheme altogether as is done with epsilon.
In the study presented on finite deformation (Chapter 5), it was supposed that the
new intrinsic Biot tensor might now be preferential to the Lagrangian measure of strain.
This was by consideration that the conventional co-rotated and Biot approaches would
suffer from significant numerical locking even under moderate finite strain, and that the
proposed formulations would not. Remembering that the orthotropic Biot strain is a
convected linear Eulerian measure, it is inferred that maybe linear strains are not limited
to infinitesimal deformation after all, and that in fact the orthotropic linear strain meas-
ure grows into a co-rotated intrinsic Eulerian frame. This idea finds support in the litera-
ture, and it is given as a corollary that the arotational forms of the orthotropic and ani-
sotropic cases complete the argument. In that sense, it is suggested that the abandon-
ment of epsilon—suggesting very small—may not have been so disadvantageous after
all. As an upshot to this, beyond the infinitesimal limit it becomes intrinsic-Eulerian and
so the lower-case graphemes—e, œ and æ—reveal an unexpected consistency to the ear-
lier noted convention. The final reflection on this is to say that perhaps we can even
draw distinction between H and e:

H, the extrinsic infinitesimal measure of strain, is limited to that scope because it cannot
with any accuracy describe the finite strains of orthotropic continua. In contrast, the in-
trinsic version—the reduced-assumption and rationally derived ILST—e is limited in its
definition to isotropic media: the scope to which it maintains accuracy. The areas of in-
adequacy of H are instead made-good by the proposed intrinsic versions, œ and æ—the
OLST and ALST respectively—and so the abandonment of the infinitesimal limitation
inferred by the notation ‘epsilon’ is put forward for thought.
Chapter 7 Summary and conclusions 132

7.2 Conclusion

7.2.1 Concluding statements


By asserting that multiple geometrically compatible deformation modes can exist, and
that these deformation modes have strain fields made determinate by a dependency on
material properties, a novel linear strain tensor has been proposed. The formulation of
asymmetric yet determinate strain states has been achieved by deriving the strain tensor
using a dependency on the material orientation and mechanical properties in maintain-
ing moment equilibrium. By referring to the conventional infinitesimal Cauchy strain
tensor as the isotropic linear strain tensor, the developed strain metric—the orthotropic
linear strain tensor—has been shown to replace the ILST in any application due to its
ability to degenerate into the conventional form. Furthermore, a novel rotation tensor
has been presented that gives change in principal material direction when applied to
orthotropic materials. This metric, the so called material principal rotation tensor, is re-
quired to solve the second-order components of strain correctly, first by means of a
corotational procedure.
These distinct yet compatible modes between material families necessitate the spa-
tial decomposition of any multi-axial continuum—such as woven fibre-reinforced com-
posites—into coincident continua. Thus for materials with inherent principal direction-
ality, the problems of non-orthonormal material basis vectors following large shear de-
formation are avoided by the use of persistently orthogonal material coordinate systems.
As part of the established continuum mechanics, a novel extended orthotropic compli-
ance tensor has been developed to be compatible with the OLST, where material frame-
indifference can be shown on the extended Mohr plot utilising an aspectual strain metric
and the necessary Eigen analysis. These were used for a two-dimensional finite element
implementation using the state tensors and Jacobian based mapping matrices, allowing a
conventional linear solution procedure.
Considering that strain metrics such as the infinitesimal Cauchy strain, Biot,
Cauchy–Green and Green–Lagrange strain are asserted to be isotropic, the validity of
the OLST leads to the proposition of a class of physical tensors described as the Intrin-
sic Field Tensors that can reasonably generalise any conventional strain tensor along
Chapter 7 Summary and conclusions 133

with their specialised counterparts such as the structural and fibre-director convected
tensors. Pursuant to this, the intrinsic stretch tensors have been derived by releasing the
forced symmetry on the polar decomposition of the deformation gradient, and thence
replacing that determining factor with an objective equilibrium function borrowed from
the earlier derivation. It follows that all the initially mentioned tensors are reformulated
so as to reflect what is derived to be the rational kinematic of orthotropic deformation,
regardless of whether they be Lagrangian, Eulerian or corotational. To this end, we ul-
timately surmise the following:
In this dissertation a continuum mechanics theory has been established for strongly
orthotropic media that is propounded to provide a kinematically exact solution over a
wide range of strain measures. It should be adequate to say that the strain tensors de-
rived benefit from the description of being rational and generalised, as they are all capa-
ble of reducing to conventional—what we call isotropic—solutions. In addition to this,
the theory has been extended to describe multi-axial materials such that the underlying
problems in composite forming could be addressed, and the requisite finite element
treatment was provided such that various demonstrative numerical examples could be
executed. From all this, as it serves to provide, it is sufficient to conclude that the objec-
tive of this thesis has been met.

7.2.2 Continuing investigation


Throughout the document a number of allusions and footnotes regarding the necessary
further developments of the work have been given. It is apparent that continuing inves-
tigation is required such that the established theory has with it the theoretical infrastruc-
ture for practical numerical implementation into a wider range of problems. This in-
cludes such things as a few essential generalised plate and shell elements [126], which
demand considerations into Kirchhoff–Love [242] and Mindlin–Reissner [243] bending
solutions, development of Q6 elements [244] and nonlinear three dimensional shells
having out-of-plane rotational degrees of freedom [245], and the many suitable numeri-
cal procedures required for implementable forms of the skew decompositions that are
compatible with mixed definition problems [246]. The intrinsic field tensor theory must
also be expanded to present the many rate, spin and vorticity measures [247] that are
demanded in rheology and visco-elastic/plastic simulation [248].
Chapter 7 Summary and conclusions 134

This work was principally completed as part of continuing research that aims to cre-
ate the necessary theory, method and framework required for the accurate prediction of
forming of advanced composite materials. This will ultimately require the aforemen-
tioned development of the strongly orthotropic continuum theory into highly material
and geometrically nonlinear finite elements.

7.2.3 Closing remark


Conceptually growing from a single comment by Augustin-Louis Cauchy, the basis
used to achieve this thesis lies in the elements of approach—geometric strain indetermi-
nacy, persistent material principal orthogonality and the intrinsic frame. Through ra-
tional derivation, development and demonstration of the methods and formulation, the
initially unsubstantiated bases and conjecture are put forward as theory: what we call
strongly orthotropic continuum mechanics.
Nonetheless, if we go so far as to presume that some kind of meaningful and
worthwhile understanding has been acquired, then, above all, it is remarkable to con-
sider that even 150 years afterwards we are still catching up to the thinking of those
great mechanicians and natural philosophers.

***
REFERENCES
1. Lagrange JL, Mécanique analytique, nouvelle édition. Courcier, Paris 1811.
2. Cauchy AL, Sur la condensation et la dilatation des corps solides. Ex. de Math 1827;
2:60–69.
3. Cauchy AL, De la pression ou tension dans un corps solide. Ex. de Math 1827; 2:42-56.
4. Cauchy AL, Sur les équations qui expriment les conditions d'équilibre on les lois du
mouvement intérieur d'un corps solide, élastique ou non élastique. Ex. Math 1828; 3:160-
187.
5. Ricci-Curbastro G, Di alcune applicazioni del calcolo differenziale assoluto alla teoria
delle forme differenziali quadratiche binarie e dei sistemi a due variabili. Veneto Ist. Atti
(7) IV. 1336-1364., 1893.
6. Green G, Mathematical Papers of the Late George Green. Macmillan and co., 1871.
7. Autonne L, Sur les groupes lineaires, reels et orthogonaux. Bulletin de la Societe Mathe-
matique de France 1902; 30:121-134.
8. Truesdell C, Toupin RA, The Classical Field Theories. Principles of Classical Mechanics
and Field Theory, Encyclopedia of Physics, ed. S. Flugge. Vol. III/1. Springer Verlag:
Berlin, Gottingen, Heidelberg, 1960.
9. Biot MA, Mechanics of Incremental Deformations: Theory of Elasticity and Viscoelastic-
ity of Initially Stressed Solids and Fluids, Including Thermodynamic Foundations and
Applications to Finite Strain. Wiley, 1965.
10. Cauchy AL, Mémoire sur les systèmes isotropes de points materiels. Institut National de
France, 1850.
11. Xiao H, Bruhns OT, Meyers A, A Natural Generalization of Hypoelasticity and Eulerian
Rate Type Formulation of Hyperelasticity. Journal of Elasticity 1999; 56(1):59-93.
12. Neff P, Fischle A, Münch I, Symmetric Cauchy stresses do not imply symmetric Biot
strains in weak formulations of isotropic hyperelasticity with rotational degrees of free-
dom. Acta Mechanica 2008; 197(1):19-30.
13. Baker AA, Dutton S, Kelly DW, Composite Materials for Aircraft Structures. AIAA,
2004.

135
References 136

14. Furukawa T, Michopoulos JG, Design of Multiaxial Tests for Characterizing Anisotropic
Materials. International Journal for Numerical Methods in Engineering 2007 (in print):25.
15. Clayton G, Howe C, Cost as a driver in advanced composites aerospace research. Pro-
ceedings of the I MECH E Part L Journal of Materials:Design and Applications 2004;
218:79-85.
16. Shuart MJ, et al., Automated fabrication technologies for high performance polymer
composites. Composites Fabrication. Vol. 14, no. 8, pp. 24-30. Aug. 1998 1998.
17. Verrey J, Wakeman MD, Michaud V, Manson JAE, Manufacturing cost comparison of
thermoplastic and thermoset RTM for an automotive floor pan. Composites Part A: Ap-
plied Science and Manufacturing 2006; 37(1):9-22.
18. Mills A, Automation of carbon fibre preform manufacture for affordable aerospace appli-
cations. Composites Part A: Applied Science and Manufacturing 2001; 32(7):955-962.
19. Mahieux CA, Cost effective manufacturing process of thermoplastic matrix composites
for the traditional industry: the example of a carbon-fiber reinforced thermoplastic fly-
wheel. Composite Structures 2001; 52(3-4):517-521.
20. Pantelakis SG, Baxevani EA, Optimization of the diaphragm forming process with regard
to product quality and cost. Composites Part A: Applied Science and Manufacturing
2002; 33(4):459-470.
21. Potluri P, Atkinson J, Automated manufacture of composites: handling, measurement of
properties and lay-up simulations. Composites Part A: Applied Science and Manufactur-
ing 2003; 34(6):493-501.
22. Belytschko T, Liu WK, Moran B, Nonlinear finite elements for continua and structures.
Wiley: New York, 2000.
23. Spencer AJM, Continuum Theory of the Mechanics of Fibre-reinforced Composites.
Springer Verlag: Wien, New York, 1984.
24. Spencer AJM, Deformations of Fibre-reinforced Materials. Clarendon Press: Oxford,
1972.
25. Rogers TG, Deformations of strongly anisotropic materials. Rheologica Acta 1977;
16(2):123-133.
26. Sun H, Pan N, Postle R, On the Poisson's ratios of a woven fabric. Composite Structures
2005; 68(4):505-510.
27. Spencer AJM, Fibre-streamline flows of fibre-reinforced viscous fluids. European Jour-
nal of Applied Mathematics 1997; 8(02):209-215.
References 137

28. Rogers TG, Squeezing flow of fibre-reinforced viscous fluids. Journal of Engineering
Mathematics 1989; 23(1):81-89.
29. Spencer AJM, Theory of fabric-reinforced viscous fluids. Composites Part A: Applied
Science and Manufacturing 2000; 31(12):1311–1321.
30. Spencer AJM, Some results in the theory of non-Newtonian transversely isotropic fluids.
Journal of Non-Newtonian Fluid Mechanics 2004; 119(1-3):83-90.
31. Cleja-Tigoiu S, Anisotropic and dissipative finite elasto-plastic composite. Universitae
Politecnico di Torino 2000; 58(1):69-83.
32. Page J, Wang J, Prediction of shear force and an analysis of yarn slippage for a plain-
weave carbon fabric in a bias extension state. Composites Science and Technology 2000;
60(7):977-986.
33. Long AC, Souter BJ, Robitaille F, Rudd CD, Effects of fibre architecture on reinforce-
ment fabric deformation. Plastics, Rubber and Composites 2002; 31(2):87-97.
34. Buet-Gautier K, Boisse P, Experimental analysis and modeling of biaxial mechanical be-
havior of woven composite reinforcements. Experimental Mechanics 2001; 41(3):260-
269.
35. Smith P, Rudd CD, Long AC, The effect of shear deformation on the processing and me-
chanical properties of aligned reinforcements. Composites Science and Technology 1997;
57(3):327-344.
36. Long AC, Process modelling for liquid moulding of braided preforms. Composites Part A
2001; 32(7):941-953.
37. McEntee SP, Obradaigh CM, Large deformation finite element modelling of single-
curvature composite sheet forming with tool contact. Composites Part A: Applied Science
and Manufacturing 1998; 29(1-2):207-213.
38. Gorczyca JL, Sherwood JA, Liu L, Chen J, Modeling of Friction and Shear in Thermo-
stamping of Composites - Part I. Journal of Composite Materials 2004; 38(21):1911-
1929.
39. Liu L, Chen J, Gorczyca JL, Sherwood JA, Modeling of Friction and Shear in Thermo-
stamping of Composites - Part II. Journal of Composite Materials 2004; 38(21):1931-
1947.
40. Liu L, Chen J, Li X, Sherwood J, Two-dimensional macro-mechanics shear models of
woven fabrics. Composites Part A: Applied Science and Manufacturing 2005; 36(1):105-
114.
References 138

41. Nguyen M, Herszberg I, Paton R, The shear properties of woven carbon fabric. Compos-
ite Structures 1999; 47(1-4):767-779.
42. Clifford MJ, Harrison P, Long AC. Can it be made? Predicting the Formability of Textile
Composite Components. Presented in XXI International Congress of Theoretical and Ap-
plied Mechanics. 2004. Warsaw, Poland.
43. Sharma SB, Sutcliffe MPF, Draping of woven fabrics: Progressive drape model. Plastics,
Rubber and Composites 2003; 32:57-64.
44. Rogers C, Schief WK, The Kinematics of the Planar Motion of Ideal Fiber-Reinforced
Fluids: An Integrable Reduction and Bäcklund Transformation. Theoretical and Mathe-
matical Physics 2003; 137(2):1598-1608.
45. Schief WK, Rogers C, The Kinematics of Fibre-Reinforced Fluids. An Integrable Reduc-
tion. Q J Mechanics Appl Math 2003; 56(4):493-512.
46. Harrison P, Clifford MJ, Long AC, Shear characterisation of viscous woven textile com-
posites: a comparison between picture frame and bias extension experiments. Composites
Science and Technology 2004; 64(10-11):1453-1465.
47. Harrison P, Clifford MJ, Long AC, Rudd CD, Constitutive modelling of impregnated
continuous fibre reinforced composites Micromechanical approach. Plastics, Rubber and
Composites 2002; 31:76-86.
48. Harrison P, Clifford MJ, Long AC, Rudd CD, A constituent-based predictive approach to
modelling the rheology of viscous textile composites. Composites Part A: Applied Sci-
ence and Manufacturing 2004; 35(7-8):915-931.
49. Aronsson A, Design, Modeling and Drafting of Composite Structures. Master's thesis,
Luleå University of Technology 2005.
50. Wang J, Paton R, Page JR, The draping of woven fabric preforms and prepregs for pro-
duction of polymer composite components. Composites Part A: Applied Science and
Manufacturing 1999; 30(6):757-765.
51. Conchúr M, O'Brádaigh R, Pipes B, Mallon PJ, Issues in diaphragm forming of continu-
ous fiber reinforced thermoplastic composites. Polymer Composites 1991; 12(4):246–256.
52. Lim T-C, Ramakrishna S, Modelling of composite sheet forming: a review. Composites
Part A: Applied Science and Manufacturing 2002; 33(4):515-537.
53. Mota Soares CM, Moreira de Freitas M, Araujo AL, Pedersen P, Identification of mate-
rial properties of composite plate specimens. Composite Structures 1993; 25(1-4):277-
285.
References 139

54. Rikards R, Identification of Mechanical Properties of Laminates. Modern Trends in Com-


posite Laminates Mechanics. Vienna, Springer-Verlag(Austria), 2003. 2003:181-225.
55. Zhenhan YAO, Shisheng QU, Identification of the Material Parameters of Laminated
Plates. ?????? 2000; 5(1):1-4.
56. Pickett AK, Review of Finite Element Simulation Methods Applied to Manufacturing and
Failure Prediction in Composites Structures. Applied Composite Materials 2002; 9(1):43-
58.
57. Rogers TG, Rheological characterization of anisotropic materials. Composites 1989;
20(1):21-27.
58. Bhattacharyya D, Composite Sheet Forming, ed. R.B. Pipes. Elsevier: Amsterdam, The
Netherlands, 1997.
59. McGuinness GB, Obradaigh CM, Development of rheological models for forming flows
and picture-frame shear testing of fabric reinforced thermoplastic sheets. Journal of Non-
Newtonian Fluid Mechanics 1997; 73(1-2):1-28.
60. McGuinness GB, Obradaigh CM, Characterisation of thermoplastic composite melts in
rhombus-shear: the picture-frame experiment. Composites Part A: Applied Science and
Manufacturing 1998; 29(1-2):115-132.
61. Youssef M, Boisse P, Soulat D, Simulation de la mise en forme des composites à renforts
tissés: Approche continue.
62. Morozov EV, Mechanics and analysis of fabric composites and structures. AUTEX Re-
search Journal 2004; 4(2):60-71.
63. Mohammed U, Lekakou C, Dong L, Bader MG, Shear deformation and micromechanics
of woven fabrics. Composites Part A: Applied Science and Manufacturing 2000;
31(4):299-308.
64. Eischen J, Bigliani R, Cloth Modeling and Animation, chapter 4, Continuum versus parti-
cle representations. AK Peters 2000:79–122.
65. Breen DE, House DH, Wozny MJ, Predicting the Drape of Woven Cloth Using Interact-
ing Particles.
66. Do D, John S, Herszberg I, 3D deformation models for the automated manufacture of
composite components. Composites Part A: Applied Science and Manufacturing 2006;
37(9):1377-1389.
67. Lin H, et al., Predictive modelling for optimization of textile composite forming. Com-
posites Science and Technology 2007; 67(15-16):3242-3252.
68. Lussier D, Shear characterization of textile composite formability'. 2002, MS. Thesis.
References 140

69. Lussier D, Chen J, Material characterization of woven fabrics for thermoforming of com-
posites. Journal of Thermoplastic Composite Materials 2002; 15(6):497-509.
70. Lussier DS, Chen J, Sherwood JA, Viscosity-based models for shear of
glass/thermoplastics fabrics. Proceedings of the 5th International ESAFORMConference
on Material Forming, Krakow 2002:283–6.
71. Chen J, Lussier DS, Cao J, Peng XQ, Materials Characterization Methods and Material
Models for Stamping of Plain Woven Composites. Material Forming Processes 2003.
72. Chow S, Lussier DS, Chen J, Shear and friction response of Co-mingled glass polypro-
pylene fabrics during stamping. Sixteenth Technical Conference of the American Society
for Composites 2001:2001.
73. de Luca P, Lefebure P, Pickett AK, Numerical and experimental investigation of some
press forming parameters of two fibre reinforced thermoplastics: APC2-AS4 and PEI-
CETEX. Composites Part A: Applied Science and Manufacturing 1998; 29(1-2):101-110.
74. Yu X, et al., Composite forming simulations - from fundamental research to industrial
applications. 50th International SAMPE Symposium & Exhibition, Long Beach, Califor-
nia 2005.
75. Dong L, Lekakou C, Bader MG, Solid-mechanics finite element simulations of the drap-
ing of fabrics: a sensitivity analysis. Composites Part A: Applied Science and Manufac-
turing 2000; 31(7):639-652.
76. Willems A, Lomov SV, Vandepitte D, Verpoest I, Thermoforming of a woven textile
thermoplastic composite: a preliminary material sensitivity study.
77. Willems A, Lomov SV, Vandepitte D, Verpoest I, Double dome forming simulation of
woven textile composites. The nineth international conference on material forming
ESAFORM, April 2006:26–28.
78. Cao J, et al., Characterization of mechanical behavior of woven fabrics: Experimental
methods and benchmark results. Composites Part A: Applied Science and Manufacturing
2008; 39(6):1037-1053.
79. Peng XQ, et al., Experimental and numerical analysis on normalization of picture frame
tests for composite materials. Composites Science and Technology 2004; 64(1):11-21.
80. Clifford MJ, Long AC, De Luca P, Forming of engineered prepregs and reinforced ther-
moplastics. Second Global Symposium on Innovations in Materials Processing and
Manufacturing: Sheet Materials 2001:303-316.
81. Long AC, Rudd CD, Clifford MJ, Draping of Engineered Prepregs and Reinforced Ther-
moplastics (DEPART) Review Report.
References 141

82. Xue P, Peng X, Cao J, A non-orthogonal constitutive model for characterizing woven
composites. Composites Part A: Applied Science and Manufacturing 2003; 34(2):183–
193.
83. Peng XQ, Cao J, A continuum mechanics-based non-orthogonal constitutive model for
woven composite fabrics. Composites: Part A 2005; 36:859–874.
84. Cao J, Xue P, Peng X, Krishnan N, An approach in modeling the temperature effect in
thermo-stamping of woven composites. Composite Structures 2003; 61(4):413-420.
85. Yu WR, et al., Analysis of flexible bending behavior of woven preform using non-
orthogonal constitutive equation. Composites Part A: Applied Science and Manufacturing
2005; 36(6):839-850.
86. Yu WR, et al., Non-orthogonal constitutive equation for woven fabric reinforced thermo-
plastic composites. Composites Part A: Applied Science and Manufacturing 2002;
33(8):1095-1105.
87. Xue P, Cao J, Chen J, Integrated micro/macro-mechanical model of woven fabric com-
posites under large deformation. Composite Structures 2005; 70(1):69-80.
88. Krebs J, Friedrich K, Bhattacharyya D, A direct comparison of matched-die versus dia-
phragm forming. Composites Part A: Applied Science and Manufacturing 1998; 29(1-
2):183-188.
89. Molyneux M, Murray P, P. Murray B, Prepreg, tape and fabric technology for advanced
composites. Composites 1983; 14(2):87-91.
90. Liu K, Piggott MR, Shear strength of polymers and fibre composites: 1. Thermoplastic
and thermoset polymers. Composites 1995; 26(12):829-840.
91. Yu X, Zhang L, Mai Y-W, Modelling and finite element treatment of intra-ply shearing
of woven fabric. Journal of Materials Processing Technology 2003; 138(1–3):47–52.
92. Sharma SB, Sutcliffe MPF, Chang SH, Characterisation of material properties for draping
of dry woven composite material. Composites Part A: Applied Science and Manufactur-
ing 2003; 34(12):1167-1175.
93. Gasser A, Boisse P, Hanklar S, Mechanical behaviour of dry fabric reinforcements. 3D
simulations versus biaxial tests. Computational Materials Science 2000; 17(1):7-20.
94. Boisse P, Zouari B, Gasser A, A mesoscopic approach for the simulation of woven fibre
composite forming. Composites Science and Technology 2005; 65(3–4):429–436.
95. Wang J, Page JR, Paton R, Experimental investigation of the draping properties of rein-
forcement fabrics. Composites Science and Technology 1998; 58(2):229-237.
References 142

96. Mohammed U, Lekakou C, Bader MG, Experimental studies and analysis of the draping
of woven fabrics. Composites Part A: Applied Science and Manufacturing 2000;
31(12):1409-1420.
97. Duffy BR, Flow of a liquid with an anisotropic viscosity tensor. Journal of Non-
Newtonian Fluid Mechanics 1978; 4(3):177-193.
98. Dykes RJ, Martin TA, Bhattacharyya D, Determination of longitudinal and transverse
shear behaviour of continuous fibre-reinforced composites from vee-bending. Composites
Part A: Applied Science and Manufacturing 1998; 29(1-2):39-49.
99. Lussier D, Chen J. Shear frame standardization for stamping of themoplastic woven fab-
ric composites. Presented in Revolutionary Materials: Technology and Economics: 32nd
International SAMPE Technical Conference. 2000. Boston, Massachusetts.
100. Potluri P, Perez Ciurezu DA, Ramgulam RB, Measurement of meso-scale shear deforma-
tions for modelling textile composites. Composites Part A: Applied Science and Manu-
facturing 2006; 37(2):303-314.
101. Peng X, Cao J, A dual homogenization and finite element approach for material charac-
terization of textile composites. Composites Part B: Engineering 2002; 33(1):45-56.
102. Takano N, Uetsuji Y, Kashiwagi Y, Zako M, Hierarchical modelling of textile composite
materials and structures by the homogenization method. Modelling and Simulation in Ma-
terials Science and Engineering 1999; 7(2):207-231.
103. Yu X, et al., Intra-ply shear locking in finite element analyses of woven fabric forming
processes. Composites Part A: Applied Science and Manufacturing 2006; 37(5):790–803.
104. Ye L, Daghyani HR, Characteristics of woven fibre fabric reinforced composites in form-
ing process. Composites Part A: Applied Science and Manufacturing 1997; 28(9-10):869-
874.
105. Mulhern JF, Rogers TG, Spencer AJM, A Continuum Model for Fibre-Reinforced Plastic
Materials. Proceedings of the Royal Society of London. Series A, Mathematical and
Physical Sciences 1967; 301(1467):473-492.
106. Potter K, Bias extension measurements on cross-plied unidirectional prepreg. Composites
Part A: Applied Science and Manufacturing 2002; 33(1):63-73.
107. Long AC, Rudd CD, Blagdon M, Smith P, Characterizing the processing and perform-
ance of aligned reinforcements during preform manufacture. Composites Part A: Applied
Science and Manufacturing 1996; 27(4):247-253.
108. Lomov SV, Barburski M, Stoilova T, Verpoest I, Experimental textile mechanics charac-
terisation of deformability of reinforcements for textile composites. 2004.
References 143

109. Stoilova T, Lomov SV, Characterization of Glasspolypropylene Fabrics. ROUND-ROBIN


Formability Study, MTM, K. U. Leuven 2004.
110. Prodromou AG, Chen J, On the relationship between shear angle and wrinkling of textile
composite preforms. Composites Part A: Applied Science and Manufacturing 1997;
28(5):491-503.
111. Conchúr M. O'Brádaigh RBPPJM, Issues in diaphragm forming of continuous fiber rein-
forced thermoplastic composites. Polymer Composites 1991; 12(4):246-256.
112. Cao J, et al., A cooperative benchmark effort on testing of woven composites. Proceed-
ings of the 7th Esaform Conference on Material Forming,” Trondheim, Norway
2004:305–308.
113. Harrison P, Wiggers J, Long AC, Clifford MJ, INTERNAL TEST STANDARD: Con-
tinuous fibre reinforced composites- Determination of the in-plane shear stress response
to shear strain and shear strain rate, using the picture-frame test. Polymer Composites Re-
search Group, University of Nottingham 2002; 1.
114. Stoilova T, Lomov SV, Glass polypropylene woven fabrics- shear diagrams and full field
strain measurements. ROUND-ROBIN Formability Study, MTM, K. U. Leuven 2005.
115. Lomov SV, et al., Full-field strain measurements in textile deformability studies. Com-
posites Part A: Applied Science and Manufacturing; In Press.
116. Ahmad S, Irons BM, Zienkiewicz OC, Analysis of thick and thin shell structures by
curved finite elements. International Journal for Numerical Methods in Engineering
1970; 2(3).
117. ‹
‰‹Œ
“‹Œ”
\ 

 


 ^
> 

element for thin multilayered elastic shells. Computational Mechanics 1995; 16(5):341-
359.
118. Cardoso RPR, Yoon JW, Valente RAF, A new approach to reduce membrane and trans-
verse shear locking for one-point quadrature shell elements: linear formulation. Interna-
tional Journal for Numerical Methods in Engineering 2006; 66(2):214-249.
119. Chang TYP, Saleeb AF, Li G, Large strain analysis of rubber-like materials based on a
perturbed Lagrangian variational principle. Computational Mechanics 1991; 8(4):221-
233.
120. Chapelle D, Bathe KJ, Fundamental considerations for the finite element analysis of shell
structures. Computers & Structures 1998; 66(1):19-36.
References 144

121. J. C. Simo FA, Geometrically non-linear enhanced strain mixed methods and the method
of incompatible modes. International Journal for Numerical Methods in Engineering
1992; 33(7):1413-1449.
122. Meek JL, Wang Y, Nonlinear static and dynamic analysis of shell structures with finite
rotation. Computer Methods in Applied Mechanics and Engineering 1998; 162(1-4):301-
315.
123. Reddy JN, On refined computational models of composite laminates. International Jour-
nal for Numerical Methods in Engineering 1989; 27(2):361-382.
124. Sandhu JS, Stevens KA, Davies GAO, A 3-D, co-rotational, curved and twisted beam
element. Computers & Structures 1990; 35(1):69–79.
125. Campanelli M, Berzeri M, Shabana AA, Performance of the Incremental and Non-
Incremental Finite Element Formulations in Flexible Multibody Problems. Journal of
Mechanical Design 2000; 122:498.
126. Braun M, Bischoff M, Ramm E, Nonlinear shell formulations for complete three-
dimensional constitutive laws including composites and laminates. Computational Me-
chanics 1994; 15(1):1-18.
127. Jr DHR, Reddy JN, Variable kinematic modelling of laminated composite plates. Interna-
tional Journal for Numerical Methods in Engineering 1996; 39(13):2283-2317.
128. Kim K-D, Lee C-S, Han S-C, A 4-node co-rotational ANS shell element for laminated
composite structures. Composite Structures 2007; 80(2):234-252.
129. Argyris J, An excursion into large rotations. Computer Methods in Applied Mechanics
and Engineering 1982; 32(1-3):85-155.
130. Moita GF, Crisfield MA, A finite element formulation for 3-D continua using the co-
rotational technique. International Journal for Numerical Methods in Engineering 1996;
39(22):3775-3792.
131. Mojia H, Ming L, Mingfu F, On several new methods to polar decomposition computa-
tion. Acta Mechanica Solida Sinica 1998; 11(4):329-340.
132. Zhongxue Li LV-Q, An efficient co-rotational formulation for curved triangular shell
element. International Journal for Numerical Methods in Engineering 2007;
9999(9999):n/a.
133. Rankin CC, Brogan FA, An element independent corotational procedure for the treatment
of large rotations. Journal of pressure vessel technology 1986; 108(2):165-174.
134. Oden JT, Finite Elements of Nonlinear Continua. Advanced Engineering Series,
McGraw-Hill 1972.
References 145

135. Zienkiewicz OC, Taylor RL, Nithiarasu P, Zhu JZ, The Finite Element Method. 6th ed.
Elsevier/Butterworth–Heinemann, 2005.
136. Truesdell C, Noll W, The Non-linear Field Theories of Mechanics. 3rd ed, ed. S.S. Ant-
man. Springer Verlag: Berlin, Heidelberg, New York, 2004.
137. Mojia H, Bufler H, Updated Lagrangian polar decomposition method in non-linear elas-
ticity and its shell element implementation. Communications in Numerical Methods in
Engineering 1999; 15(3):167–181.
138. Reddy JN, A generalization of two-dimensional theories of laminated composite plates.
Communications in Applied Numerical Methods 1987; 3(3):173–180.
139. Mares T, Anisotropic elasticity in curvilinear coordinates. Bulletin of Applied Mechanics
2007; 3(9).
140. El-Abbasi N, Meguid S, A Continuum Based Thick Shell Element for Large Deformation
Analysis of Layered Composites. International Journal of Mechanics and Materials in
Design 2005; 2(1):99-115.
141. Pipkin AC, Sanchez VM, Existence of Solutions of Plane Traction Problems for Ideal
Composites. SIAM Journal on Applied Mathematics 1974; 26(1):213-220.
142. Conchúr M. Ó Brádaigh RBP, A finite element formulation for highly anisotropic incom-
pressible elastic solids. International Journal for Numerical Methods in Engineering
1992; 33(8):1573-1596.
143. Morland LW, Elastic anisotropy of regularly jointed media. Rock Mechanics and Rock
Engineering 1976; 8(1):35-48.
144. Zheng QS, Betten J, The formulation of constitutive equations for fibre-reinforced com-
posites in plane problems: Part II. Archive of Applied Mechanics (Ingenieur Archiv) 1995;
65(3):161–177.
145. Zheng QS, Betten J, Spencer AJM, The formulation of constitutive equations for fibre-
reinforced composites in plane problems: Part I. Archive of Applied Mechanics (Ingenieur
Archiv) 1992; 62(8):530–543.
146. Zheng QS, Two-Dimensional Tensor Function Representation for All Kinds of Material
Symmetry. Proceedings: Mathematical and Physical Sciences 1993; 443(1917):127-138.
147. Qiu GY, Pence TJ, Remarks on the Behavior of Simple Directionally Reinforced Incom-
pressible Nonlinearly Elastic Solids. Journal of Elasticity 1997; 49(1):1-30.
148. Nedjar B, An anisotropic viscoelastic fibre-matrix model at finite strains: Continuum
formulation and computational aspects. Computer Methods in Applied Mechanics and
Engineering 2007; 196(9–12):1745–1756.
References 146

149. Klinkel S, Sansour C, Wagner W, An anisotropic fibre-matrix material model at finite


elastic-plastic strains. Computational Mechanics 2005; 35(6):409–417.
150. Klinkel S, Gruttmann F, Wagner W, A continuum based three-dimensional shell element
for laminated structures. Computers & Structures 1999; 71(1):43–62.
151. Brannon RM, Using transverse isotropy to model arbitrary deformation-induced anisot-
ropy, 1996; 28 p. ; PL:.
152. Brannon RM, Large deformation analysis of axisymmetric inhomogeneities including
coupled elastic and plastic anisotropy, 1996; 11 p. ; PL:.
153. Brannon RM, Caveats concerning conjugate stress and strain measures for frame indiffer-
ent anisotropic elasticity. Acta Mechanica 1998; 129(1):107-116.
154. Brannon RM, On the distinction between large deformation and large distortion for ani-
sotropic materials, 2000; 3 pages.
155. Kyriacou SK, Humphrey JD, Schwab C, Finite element analysis of nonlinear orthotropic
hyperelastic membranes. Computational Mechanics 1996; 18(4):269-278.
156. Chung DH, Kwon TH, Fiber orientation in the processing of polymer composites. Korea-
Australia Rheology Journal 2002; 14(4):175-188.
157. Manach PY, Rio G, Analysis of orthotropic behavior in convected coordinate frames.
Computational Mechanics 1999; 23(5):510-518.
158. Zheng QS, On the generalization of constitutive laws from their arotational forms. Acta
Mechanica 1992; 91(1):97-105.
159. Dienes JK, On the analysis of rotation and stress rate in deforming bodies. Acta
Mechanica 1979; 32(4):217-232.
160. Dienes J, A discussion of material rotation and stress rate. Acta Mechanica 1987; 65(1):1-
11.
161. Rajagopal KR, Srinivasa AR, Modeling anisotropic fluids within the framework of bodies
with multiple natural configurations. Journal of Non-Newtonian Fluid Mechanics 2001;
99(2-3):109-124.
162. Itskov M, A generalized orthotropic hyperelastic material model with application to in-
compressible shells. International Journal for Numerical Methods in Engineering 2001;
50(8):1777-1799.
163. Christoffersen J, A Class of Initially Isotropic Elastic Materials. Proceedings of the Royal
Society of London. Series A, Mathematical and Physical Sciences 1988; 417(1852):133-
154.
References 147

164. Haupt P, Tsakmakis C, On the application of dual variables in continuum mechanics.


Continuum Mechanics and Thermodynamics 1989; 1(3):165-196.
165. Xiao H, Bruhns OT, Meyers A, A consistent finite elastoplasticity theory combining addi-
tive and multiplicative decomposition of the stretching and the deformation gradient. In-
ternational Journal of Plasticity 2000; 16(2):143-177.
166. Poisson SD, Mémoire sur l'équilibre et le mouvement des corps cristallisés. Mémoires de
l'Académie des Sciences, Paris 1842; 18.
167. Duhem P, Sur les lois générales de l'induction électrodynamique. Annales de la faculté
des sciences de Toulouse 1're Série 1893; 7(1):B1–B28.
168. Cosserat E, les congruences de droites et sur la théorie des surfaces. Annales de la faculté
des sciences de Toulouse 1're Série 1893; 7(1):N 1-B 62.
169. Cosserat E, Cosserat F, Théorie des corps déformables. Bull. Amer. Math. Soc. 19 (1913),
242-246. 1913.
170. Stokes VK, Theories of Fluids With Microstructure: An Introduction. Springer Verlag,
1984.
171. Green AE, Naghdi PM, On Electromagnetic Effects in the Theory of Shells and Plates.
Philosophical Transactions of the Royal Society of London. Series A, Mathematical and
Physical Sciences (1934-1990) 1983; 309(1510):559-610.
172. Rovati M, Veber D, Optimal topologies for micropolar solids. Structural and Multidisci-
plinary Optimization 2007; 33(1):47-59.
173. Eringen AC, SIMPLE MICROFLUIDS. 1963.
174. Eringen AC, Theory of micropolar elasticity. Fracture. An advanced treatise”(Ed. H.
Liebowitz), Mathematical Fundamentals 1968; 2:622–729.
175. Mindlin RD, Influence of rotatory inertia and shear on flexural motions of isotropic, elas-
tic plates. Journal of Applied Mechanics 1951; 18(1):31-38.
176. Mindlin RD, Micro-structure in linear elasticity. Archive for Rational Mechanics and
Analysis 1964; 16(1):51-78.
177. Mindlin RD, Stress functions for a Cosserat continuum. International Journal of Solids
and Structures 1965; 1(3):265-271.
178. Mindlin RD, Tiersten HF, Effects of couple-stresses in linear elasticity. Archive for Ra-
tional Mechanics and Analysis 1962; 11(1):415-448.
179. Cowin SC, An incorrect inequality in micropolar elasticity theory. Zeitschrift für Ange-
wandte Mathematik und Physik (ZAMP) 1970; 21(3):494-497.
References 148

180. Green AE, Rivlin RS, Multipolar continuum mechanics. Archive for Rational Mechanics
and Analysis 1964; 17(2):113-147.
181. Smith AC, Deformations of micropolar elastic solids. International Journal of Engineer-
ing Science 1967; 5(8):637-651.
182. Toupin RA, Theories of elasticity with couple-stress. Archive for Rational Mechanics and
Analysis 1964; 17(2):85-112.
183. Ramezani S, Naghdabadi R, Energy pairs in the micropolar continuum. International
Journal of Solids and Structures 2007; 44(14-15):4810-4818.
184. Achenbach JD, A theory of elasticity with microstructure for directionally reinforced
composites. CISM Courses and Lectures No. 167. 1975, New York: Springer.
185. Hutapea P, Qiao P, Micropolar in-Plane Shear and Rotation Moduli of Unidirectional Fi-
ber Composites with Fiber-Matrix Interfacial Debonding. Journal of Composite Materials
2002; 36(11):1381-1399.
186. Xun F, Hu G, Huang Z, Effective in plane moduli of composites with a micropolar matrix
and coated fibers. International Journal of Solids and Structures 2004; 41(1):247-265.
187. Bigoni D, Drugan WJ, Analytical derivation of Cosserat moduli via homogenization of
heterogeneous elastic materials. Journal of Applied Mechanics-Transactions of the Asme
2007; 74(4):741-753.
188. Lakes R, Experimental methods for study of Cosserat elastic solids and other generalized
elastic continua. Continuum Models for Materials with Microstructure 1995:1-25.
189. Nakamura S, Benedict R, Lakes R, Finite element method for orthotropic micropolar
elasticity. International Journal of Engineering Science 1984; 22(3):319-330.
190. Yeh J-T, Chen W-H, Shell elements with drilling degree of freedoms based on micropolar
elasticity theory. International Journal for Numerical Methods in Engineering 1993;
36(7):1145–1159.
191. Zhu Y, Zacharia T, A new one-point quadrature, quadrilateral shell element with drilling
degrees of freedom. Computer Methods in Applied Mechanics and Engineering 1996;
136(1-2):165-203.
192. D. P. Adhikary, Mühlhaus HB, Dyskin AV, Modelling the large deformations in stratified
media - the Cosserat continuum approach. Mechanics of Cohesive-frictional Materials
1999; 4(3):195-213.
193. Figueiredo RPd, Vargas EdA, Moraes A, Analysis of bookshelf mechanisms using the
mechanics of Cosserat generalized continua. Journal of Structural Geology 2004;
26(10):1931-1943.
References 149

194. Fernlund G, et al., Finite element based prediction of process-induced deformation of


autoclaved composite structures using 2D process analysis and 3D structural analysis.
Composite Structures 2003; 62(2):223-234.
195. Crookston JJ, Long AC, Jones IA, Modelling effects of reinforcement deformation during
manufacture on elastic properties of textile composites. Plast Rubber Compos 2002; 31:1-
7.
196. Cherouat A, Billoet JL, Mechanical and numerical modelling of composite manufacturing
processes deep-drawing and laying-up of thin pre-impregnated woven fabrics. Journal of
Materials Processing Technology 2001; 118(1-3):460-471.
197. Yu X, Ye L, Mai Y, Spurious Wrinkles in Forming Simulations of Woven Fabric. Inter-
national Journal of Forming Processes Special Issue 2005:141-155.
198. ten Thije RHW, Akkerman R, Intra-ply shear locking. AIP Conference Proceedings
2007; 907(1):1017-1022.
199. Furukawa T, Yagawa G, Implicit constitutive modelling for viscoplasticity using neural
networks. Int. J. Numer. Meth. Engng 1998; 43:195-219.
200. Al-Haik MS, Garmestani H, Savran A, Explicit and implicit viscoplastic models for
polymeric composite. International Journal of Plasticity 2004; 20(10):1875-1907.
201. Michopoulos J, Lambrakos S, On the Fundamental Tautology of Validating Data-Driven
Models and Simulations. See Sunderam, van Albada, Sloot, and Dongarra (2005)
2005:738–745.
202. Boisse P, Borr M, Buet K, Cherouat A, Finite element simulations of textile composite
forming including the biaxial fabric behaviour. Composites Part B: Engineering 1997;
28(4):453–464.
203. McBride TM, Chen J, Unit-cell geometry in plain-weave fabrics during shear deforma-
tions. Composites Science and Technology 1997; 57(3):345-351.
204. Hsiao S-W, Kikuchi N, Numerical analysis and optimal design of composite thermoform-
ing process. Computer Methods in Applied Mechanics and Engineering 1999; 177(1-2):1-
34.
205. Boisse P, Meso-macro approach for composites forming simulation. Journal of Materials
Science 2006; 41(20):6591–6598.
206. Boisse P, Zouari B, Daniel J-L, Importance of in-plane shear rigidity in finite element
analyses of woven fabric composite preforming. Composites Part A: Applied Science and
Manufacturing 2006; 37(12):2201-2212.
References 150

207. Boisse P, Gasser A, Hagege B, Billoet J-L, Analysis of the mechanical behavior of woven
fibrous material using virtual tests at the unit cell level. Journal of Materials Science
2005; 40(22):5955-5962.
208. Cook RD, Malkus DS, Plesha ME, Witt RJ, Concepts and Applications of Finite Element
Analysis. John Wiley & Sons, 2007.
209. Badel P, Vidal-Sallé E, Boisse P, Computational determination of in-plane shear me-
chanical behaviour of textile composite reinforcements. Computational Materials Science
2007; 40(4):439-448.
210. Hivet G, et al., Mechanical Behavior of Woven Composite Reinforcements While Form-
ing. Journal of Thermoplastic Composite Materials 2002; 15(6):545-555.
211. Creech G, Pickett A, Meso-modelling of Non-Crimp Fabric composites for coupled drape
and failure analysis. Journal of Materials Science 2006; 41(20):6725-6736.
212. Pickett AK, Creech G, Mesoscopic Finite Element modelling of non-crimp fabrics for
drape and failure analyses. 2006.
213. Robitaille F, et al., Geometric modelling of industrial preforms: warp-knitted textiles.
Proceedings of the I MECH E Part L Journal of Materials:Design and Applicat 2000;
214:71-90.
214. Robitaille F, et al., Geometric modelling of industrial preforms: woven and braided tex-
tiles. Proceedings of the Institution of Mechanical Engineers, Part L: Journal of Materi-
als: Design and Applications 1999; 213(2):69-83.
215. Robitaille F, Long AC, Rudd CD, Geometric modelling of textiles for prediction of com-
posite processing and performance characteristics. Plastics, Rubber and Composites
2002; 31(2):66-75.
216. Robitaille F, Long AC, Jones IA, Rudd CD, Automatically generated geometric descrip-
tions of textile and composite unit cells. Composites Part A: Applied Science and Manu-
facturing 2003; 34(4):303-312.
217. Desplentere F, et al., Geometrical characterization of 3-D warp-interlaced fabrics. Inter-
national SAMPE symposium and exhibiton 2003:1335-1347.
218. Sidhu RMJS, Averill RC, Riaz M, Pourboghrat F, Finite element analysis of textile com-
posite preform stamping. Composite Structures 2001; 52(3-4):483-497.
219. King MJ, Jearanaisilawong P, Socrate S, A continuum constitutive model for the me-
chanical behavior of woven fabrics. International Journal of Solids and Structures 2005;
42(13):3867-3896.
References 151

220. Quintanilla R, Straughan B, Growth and Uniqueness in Thermoelasticity. Proceedings:


Mathematical, Physical and Engineering Sciences 2000; 456(1998):1419-1429.
221. Farrashkhalvat M, Miles JP, Tensor Methods for Engineers and Scientists. Ellis Horwood
Ltd.: West Sussex, 1990; 213ppp.
222. Green AE, Naghdi PM, A new thermoviscous theory for fluids. Journal of Non-
Newtonian Fluid Mechanics 1995; 56(3):289-306.
223. Green AE, Naghdi PM, A Re-Examination of the Basic Postulates of Thermomechanics.
Proceedings: Mathematical and Physical Sciences 1991; 432(1885):171-194.
224. Brannon RM, Mohr's Circle And More Circles. University of New Mexico, 2003.
225. Matthews FL, Davies GAO, Hitchings D, Soutis C, Finite Element Modelling of Compos-
ite Materials and Structures. Woodhead Publishing Ltd: Cambridge, 2000.
226. Murdoch AI, On Material Frame-Indifference. Proceedings of the Royal Society of Lon-
don. Series A, Mathematical and Physical Sciences 1982; 380(1779):417-426.
227. Müller I, On the frame dependence of stress and heat flux. Archive for Rational Mechan-
ics and Analysis 1972; 45(4):241-250.
228. Murdoch AI, The motivation of continuum concepts and relations from discrete consid-
erations. Q J Mechanics Appl Math 1983; 36(2):163-187.
229. Daniel IM, Ishai O, Engineering Mechanics of Composite Materials. Oxford University
Press: New York, 1994.
230. Chandrupatla TR, Belegundu AD, Introduction to Finite Elements in Engineering. Pren-
tice Hall: New Jersey, 2002.
231. Yang HTY, Saigal S, Masud A, Kapania RK, A survey of recent shell finite elements.
International Journal for Numerical Methods in Engineering 2000; 47(1–3):101–127.
232. Felippa CA, A Systematic Approach to the Element-Independent Corotational Dynamics
of Finite Elements. IASS-IACM 2000, fourth international colloquium on computation of
shell and spatial structures, Chania-Crete, Greece 2000.
233. Day RA, Potts DM, Curved Mindlin beam and axi-symmetric shell elements - A new ap-
proach. International Journal for Numerical Methods in Engineering 1990; 30(7):1263-
1274.
234. Munn RW, Finite strain and finite rotation of solids. Journal of Physics C: Solid State
Physics 1978(2):L61.
235. Bufler H, On drilling degrees of freedom in nonlinear elasticity and a hyperelastic mate-
rial description in terms of the stretch tensor. Part 1: Theory. Acta Mechanica 1995;
113(1):21-35.
References 152

236. Sansour C, Bufler H, An exact finite rotation shell theory, its mixed variational formula-
tion and its finite element implementation. International Journal for Numerical Methods
in Engineering 1992; 34(1):73-115.
237. Sansour C, Bufler H, Müllerschön H, On drilling degrees of freedom in nonlinear elastic-
ity and a hyperelastic material description in terms of the stretch tensor. Part 2: Applica-
tion to membranes. Acta Mechanica 1996; 115(1):103-117.
238. Simo JC, Taylor RL, Quasi-incompressible finite elasticity in principal stretches. contin-
uum basis and numerical algorithms. Computer Methods in Applied Mechanics and Engi-
neering 1991; 85(3):273–310.
239. Wriggers P, Gruttmann F, Thin shells with finite rotations formulated in biot stresses:
Theory and finite element formulation. International Journal for Numerical Methods in
Engineering 1993; 36(12):2049-2071.
240. Padovan J, Parris H, Ma J, Large Deformation Micropolar Theory for Cord–Rubber
Composites. Rubber Chemistry and Technology Journal 1995; 68(2):77-96.
241. Macosko CW, Rheology: Principles, Measurements and Applications. Poughkeepsie, NY
Wiley-VCH, 1994.
242. Naghdi PM, Vongsarnpigoon L, A theory of shells with small strain accompanied by
moderate rotation. Archive for Rational Mechanics and Analysis 1983; 83(3):245-283.
243. Hughes TJR, Tezduyar TE, Finite elements based upon Mindlin plate theory with particu-
lar reference to the four-node bilinear isoparametric element. New concepts in finite ele-
ment analysis 1981:81-106.
244. Taylor RL, Beresford PJ, Wilson EL, A non-conforming element for stress analysis. In-
ternational Journal for Numerical Methods in Engineering 1976; 10(6):1211-1219.
245. Dvorkin EN, Bathe KJ, A continuum mechanics based four-node shell element for gen-
eral nonlinear analysis. Engineering Computations 1984; 1(1):77-88.
246. Hughes TJR, Numerical implementation of constitutive models: rate-independent devia-
toric plasticity. Theoretical Foundation for Large Scale Computations for Nonlinear Ma-
terial Behavior, Martinus Nijhoff Publishers 1983.
247. Xiao H, Bruhns OT, Meyers A, On objective corotational rates and their defining spin
tensors. International Journal of Solids and Structures 1998; 35(30):4001-4014.
248. Coleman BD, Noll W, Foundations of Linear Viscoelasticity. Reviews of Modern Physics
1961; 33(2):239.
APPENDICES

Appendix A IJNME

Strongly orthotropic continuum mechanics and finite element treatment

Appendix B ACAM

A theory of strongly orthotropic continuum mechanics

Appendix C APCOM

Finite element implementation and sub-laminar decomposition for strongly orthotropic


continuum mechanics

Appendix D ICTAM

Intrinsic field tensors for strongly orthotropic continua

Appendix E USNCCM

A continuum mechanics solution for in-plane shear locking in plate and shell elements

Appendix F Composites Australia

Intrinsic field tensors for strongly orthotropic continua

153

S-ar putea să vă placă și