You are on page 1of 232

Multifunctional Elastomers with

Tailored Anisotropic Response

David C. Stanier

A dissertation submitted to the University of Bristol in accordance with the


requirements for award of the degree of Doctor of Philosophy in the Faculty
of Engineering

May 2016

Word Count: 71140

Abstract

Traditional engineering design focusses on the load-bearing capabilities of materials, and


develops functionality by adding the structure. However, hybridising with functional materials allows non-load bearing capabilities to be integrated, that simplify the design. Material
anisotropy can effect this multi-functionality in complimentary ways, however its effects
need to be understood.
The aims of this study were to produce tailored anisotropy in elastomers, and to investigate the resulting mechanical and magnetic performance. Hence, a combination of experimental testing, and numerical & constitutive modelling were used to analyse the mechanical
behaviour.
The placement of fibres with a homogeneous magnetic field is demonstrated and described by a simple model. The resulting transversely isotropic elastomers, reinforced by
short nickel-coated carbon fibres, are compared to a number of numerical and constitutive
mechanical models. Current methods for describing the behaviour of aligned reinforcements assume an inherent anisotropy, analogous to continuous fibre reinforcement, rather
than discontinuous reinforcements. However, it is demonstrated that a simple constitutive
model can adequately describe the behaviour up to moderate strains (<30%). In addition,
a simplified numerical model is shown to represent the behaviour, and could be adapted to
investigate complex effects, such as failure and interfacial properties.
In addition, the specimens actuate in a magnetic field, due to the nickel-functionalised
fibres. The actuation is dependent on the reinforcement angle and is described by a simple
model; furthermore, the combination of the magnetic and mechanical models allows the
complimentary behaviour of these properties to be described.
The multi-functional material could be envisaged in a number of high performance applications, such as the active surface of micro-swimmers & -controllers. However, there are
a number of challenges in experimental testing of anisotropic materials that require further
investigation. Never-the-less, the orientation of reinforcements could be used to produce
bespoke fibre alignments or for through-thickness composite repair.

Publications

In the following, the reader is directed to a list of publications produced during the course of
this Research Project. In particular, the author would like to acknowledge that Section 2.2
reuses work published in [1 ], [6 ] and [7 ], Section 2.3 contains results from [2 ] and [3 ],
whilst Section 3.3 and Section 3.4 use work from [4 ] and [3 ], respectively. Section 4.3
uses work from [2 ], whilst [8 ] contains work from Section 2.3, Section 4.2 and Section
4.3.

Journal Publications
[1 ]
Stanier, D.C., Patil, A., Sriwong, C., Rahatekar, S., and Ciambella, J. (2014). The
reinforcement effect of exfoliated graphene oxide nanoplatelets on the mechanical and viscoelastic properties of natural rubber. Composites Science and Technology, 95:59-66.

Conference Proceedings
[2 ]
Stanier, D. C., Ciambella, J. and Rahatekar, S. (2015). Magnetic Response of
aligned nickel coated carbon fibres in a PDMS matrix. In Proceedings of the 9th European
Conference on Constitutive Models for Rubbers (ECCMR IX), Prague.
[3 ]
Ciambella, J. and Stanier, D.C. (2014). Orientation Effects in Short Fibre-Reinforced
Elastomers. In Proceedings of ASME 2014 International Mechanical Engineering Congress
and Exposition, Montreal. American Society of Mechanical Engineers (ASME).
[4 ]
Stanier, D.C. and Ciambella, J. (2014). Multi scale Modelling Nano-Platelet Reinforced Composites at Large Strain. In Proceedings of ECCM 16th European Conference on
Composite Materials, Seville. European Society for Composite Materials (ESCM).
[5 ]
Stanier, D.C., Patil, A.J., Rahatekar, S.S., and Ciambella, J. (2013). Microstructure
Evolution in Graphene Oxide Reinforced Elastomers. In Proceedings of the 7th International Nanoscience Student Conference (INASCON), London.

[6 ]
Stanier, D.C., Ciambella, J., Rahatekar, S., Patil, A.J. and Mann, S. (2013). The
Reinforcement Effects of Graphene Oxide on the Mechanical and Viscoelastic Properties of
Natural Rubber. In Proceedings of the 8th European Conference on Constitutive Models for
Rubbers (ECCMR VIII), San Sebastian.

Internal Publications
[7 ]
Stanier, D.C., Ciambella, J., Rahatekar, S., Patil, A.J. and Mann, S. (2013) The Reinforcement Effects of Graphene Oxide Nanoplatelets on the Mechanical and Viscoelastic
Properties of Natural Rubber. University of Bristol (internal unpublished document).

Pending Publications
[8 ]
Stanier, D.C., Ciambella, J. and Rahatekar, S. Bending and Twisting Actuation of
Nickel Coated Carbon Fibres and Elastomer Composites using Low Magnetic Field. Composites Part A, Under Review 2016.
[9 ]
Ciambella, J., Stanier, D.C., and Rahatekar, S. Magnetic Alignment of Short Carbon
Fibres in Curing Composites by Uniform and Non-Uniform Fields. To be submitted in 2016.

Acknowledgements

This thesis has been realised thanks to the contributions of a number of individuals.
I would like to acknowledge the assistance of my supervisors. In particular, I have been
extremely lucky to have such a dedicated and passionate supervisor as Jacopo, who was able
to respond to my enquires and concerns so promptly, even when a thousand miles away. He
has given me an enormous amount of freedom, and given me the opportunity to experience
so much.
I wish to express my gratitude for the unerring belief and support of my parents and
close family, and the encouragement of my friends. I am also indebted to Iryna, who has
been closest during the most difficult periods; offering continued patience and optimism that
I hope to repay.
Finally, I must thank the Engineering and Physical Sciences Research Council, for providing the funding which has made all of this possible.

Declaration

I declare that the work in this dissertation was carried out in accordance with the requirements of the Universitys Regulations and Code of Practice for Research Degree Programmes
and that it has not been submitted for any other academic award. Except where indicated by
specific reference in the text, the work is the candidates own work. Work done in collaboration with, or with the assistance of, others, is indicated as such. Any views expressed in
the dissertation are those of the author.

SIGNED:................................................................................ DATE:....................................

Table of contents
Table of contents

xiii

List of figures

xvii

List of tables

xxix

Acronyms / Abbreviations

xxxii

Introduction

1.1

Motivations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.2

Technical Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.3

Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.4

Structure of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Experimental Mechanical Characterisation

2.1

2.2

2.3

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1

The Benefits of Homogeneously Distributed Reinforcements . . . .

10

2.1.2

The Benefits of Oriented and Aligned Reinforcements . . . . . . .

15

The anisotropic effects of nano-inclusions . . . . . . . . . . . . . . . . . .

20

2.2.1

Natural Rubber Reinforced by Graphene Oxide . . . . . . . . . . .

20

2.2.2

Materials characterization and experimental techniques . . . . . . .

21

2.2.3

Results and discussion . . . . . . . . . . . . . . . . . . . . . . . .

24

2.2.4

Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . .

34

The anisotropic effects of micro-inclusions . . . . . . . . . . . . . . . . . .

36

2.3.1

PDMS reinforced by Carbon Fibres . . . . . . . . . . . . . . . . .

36

2.3.2

Material Preparation and Experimental Techniques . . . . . . . . .

37

2.3.3

Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . .

40

2.3.4

Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . .

49

xiv
3

Table of contents
Numerical Modelling
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.1 Literature Review . . . . . . . . . . . . . . . . . . . .
3.2 Constitutive Modelling . . . . . . . . . . . . . . . . . . . . .
3.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
3.2.2 The Transversally Isotropic Neo-Hookean Model . . .
3.3 Numerical Homogenisation . . . . . . . . . . . . . . . . . . .
3.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
3.3.2 Homogenization Procedure . . . . . . . . . . . . . . .
3.3.3 Description of the RVE . . . . . . . . . . . . . . . . .
3.4 Results of the FE Analysis for Fibre Reinforced Elastomers . .
3.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
3.4.2 Numerical Methods . . . . . . . . . . . . . . . . . . .
3.4.3 Results of the SFRVE FE Analysis . . . . . . . . . . .
3.4.4 Results of the MFRVE FE Analysis . . . . . . . . . .
3.4.5 Discussion . . . . . . . . . . . . . . . . . . . . . . .
3.5 Results of the FE Analysis for Platelet Reinforced Elastomers
3.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
3.5.2 Numerical Methods . . . . . . . . . . . . . . . . . . .
3.5.3 Results of the FE analysis . . . . . . . . . . . . . . .
3.5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . .
3.6 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . .
Magnetic Response
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.1 Manufacturing Tailored Reinforced Networks . . . . .
4.1.2 Stimuli Responsive Elastomers . . . . . . . . . . . . .
4.2 Magnetically tailoring the orientation of reinforcements . . . .
4.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
4.2.2 Materials characterization and experimental techniques
4.2.3 Magnetic Experimental Setup and Discussion . . . . .
4.3 Magnetic Actuation . . . . . . . . . . . . . . . . . . . . . . .
4.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . .
4.3.2 Experimental Techniques . . . . . . . . . . . . . . . .
4.3.3 Magnetic Response in a Static Homogeneous Field . .
4.3.4 Results and Discussion . . . . . . . . . . . . . . . . .
4.4 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

51
51
53
66
66
67
78
78
80
81
84
84
85
86
91
94
97
97
99
100
104
110

.
.
.
.
.
.
.
.
.
.
.
.
.

113
113
114
120
126
126
127
130
142
142
142
144
150
153

Table of contents
5

xv

Conclusions
157
5.1 Original Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
5.2 Further Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

References

163

Appendix A Supplementary Material - Chapter 2


183
A.1 The Stress-Strain Curve for Reinforced PDMS . . . . . . . . . . . . . . . . 183
Appendix B Supplementary Material - Chapter 3
B.1 Comparison of the RVE to the Constitutive Model - Plane Stress . . . . . .
B.2 Comparison of the RVE to the Constitutive Model - Plane Strain . . . . . .
B.3 Comparisons of the SPRVE and MPRVE - Plane Strain . . . . . . . . . . .

185
185
191
197

Appendix C Supplementary Material - Chapter 4


199
C.1 The Magneto-Mechanical Experimental Data . . . . . . . . . . . . . . . . 199

List of figures

1.1

2.1

2.2

2.3

2.4

2.5

2.6

Examples of anisotropic architecture in nature. (a) A spruce tree, and (b)


The fibril structure at the spruce tree branch joint (Adapted from [27]). (c)
Venuss Fly Basket Euplectella Aspergillum [235], and (d) A section of the
mineralized skeletal cage of Euplectella Aspergillum (Adapted from [172]).
(e) The marine mussel, anchored by the byssus [93], and (f) A Byssal thread,
with a soft fibrous core and protective cuticles (Adapted from [218]). . . . .

Schematic representation of two types of crystallisation encountered in rubber materials. (a) Static crystallisation. (b) Strain-induced crystallisation. .

Distribution vs Dispersion. (a) A homogeneous distribution of the fillers,


(b) A homogeneous distribution of the fillers with an increased dispersion. .

The effect of particle size on the interfacial area. (a) Large particles of
diameter, d. (b) Small particles of diameter, d/2. Particles of equivalent
area have an interfacial zone over twice the size when the diameter is halved.
The inter-particle spacing is also decreased. . . . . . . . . . . . . . . . . .

10

The tortuous path caused by aligned anisotropic reinforcements in a matrix


medium. The blue arrow shows the path of least resistance for, e.g. water
molecules travelling through a permeable material. . . . . . . . . . . . . .

15

The crack path for aligned anisotropic fillers. The blue arrow indicates the
crack path. (a) parallel to the loading direction, (b) perpendicular to the
loading direction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

17

Schematic structure of (a) graphene and (b) graphene oxide. Adapted from
[150]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

20

xviii
2.7

2.8

2.9

List of figures
(a) AFM images of the GO platelets reveal the delaminated sheets to be
1-1.4 nm in thickness, consistent with completely exfoliated single layer
GO; (b) TEM image of GO platelet reveals ultra-thin flat sheet between 500
nm and 1 lm in size; (c) Fourier-transform spectra of GO, graphite, natural rubber and NR-GO nanocomposites containing 1.00 wt% GO; and (d)
SEM images of fractured 1.00 wt% NR-GO specimens reveal homogeneous
dispersions of filler particles. . . . . . . . . . . . . . . . . . . . . . . . . .

22

(a) Specimens of NR-GO used in experimental testing. Pure NR on the left,


increasing in GO conc. up to 1.00 wt% on the right. (b) The petri dish
arrangement from which specimens are cut, showing a pure NR material. .

23

Uniaxial tensile test results of NR-GO specimens (a) The stress-strain curve
up to failure. (b) The small-strain fitting of the Mooney-Rivlin model (Eq. 2.1)
up to 30% strain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.10 Uniaxial tensile testing results of cycling specimens up to 1.4 MPa, in between 24 hour rests. (a)&(d) 0.00 wt%, (b)&(e) 0.25 wt%, (c)&(f) 0.50
wt%. (a)-(c) Stress-strain curves. (d)-(f) Mooney plots. . . . . . . . . . . .
min1

27

102 ).

2.11 Loading\unloading cyclic test at 50 mm


(strain rate of 5.55
(a) Strain versus time; (b) stress versus time for 0.25 wt.% against the constitutive model in Eq. 2.1; (c) stress versus strain at 50 mm\min. The model
for 0.00 wt.%, 0.25 wt.% and 0.50 wt.% are represented by the solid, dashed
and dash-dot lines respectively. . . . . . . . . . . . . . . . . . . . . . . . .

29

2.12 (a) Normalised dissipated energy D during cyclic tests estimated through
Eq. 2.2.b normalised with respect to dissipation of the natural rubber sample
(i.e. 0.00 wt.%). (b) Ratio between the dissipated and stored energies during
the cyclic test (estimated through Eq. 2.2.b). . . . . . . . . . . . . . . . . .

30

2.13 Comparisons of the stress-strain plots at different cross-head speeds. Tested


at 50, 100 and 200 mm min-1 for specimens v = 0.00 wt.%(continuous line),
0.25 wt.% (dashed line) and 0.50 wt.% (dash-dotted line). . . . . . . . . . .

31

2.14 (a) Stress-time plot of the relaxation tests for the different GO concentrations. The specimens were loaded at 100 mm min1 up to 50% strain, then
held fixed for 600 seconds. (b) The stress-time plot, without the initial loading ramp, normalised against the maximum stress. . . . . . . . . . . . . . .

32

2.15 Stress-ratetching of NR-GO specimens. (a) The first 30 seconds of the NR


specimen strain cycle. (b) NR stress-strain cycle up to failure. (c) The first
30 seconds of the NR-GO specimen (1.00.wt%) strain cycle. (d) The NRGO
(1.00wt.%) stress-strain cycle up to failure. . . . . . . . . . . . . . . . . .

33

List of figures

xix

2.16 The evolution of the maximum strain reached at the peak of each stress cycle
during stress-ratcheting between 0.15 MPa and 0.30 MPa of specimens of
NR-GO with concentrations of GO between 0 wt% and 1.00 wt%. . . . . .

34

2.17 (a) The neodymium magnetic setup for aligning magnetic fibres in viscous
PDMS solution, and (d) The dispersion of fibres through the thickness of
the specimen, observed by SEM on a fractured specimen. . . . . . . . . . .

37

2.18 The dumbbell test specimen according to ASTM D1708, and an inset showing the alignment of the fibres within a small central sample of the specimen
(Highlighted fibres shown for illustrative purposes only). . . . . . . . . . .

38

2.19 The fibre distributions from a specimen aligned between the four Neodymium
N52 magnets. (a) The fibre orientation distribution, fitted to Eq. 2.3 (b=4.3),
(b) The fibre length distribution, fitted to Eq. 2.4 (c=23.0, d=1.59). The averaged values of 6 specimens are b= 4.5 0.57, c=25.7 4.9 and d=1.61
0.15. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

2.20 Interface of the fractured PDMS specimens. (a) A poor interface showing
fibres without matrix attached [96]. (b) Trapped rubber between bundles of
fibres. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

39

2.21 Large strain tensile experimental results of magnetically aligned fibre reinforced PDMS, up to 60% strain. (a) Reinforcement angles 0 , 30 , 60
and 75 , (b) Reinforcement angles 15 , 45 and 90 . The 90% confidence
bars, calculated from 6 tested specimens at each angle, are shown at strain
intervals for each specimen. . . . . . . . . . . . . . . . . . . . . . . . . . .

42

2.22 Large strain tensile experimental results compared to the transversally isotropic
constitutive model, Eq. 2.5, up to 60% strain. Markers indicate the experimental results, lines indicate the constitutive model fitting. (a) Reinforcement angles 0 , 30 , 60 and 75 , (b) Reinforcement angles 15 , 45 and
90 . The 90% confidence bars, calculated from 6 tested specimens at each
angle, are shown at strain intervals for each specimen. . . . . . . . . . . . . 44
2.23 Small strain behaviour of the material, averaged from all specimens and
showing the standard deviation. The model (Eq. 2.10) fitting is shown to fit
the behaviour well. ( = 0.252, = 1.85). . . . . . . . . . . . . . . . . . .

44

2.24 Experimental micrographs of a reinforcing fibre during uniaxial tensile tests


at different strain levels. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

45

2.25 The experimental rotation of the fibres compared to the rotations predicted
h
i
by the model, obtained through Nansons formula, i.e. = tan1 3/2 tan(0 ) . 45

xx

List of figures
2.26 The lateral stretch, 2 vs longitudinal stretch, 1 of selected specimens of
each reinforcement angle, compared to the theoretical plane stress (dashed
line) prediction, i.e. 2 = 10.5 . . . . . . . . . . . . . . . . . . . . . . . .

46

2.27 (a) The COMSOL Multiphysics (ver. 5.0, 2014) dumbbell numerical model,
compared to the idealised constitutive behaviour, for angles 0 , 30 , 60 &90
(=0.252 MPa & =1.85). Dashed lines refer to the idealised constitutive
behaviour, markers refer to the behaviour of a dumbbell according to ASTM
D1708. (b-c) The deformation of the 45 dumbbell numerical model, indicating the lateral misalignment between the top and bottom of the gauge
length, at 20% strain for parametric values (b) =0.252 MPa & =1.85
(Lateral misalignment = 0.31mm) and (c) =0.252 MPa & =18.5 (Lateral
misalignment = 0.58mm). The parameter x refers to the lateral misalignment in millimetres, compared to the undeformed configuration. . . . . . .

47

3.1
3.2

3.3

3.4
3.5

3.6

The transverse isotropy of a fibre. Indicating the plane of isotropy orthogonal to the principal fibre direction. . . . . . . . . . . . . . . . . . . . . . .

59

The smallest section of the material microstructure that can model the properties of the overall composite is called the representative volume element.
An assumption of periodicity in the microstructure allows a further simplification to a unit-cell. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

63

A fibre reinforced thin elastomer specimen (t << L) can be simplified to a


multiple fibre RVE, assuming plane stress in the through thickness direction.
In addition, a platelet reinforced elastomer can be assumed to have plane
strain in the through thickness direction, if the thickness is much greater
than the length, i.e. t >> L. In both cases, a further simplification to a
single reinforcement RVE can be made if an assumption of periodicity and
dilute inclusions is made. . . . . . . . . . . . . . . . . . . . . . . . . . . .

66

A schematic representation of the RVE in the cartesian coordinates plane


E1 ,E2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

A schematic representation of the different models examined: from the full


3D RVE to the single platelet one. In each case, the model is simplified
further. (a) The platelet reinforced elastomer block, (b) The fibre reinforced
thin elastomer specimen. . . . . . . . . . . . . . . . . . . . . . . . . . . .

82

A thin elastomer (length much greater than the thickness) with fibres restricted to a single plane (E1 , E2 ) can be modelled as a 2D plane stress
structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

85

List of figures
3.7

xxi

The two RVE configurations in the undeformed configurations. (a) The


multiple aligned fibre RVE model for 0 = 30 (2D-MFRVE). The inset
shows the mesh surrounding a reinforcement. (b) The single fibre reinforced
RVE model for 0 = 30 (2D-SFRVE). The inset shows the mesh around the
inclusion edge. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

86

The distribution of the fibre angles, as defined by the parameter b in Eq. 3.75.
The images show an arrangement whereby the fibres are aligned around the
0 = 0 configuration. The fibre distributions of the 2D-MFRVE are shown
for b = 5 and the aligned configuration. . . . . . . . . . . . . . . . . . . .

87

Cauchy stress 11 against stretch for simple tension in terms of reinforcement orientations 0 = {0 , 25 , 45 , 65 , 90 }. In all cases, (r) = 2 MPa,
f = 1000, AR = 40 and vol = 1.2%. . . . . . . . . . . . . . . . . . . . . .

87

3.10 (a) The tangent modulus during tensile loading of initial reinforcement angles 0 = {0 , 25 , 45 , 65 , 90 }. (b) The fibre change of orientation during
simple tension for different initial configuration of the reinforcement 0 =
{0 , 25 , 45 , 65 , 90 } plotted against the predicted behaviour, Eq. 3.69. In
all cases the parameters of the RVE are: AR=40, vol%=1.2, (r) = 2 MPa
and f = 1000. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

89

3.11 Cauchy stress, 12 , against shear, , for simple tension in terms of reinforcement orientations 0 = {0 , 25 , 45 , 65 , 90 }. In all cases, (r) = 2 MPa,
f = 1000, AR = 40 and vol = 1.2%. . . . . . . . . . . . . . . . . . . . . .

89

3.12 (a) The tangent modulus during simple shear loading of initial reinforcement angles 0 = {0 , 25 , 45 , 65 , 90 }. (b) The fibre change of orientation during simple tension for different initial configuration of the reinforcement 0 = {0 , 25 , 45 , 65 , 90 } plotted against the predicted behaviour,
Eq. 3.69. In all cases the parameters of the RVE are: AR=40, vol%=1.2,
(r) = 2 MPa and f = 1000. . . . . . . . . . . . . . . . . . . . . . . . . .

90

3.13 The change in initial modulus during simple tension for different initial
configuration of the reinforcement 0 = {0 90 }, (r) = 2 MPa and
f = 1000 against Eq. 3.69 . . . . . . . . . . . . . . . . . . . . . . . . . .

91

3.14 A comparison between the uniaxial cauchy stress of the multiple and single
fibre 2D plane stress RVE models up to 60% strain. . . . . . . . . . . . . .

92

3.15 The variation of the stress-strain curves with varying amounts of fibre distribution for different angles. (a) 0 , (b) 30 , (c) 60 , (d) 90 . . . . . . . . .

93

3.8

3.9

xxii

List of figures

3.16 The small strain effects of a distribution of fibres in the RVE. (a) The reduction in transverse isotropy is shown by the reduction in the stiffness difference between longitudinal and lateral directions, (b) Show the effect of
increased alignment on the stiffness for each angle. . . . . . . . . . . . . .

93

3.17 The transverse isotropy of the model is shown by the close fitting to Eq. 3.22
for different values of the heterogeneity contrast (a) f = 10, (b) f = 100 (c)
and f = 1000. The other values are 0 = 35 , (r) = 2 MPa, AR = 40
and = 0.8%. The planes show the constitutive model fitted to the data.
The plotted dots show the numerical values of the strain energy function as
computed by Abaqus for simple tension/compression (red dots) and simple
shear (magenta dots). . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

95

3.18 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 0 , (c)-(d) 0 = 45 and
(e)-(f) 0 = 75 . The other values were (r) = 2 MPa and f = 1000. The
remaining angles can are found in the Appendix B.1 . . . . . . . . . . . . .

96

3.19 Long Platelets oriented in a single plane (E1 , E2 ) of a block elastomer,


wherein the thickness is much greater than the length, can be modelled by a
2D plane strain assumption. . . . . . . . . . . . . . . . . . . . . . . . . . .

98

3.20 The two RVE configurations in the undeformed configurations. (a) The
3D multiple aligned platelet RVE model for 0 = 30 (3D-MPRVE). The
inset shows the mesh surrounding a reinforcement. (b) The single platelet
reinforced RVE model for 0 = 30 (2D-SPRVE). The inset shows the mesh
around the inclusion edge. . . . . . . . . . . . . . . . . . . . . . . . . . .

99

3.21 Comparisons between the 3D-MPRVE and 2D-SPRVE for the cauchy stress
difference (11 33 ) against strain for simple tension and compression,
and the cauchy shear stress (12 ) against the shear strain. Displayed for
reinforcement orientations, (a) 30 , (b) 45 , (c) 75 , (d) 90 . . . . . . . . . 101
3.22 Cauchy stress difference against stretch for (a) simple compression and
(b) tension in terms of reinforcement orientations, 0 = {0 , 25 , 45 , 65 , 90 }.102
3.23 (a) Tensile tangent modulus against stretch in terms of reinforcement orientations, 0 = {0 , 25 , 45 , 65 , 90 }. (b) The platelet change of orientation for different initial configurations of the reinforcement against Eq. 3.38.
The absence of points for large tensile and compressive strains both at 0
and 90 is caused by the premature failure of the FE simulation. . . . . . . 103

List of figures

xxiii

3.24 Shear stress against (a) negative and (b) positive shear strain in terms of
reinforcement orientations 0 = {0 , 25 , 45 , 65 , 90 }. . . . . . . . . . . . 104
3.25 (a) Tangent shear modulus against shear strain in terms of reinforcement
orientations, 0 = {0 , 25 , 45 , 65 , 90 }. (b) The platelet change of orientation during simple shear for different initial configuration of the reinforcement against Eq. 3.38. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.26 The transverse isotropy of the model is shown by the close fitting to Eq. 3.22
for different values of the heterogeneity contrast (a) f = 10, (b) f = 100 (c)
and f = 1000. The other values are 0 = 35 , (r) = 2 MPa, AR = 40
and = 1.2%. The planes show the constitutive model fitted to the data.
The plotted dots show the numerical values of the strain energy function as
computed by Abaqus for simple tension/compression (red dots) and simple
shear (magenta dots). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
3.27 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 25 , (c)-(d) 0 = 45 and
(e)-(f) 0 = 90 . The other values were (r) = 2 MPa and f = 1000. The
remaining angles can are found in the Appendix B.2 . . . . . . . . . . . . . 107
3.28 Effective shear modulus c in terms of platelet orientations 0 for various
heterogeneity contrast {1, 10, 100, 1000} against Eq. 3.43. The other parameters were (r) = 2 MPa, = 1.2% and AR = 40. Note that the case
f = 1 corresponds to the homogeneous solution. . . . . . . . . . . . . . . . 108
3.29 Dependence of the constitutive parameters (a) & (c) and (b) & (d) on the
heterogeneity contrast, f , the aspect ratio, and the volume fraction against
the model introduced in [160] and the one introduced with Eqs. (3.76)-(3.78). 109
4.1

Examples of magnetic actuation and locomotion in the literature. (a) Inchworm like motion of a composite sample by varying the magnetic field on
a ratchet strip [132], (b) Modular, finger-like actuator responsive to a magnetic field [202], (c) A modular micro-actuator, in which each module has
a different reinforcement direction and responds to a magnetic field accordingly [127], (d) A soft cellular elastomer with embedded magnets that collapses in a magnetic field [236]. . . . . . . . . . . . . . . . . . . . . . . . 124

4.2

NiC fibre characterisation under SEM. (a) good fibre (b) agglomeration of
fibres (pre-cure) (c) uneven nickel coating (d) Broken fibre (e) broken interface (g) almost fully broken interface of small fibre. . . . . . . . . . . . . . 129

xxiv

List of figures

4.3

(a) The fibre length distribution, fitted to Eq. 2.4 (c=23.0, d=1.59). The
averaged values of 6 specimens are c=25.7 4.9 and d=1.61 0.15 (b) Micrograph image of the aligned fibres, showing the variation in length and
orientation (Selected fibres highlighted in white). . . . . . . . . . . . . . . 129

4.4

(a) The curing PDMS specimen during the alignment procedure between
the coils of the Electromagnet (Newport Instruments Electromagnet Type
C), (b) The fibre distribution after 24 hours under various magnetic fields
(|B0 | = {0.01, 0.03, 0.05, 0.08} T), fitted to the fibre distribution function,
Eq. 2.3 (b=0.44, 0.92, 1.58, 2.95). . . . . . . . . . . . . . . . . . . . . . . 131

4.5

Orientation distribution of the fibre as predicted by the model (4.10) at different times, tr , for (a) = 5 and (b) = 0.5. A random orientation of the
fibres is assumed at tr = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . 134

4.6

Contour plots of the theoretical minimum magnetic field intensity B0 min


(eq. 4.13) to have 90% of the fibres oriented at 5 deg. B0 min is plotted
against (a) the initial curing rate 0 = 0 /c and the fibre aspect ratio AR
(for a = 2.20 103 ), and (b) the initial curing rate 0 = 0 /c and the
fibre magnetic susceptibility a (for ar = 25). The labels show the magnetic
filed intensity in Tesla. The dashed red area represents the region where B0
is higher than 1 T. Due to the increasing viscosity of the solution, the results
are shown for the steady state (i.e. cured) condition. . . . . . . . . . . . . . 135

4.7

The effect of fibre damage and fibre morphology on the magnetic susceptibility. Orientation of different fibre aspect ratios after 60 minutes under the
homogeneous magnetic field of an electromagnet at a field strength of 0.08T.
Fitted to Eq. 2.3 (a) Orientation of fibres of aspect ratio: {0 < AR 12.7},
b=0.7 (b) {12.7 < AR 24.1}, b=1.4 (c) {24.1 < AR}. b=37.7. . . . . . . 136

4.8

Fibre distribution measured at intervals over a period of 24 hours, showing


the fitting of Eq. 4.10. (a) One value of is fitted to the interval data,
o1 = 15.4. (b) Three characteristic orientation times improve the fitting of
the data, o1 = 20.2, o2 = 15.7, o3 = 12.6 . . . . . . . . . . . . . . . . . . 137

4.9

(a) The Neodymium magnetic set-up to create an approximately homogeneous field for alignment of the reinforcement, (b) The magnetic field
strength (0.07-0.08T) approximation calculated using FEMM (ver.4.2), (c)
The cured specimen; the arrow indicates the in-plane direction of the reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138

List of figures

xxv

4.10 (a) The post-cure fibre distribution from a specimen aligned between the
four Neodymium N52 magnets, fitted to Eq. 2.3 (b=4.3). (b) Manufactured
defects and fibre interaction during the curing process cause some agglomerations to form. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
4.11 (a) The neodymium magnetic set-up to create an out-of-plane reinforcement, (b) The magnetic field strength approximation calculated using the
FEMM software [170], (c) The cured specimen; the arrow indicates the inplane direction of the reinforcement. The centre dot indicates the fibres are
oriented orthogonal to the viewed plane. . . . . . . . . . . . . . . . . . . . 140
4.12 The out-of-plane cured specimen has a fibre reinforcement angle that varies
along the length of the specimen. The length along the specimen is shown
in each image. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
4.13 (a) The non-homogeneous neodymium magnetic set-up to create an, inplane, variable angle reinforcement, (b) The magnetic field strength approximation calculated using FEMM (ver.4.2), (c) The cured specimen; the arrow indicates the in-plane direction of the reinforcement. . . . . . . . . . . 141
4.14 Experimental setup of the electromagnet for static actuation of the specimens, with the magnetic field directed orthogonal to the plane of the fibres.
Inset: A sample specimen of 0 = 90 , actuated in the static magnetic field. 143
4.15 Static Actuation of specimens with reinforcement angles, 0 , at 0 , 30 , 60
& 90 . Specimens of length 27.5mm, width 7mm and thickness 0.5mm are
presented to a homogeneous magnetic field. A transition from pure bending
for 0 , to pure twisting for 90 is observed. For the orientation of the axes
see Fig. 4.18. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
4.16 Static bending/twisting actuation of specimens with different fibre alignments; (a) 0 = 0 , (b) 0 = 15 , (c) 0 = 30 , (d) 0 = 45 , (e) 0 = 60 ,
(f) 0 = 75 , (g) 0 = 90 . The dominant actuation angle (bending or twisting) is recorded in each case. . . . . . . . . . . . . . . . . . . . . . . . . . 146
4.17 The magneto-mechanical properties of the magnetically responsive specimens, in a homogeneous magnetic field of maximum strength 0.09 Tesla.
(a) The perpendicular arrangement, whereby 0 is orthogonal to the field.
(b) The transverse arrangement, whereby the field lines are in the plane of
the reinforcements. At 0 = 90 the field is parallel to the field lines. . . . . 147
4.18 Schematic representation of the beam model. (a) The undeformed beam
showing the reinforcement angle in the x-z plane, 0 . (b) The twisting angle,
. (c) The bending angle, . . . . . . . . . . . . . . . . . . . . . . . . . . 148

xxvi

List of figures
(crit)

at which the undeformed configura4.19 Critical magnetic field intensity By


tion becomes unstable, assuming the model (4.21). (Parameters: E(0 ) and
G(0 ) taken from the results of the mechanical testing (i.e. = 0.252MPa, =
1.85), I = 1.261013 (m4 ), J = 4.771013 (m4 ), L = 27.5103 (m), a
= 2.20103 and f = 6 %). The black dots with error bars represent the results of the experiment carried out by the authors for specimens with fibres
at 0 = {0 , 15 , 30 , 45 , 60 , 75 , 90 }. . . . . . . . . . . . . . . . . . . . 151
4.20 The bending and twisting angles achieved for an imposed orthogonal magnetic field at a field strength just greater than at the occurrence of instability
(crit)
(i.e. 1.01By ), assuming the model (4.19). (Parameters: E(0 ) and G(0 )
taken from the results of the mechanical testing (i.e. = 0.252MPa, =
1.85), I = 1.261013 (m4 ), J = 4.771013 (m4 ), L = 27.5103 (m), a
= 2.20103 and f = 6 %). The orange dots and purple dots, represent the
bending and twisting experimental results, respectively, for specimens with
fibre angles at 0 = {0 , 15 , 30 , 45 , 60 , 75 , 90 }. The solid line represents the model (4.19). The error bars represent the experimental prediction
of the actuation angle, when extrapolated from the magnetic field 1% before and after the occurence of the instability (Given by the dashed line in
Fig. 4.16). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
A.1 The stress-strain behaviour of 7 selected specimens of NiC fibre reinforced
PDMS samples, shown up to mechanical failure. . . . . . . . . . . . . . . 183
B.1 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The plots represent (a)-(b) 0 = 5 . The other values were (r) =
2 MPa and f = 1000. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
B.2 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 10 , (c)-(d) 0 = 15 , and
(e)-(f) 0 = 20 . The other values were (r) = 2 MPa and f = 1000. . . . . 186
B.3 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 25 , (c)-(d) 0 = 30 , and
(e)-(f) 0 = 35 . The other values were (r) = 2 MPa and f = 1000. . . . . 187

List of figures

xxvii

B.4 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 40 , (c)-(d) 0 = 50 , and
(e)-(f) 0 = 55 . The other values were (r) = 2 MPa and f = 1000. . . . . 188
B.5 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 60 , (c)-(d) 0 = 65 , and
(e)-(f) 0 = 70 . The other values were (r) = 2 MPa and f = 1000. . . . . 189
B.6 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 80 , (c)-(d) 0 = 85 , and
(e)-(f) 0 = 90 . The other values were (r) = 2 MPa and f = 1000. . . . . 190
B.7 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 0 , and (c)-(d) 0 = 5 .
The other values were (r) = 2 MPa and f = 1000. . . . . . . . . . . . . . 191
B.8 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 10 , (c)-(d) 0 = 15 , and
(e)-(f) 0 = 20 . The other values were (r) = 2 MPa and f = 1000. . . . . 192
B.9 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 30 , (c)-(d) 0 = 35 , and
(e)-(f) 0 = 40 . The other values were (r) = 2 MPa and f = 1000. . . . . 193
B.10 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 50 , (c)-(d) 0 = 55 , and
(e)-(f) 0 = 60 . The other values were (r) = 2 MPa and f = 1000. . . . . 194
B.11 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 65 , (c)-(d) 0 = 70 , and
(e)-(f) 0 = 75 . The other values were (r) = 2 MPa and f = 1000. . . . . 195
B.12 Comparison between model (3.22) and FE results for different values of the
orientation angle in terms of strain energy vs. strain and stress vs. strain
curves. The different plots represent (a)-(b) 0 = 80 , and (c)-(d) 0 = 85 .
The other values were (r) = 2 MPa and f = 1000. . . . . . . . . . . . . . 196

xxviii

List of figures

B.13 Comparisons between the 3D-MPRVE and 2D-SPRVE for the cauchy stress
difference (11 33 ) against strain for simple tension and compression,
and the cauchy shear stress (12 ) against the shear strain. Displayed for
reinforcement orientations, (a) 0 , (b) 15 , (c) 60 . . . . . . . . . . . . . . 197
C.1 The cyclic testing of a 0 specimen. The specimen is cycled 10 times, to
10% strain at a loading rate of 10mm min1 . . . . . . . . . . . . . . . . . . 199

List of tables
2.1
2.2

2.3

4.1

NR-GO nanocomposite tensile properties at concentrations between 0.0 wt%


and 1.0 wt%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Youngs modulus, E, calculated from based on the fitting of the Mooney
Rivlin formulation (Eq. 2.1) to the stress-strain response up to 30% strain,
alongside an estimate of the strain value at which the upturn in stress occurs
(associated with the onset of SIC [37]), which is estimated from a Mooney
plot. The testing consisted of the first uniaxially stretching the specimen up
to 1.4MPa and then resting it, followed by again testing to 1.4 MPa after
24 hours; resulting in two sets of results. . . . . . . . . . . . . . . . . . .
The average mechanical results of specimens tested to failure, standard deviations shown in brackets. Indicating, the Youngs modulus, E, calculated
from based on the fitting of the Transversely Isotropic constitutive model
(Eq. 2.10) to the stress-strain response up to 30% strain, the strain-to-failure
and the strength of the specimens. . . . . . . . . . . . . . . . . . . . . . .

25

26

41

The percentage of fibres aligned to the magnetic field direction in experimental testing. The model fittings are shown in square brackets, when considering three characteristic orientation times (o1 = 20.2, o2 = 15.7, o3 =
12.6). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

Acronyms / Abbreviations
AFM Atomic Force Microscopy
CB

Carbon Black

FE

Finite Element

FEA Finite Element Analysis


FT IR Fourier transform infrared spectroscopy
GE

Graphene

GO

Graphene Oxide

H T Halpin-Tsai
LAA

Laminate Analogy Approach

MFRV E Multiple Fibre Representative Volume Element


MPRV E Multiple Platelet Volume Element
NR

Natural Rubber

PDMS Polydimethylsiloxane Silicone


RV E Representative Volume Element
SAXS Small angle x-ray scattering
SEM Scanning electron microscopy
SFRV E Single Fibre Representative Volume Element
SIC

Strain-Induced Crystallisation

xxxii

Acronyms / Abbreviations

SPRV E Single Platelet Representative Volume Element


T EM Transmission electron microscopy
Tg

Glass Transition Temperature

WAXD Wide angle x-ray diffraction

Chapter 1
Introduction
1.1

Motivations

Optimisation in the design of engineering structures is a vital requirement. The optimisation can involve consideration of many variables that include, e.g. time, cost and weight.
However, this is not possible without (1) a good understanding of the processes involved
and (2) a knowledge of the design space available. Research is continually involved in the
expansion of both of these areas, and often takes inspiration from nature [61, 62]; which has
had many thousands of years to optimise its behaviour to its environment.
The mechanical properties are generally regarded as of primary concern in the majority of structural applications, however there is also great interest in other properties such
as electrical conductance [15] and magnetic responsiveness [202]. In fact, it is possible to
control multiple properties simultaneously to develop multi-functionality. In composite materials, major advances are being made to develop novel and multi-functional materials, for
instance morphing structures [234] and self-healing composites [123].
In elastomer materials, the design space can be expanded to consider the large strain
behaviour, although the continuous fibre reinforcements used in composite materials are not
suitable for withstanding large strains. Hence, discrete reinforcements such as carbon black
(CB) are commonly used to reinforce the material, although anisotropic reinforcements are
increasingly used due to the excellent excellent physical and mechanical properties of the
resulting elastomers [14]. However the reinforcing potential of anisotropic reinforcements is
rarely maximised due to the homogenous distributions that are typically used. To maximise
their potential, deliberate placement of the reinforcement can be considered that will create
an anisotropic material that optimises it to, for example, carry load or resist damage.
There are many of example in nature of optimising the architecture of the constituents
to achieve high performance. Trees construct their architecture in such a manner by ori-

Introduction

Fig. 1.1 Examples of anisotropic architecture in nature. (a) A spruce tree, and (b) The fibril structure at the spruce tree branch joint (Adapted from [27]). (c) Venuss Fly Basket
Euplectella Aspergillum [235], and (d) A section of the mineralized skeletal cage of Euplectella Aspergillum (Adapted from [172]). (e) The marine mussel, anchored by the byssus
[93], and (f) A Byssal thread, with a soft fibrous core and protective cuticles (Adapted from
[218]).

enting the fibrils, and tailoring the response of the material [59]. In this way, rather than
homogenising the material properties, the strain field can be homogenised; subsequently reducing the stress concentrations and delaying failure around features such as branch stems
[178] (See Fig. 1.1a&b). Similar examples include the Venuss Fly Basket (Euplectella Aspergillum), which makes use of low strength and brittle glass to create a stable structure
that is able to withstand ocean currents [51] (See Fig. 1.1c&d) and the byssus of marine
mussels, that has a hard protective coating that withstands strains as high as 100% [94] (See
Fig. 1.1e&f).

1.2 Technical Challenges

1.2

Technical Challenges

Control of the reinforcements in elastomer materials, in terms of orientation and spatial position, could provide potential weight savings, increased performance and maximum use
of multi-functional capabilities (e.g. electrical conductance, magnetic susceptibility). In
this study, an understanding of the stress-strain behaviour will generally only be considered
up to moderate strain values (defined here as 60%); due to the technical challenges that
remain unresolved in this strain range, and the large number of applications that operate
within this range.
The identified key technical challenges and unanswered questions are highlighted below:
Despite the recent interest in nano-dimensioned fillers, many aspects of the effects of
anisotropic nano-reinforcements on the mechanical and viscoelastic properties of elastomers
remain unclear. This is particularly relevant to its acceleration of formed stiffening networks
compared to spherical particles, and the induced anisotropic behaviour that may occur at
large strain. Anisotropic reinforcements are known to cause instabilities under compression,
however the rotation of the reinforcement is also expected to affect the large strain behaviour
by changing the stress transfer mechanism.
A transversely isotropic material has properties which are identical in one plane, with
different properties in the direction normal to this plane. This is the simplest form of
anisotropy, and is often used in industrial applications. However the manufacture of these
materials with dis-continuous fillers is challenging, as it requires the alignment of the reinforcement in a single plane. Further complications exist when spatial change in the alignment angle exists. A number of strategies exist that can help to predict and control the
alignment of fillers in a viscous solution, however the lack of experimental data to validate the approaches raises questions regarding the extend to which such models are able to
predict the behaviour.
Modelling the behaviour of these dis-continuous filler reinforced elastomer materials
can also be challenging at large strain. Complex models are often required, particularly
when considering material phenomenon, such as strain-induced crystallisation, that result
in highly non-linear behaviour. However these models can be computationally expensive
and unstable. A simple model is generally preferred, however it is important to be aware
of the limitations of each approach and assess the degree to which it is able to describe the
material behaviour.
Magnetic fibres are added to an elastomer material, and so the resultant material is responsive to a magnetic field. When the fibres are aligned, the magnetic responsiveness and

Introduction

mechanical properties both depend upon the fibre orientation, and are expected to behave
in a complimentary manner, i.e. the mechanical properties affect the magnetic response. It
is expected that interesting actuation behaviour can be achieved by careful consideration of
the reinforcement direction, in relation to the magnetic field direction.

1.3

Objectives

To investigate the benefits of anisotropic reinforcements, in particular it is expected


that these benefits will be most apparent when the reinforcements are aligned.
To demonstrate the controlled manufacture of aligned reinforcements.
To understand the mechanical behaviour up to moderate strains (60%).
To see how best to model the behaviour and understand the limitations of each method.
To demonstrate the usefulness of anisotropic behaviour in the context of actuation,
and the complimentary manner in which mechanical/magnetic response will interact.

1.4

Structure of the Thesis

This thesis discusses the benefits of anisotropic reinforcements on the mechanical and multifunctional response of elastomers, and is organised into three distinct sections. Each section
has a self-contained literature review that is intended to focus the reader on the relevant literature concerning the chapter ahead. It is organised as follows:
In the second chapter, "Experimental Mechanical Characterisation", the experimental
results of elastomers containing anisotropic reinforcements are discussed. In particular, the
effects of the orientation and rotation of the reinforcements is explored. This chapter introduces the reader to the mechanical properties of the anisotropic materials, and shows that
a simple constitutive model is unable to describe the material behaviour at larger strains.
A literature review is included at the beginning of the chapter to give the reader a general
introduction to rubber technology and a number of its physical characteristics.
In the third chapter, "Numerical Modelling", a set of representative volume elements
are used to explore the constitutive models applicability to the specific problem of discontinuous fibres in an elastomer. The numerical model considers the micro-scale behaviour

1.4 Structure of the Thesis

of the constituents, and the constitutive model acts as an intermediary step towards the experimental testing of chapter 1. Only strain effects up to 60% are considered. A literature
review at the beginning of the chapter discusses the progress of numerical modelling towards describing complicated non-linear effects at large strain; illustrating the difficulties
and limitations of both constitutive modelling and numerical modelling.
In the fourth chapter, "Magnetic Response", the effects of the magnetically responsive
fibres are discussed in the material when the PDMS matrix is uncured and cured (i.e. during
alignment, and the subsequent actuation of the material upon cure). As such, a literature
review is included at the beginning of the chapter to display the state-of-the-art in this field,
and the projected applications which it is intended for.

Chapter 2
Experimental Mechanical
Characterisation
2.1

Introduction

Rubber is an important industrial material with remarkable properties, such as excellent


dissipation and the ability to deform to large strain [42], that see it used in applications that
range from car tyres to vibration mounts to artificial hearts [162].
It is able to recover quickly to its approximate original shape after large deformations
[241], which is a quality shared by a number of materials, that includes some biological
tissues. The behaviour is as a result of the long chains in the material, and the unique way
in which they interact. For example, in natural rubber (NR) there are three main characteristics of the chemical structural composition that are thought to be most responsible for this
behaviour, and differentiate it from a viscous liquid [241]: (1) the presence of long polymer
chains with freely rotating links, (2) weak secondary forces between polymer chains that
allow it to take a variety of statistical configurations without being massively impeded by a
large number of primary bonds (e.g. covalent bonds), (3) a limited 3D network formed by
the interlocking of molecules at certain points, allowing it to return to its original shape.
This interlocking 3D network is the result of a cross-linking agent added to the elastomer. In natural rubber it is known as vulcanisation and typically involves the addition of
sulphur to form cross-links, although in unvulcanised NR there is no direct 3D network, and
instead it is thought that the chain entanglements and end-linking networks help to form a
pseudo network [4].
An important phenomenon of NR, and some other rubbers, is the ability to crystallise at
temperatures approaching 0 C, which involves a gradual reordering of the polymer chains

Experimental Mechanical Characterisation

to form crystalline regions (see Fig. 2.1a). This change in structure is a first order transition,
and is therefore accompanied by an increase in density and the release of latent heat [241].
It can result in the ordering of up to 30% of the polymer chains, after which the motion of
polymer chains is restricted, resulting in a hardening/reinforcing of the material analogous
to filler reinforcement. The rate of crystallisation is much slower in vulcanised NR than in
unvulcanised NR because of the relative decrease in chain motion possible.
A related, but independent, type of crystallisation that occurs due to the ordering of
the chains in the direction of an applied stretch is strain-induced crystallisation (SIC) (see
Fig. 2.1b). This reinforces the material in the direction of the applied strain (analogous to
transversely isotropic fibre reinforcement), causing a stiffening of the material at high strain
that can be observed as an upturn in stress on the stress-strain curve [226] (1) . However,
upturns in stress are observed in non-crystallising rubber such as PDMS due to the effects
of limited extensibility of the polymer chains. The overall result can be a toughening of the
rubber, whereby the initial softening protects the material and reduces stress concentrations,
but as strain hardening occurs it protects areas from damage and further deformation [252].

Fig. 2.1 Schematic representation of two types of crystallisation encountered in rubber materials. (a) Static crystallisation. (b) Strain-induced crystallisation.
The rubbery behaviour of elastomers is temperature dependent, and due to them operating above the glass transition temperature, Tg. The behaviour depends on the random
thermal motion of polymer chains by the rotation of single bonds, an effect that is partially
restricted in the materials. However, the addition of thermal energy aids this motion. Con1 The

upturn in stress is an indication of SIC rather than a direct measure of it, and does not necessarily
indicate the initiation point of SIC.

2.1 Introduction

versely, however, by restricting the thermal energy to the system it is possible to reach a
temperature at which the material cant behave as such, and effectively behaves as a solid,
i.e. glassy. The Tg alters depending on the materials internal structure, but there is no
change in density, volume or in fact any geometrical structural change within the material
during the transition, and so it is termed a secondary phase transition. It does however involve an increase in the elastic properties (i.e. stiffness) as well as a change in the expansion
coefficient [241].
The addition of filler materials to rubber is the general method of improving the mechanical properties, and in fact the reinforcements are almost ubiquitous to rubbers (i.e. consider
that rubber is generally thought of as black, which is the case only because of the addition
of large quantities of carbon black). The primary purpose of filler materials is to reinforce
and improve the mechanical and physical properties of a material, however non-reinforcing
fillers can also be used to bulk out a material. For instance, the use of CB in rubbers also
reduces both the cost and weight of the material.
The effectiveness of reinforcements is largely dependent on the state of dispersion and
distribution of the fillers. A good distribution of the fillers requires that they are evenly
spaced in the material to give a homogenous mechanical response, but to maximise the surface interaction to the elastomer it is important to separate agglomerations and maximise
dispersion [265] (See Fig. 2.2). In this way it is possible to avoid defect sites, voids and
stress concentrations; all of which will reduce the mechanical properties and lead to premature failure [198].

Fig. 2.2 Distribution vs Dispersion. (a) A homogeneous distribution of the fillers, (b) A
homogeneous distribution of the fillers with an increased dispersion.
The morphology of the filler can also have a significant effect upon the properties, and
can be separated into three main categories: (1) Equi-axed (spherical) [9, 156], (2) fibres
such as carbon fibres [66] and carbon nanotubes (CNT) [13], and (3) platelets such as

10

Experimental Mechanical Characterisation

graphene (GE) [226] and layered clays [76]. The choice of reinforcements generally depends upon the ease of manufacture and material cost, however anisotropic reinforcements
can be tailored to give a controlled anisotropic response to the material [52] and have a
higher specific surface area than equi-axed particles. The surface area, in particular, is
a well considered advantage of a reinforcement and has led to an explosion in the use
of nano-dimensioned materials [14, 262], particularly since the discovery of methods to
produce graphene [182] and CNTs [109]. Their large specific surface area and excellent
physical properties give the engineer a large amount of control over the bulk properties of
the material, if they can control the transfer of stress between the matrix and reinforcement.
The extent of this increase can be seen in Fig. 2.3, which shows that a reduction of the particle size leads to an increase in the specific surface area and a reduction in the inter-particle
spacing. This effect is further exacerbated by an increase in the aspect ratio of the particle
towards a fibre or platelet morphology. This has led to an enormous amount of research
on improving the interface of nano-reinforcements, as a small change in the interface has a
comparatively large effect compared to micro reinforcements and can help to improve the
dispersion of the fillers (2) .

Fig. 2.3 The effect of particle size on the interfacial area. (a) Large particles of diameter,
d. (b) Small particles of diameter, d/2. Particles of equivalent area have an interfacial zone
over twice the size when the diameter is halved. The inter-particle spacing is also decreased.

2.1.1

The Benefits of Homogeneously Distributed Reinforcements

The mechanical benefits of elastomers reinforced by stiff filler materials are well documented. The benefit is dependent on both the mechanical properties of the reinforcement,
2 The

interfacial region is the region of altered properties, in comparison to the main constituent parts, that
surround the fillers due to, for example, chain mobility, altered chemistry or crystallinity [2].

2.1 Introduction

11

and its adhesion to the elastomer i.e. the interfacial properties. Hence, the recent interest in nano-dimensioned materials in elastomers is unsurprising [34, 100, 264], as their
large surface areas and excellent physical properties [262] help to form an excellent transfer
mechanism between the elastomer and the reinforcement. In order to affect these property
enhancements, it is important to carefully control both the processing techniques and the
interfacial properties of the materials. It is generally considered that a good dispersion of
reinforcements is advantageous, as shown in [265] where a better dispersion is achieved
by sonication of an aqueous dispersion of NR and Graphene Oxide (GO), followed by insitu reduction to GE. However, by carefully controlling the formation of the reinforcement
networks within the rubber, it is also possible to improve the small-strain mechanical properties, as well as the electrical and thermal conductivity, at the expense of the large strain
ability to withstand rupture [196]. In this case, the strong inter-particle links at small strain
become stress concentrations at large strain. Although, in general it is agglomerations which
have a significant effect on the large strain performance due to the formation of cracks and
voids [15].
In addition to the processing of the reinforced rubber, the interface between the constituents is equally important. Reduced GO (i.e. GE) can be produced with defects and
chemical bonding sites that reduce the property of the platelet, but improve the interfacial
properties by creating extra bonding sites and mechanical interlocking, and which help to
enhance the cross-linking process [100]. In fact, the interfacial chemistry is very similar in
most reinforcing materials and involves forming an immobilized region of rubber around
the reinforcement; the effects of which can stretch far into the polymer chains of the rubber
if the functionalization of the reinforcement is good [200]. However, this is only possible when the interface between the rubber polymer and the reinforcement is good. This
can be achieved by modification of the reinforcement prior to mixing [23, 143], although
advantages in dispersion have been demonstrated with in-situ functionalization [200].
The result is that incredible improvements in rubber mechanical properties (e.g. stiffness and strength) can be achieved by the addition of only a small amount of nano-filler
[13, 196]. The consequence, however, is that for heavy polymers, as elastomers typically
are, the use of cheap fillers such as CB have the effect of reducing not just the price but also
the weight of the final material. It is for this reason that there is also work involving the
combination of cheap conventional fillers, such as CB, with nano-fillers such as GE. This
combination has been shown to produce beneficial effects in the properties of NR, showing
an appreciable increase in the dynamic, thermal and mechanical properties of the material
[164].

12

Experimental Mechanical Characterisation

Large Strain Effects


One of the advantages of using discontinuous reinforcements, in contrast to, for example, continuous carbon fibres used in conventional composite materials, is that they allow
the rubber to achieve the large strains that are synonymous with rubbers and are a major
advantage in many of their applications. This has significant effects upon the large strain
performance of reinforced elastomers, which is evidenced particularly in the upturn in stress
at large strain, which occurs at lower strains due to strain amplification effects of the reinforcements (3) [13, 15, 265].
The reinforcement of NR with CB has been shown to affect this change [186], and
although the exact mechanism is uncertain [239], it is clear that the reinforcement promotes
an early formation of SIC due to a strain amplification effect. However, the final percentage
of SIC is independent of reinforcement concentration, suggesting that the reinforcement
merely initiates SIC earlier and does not help to directly promote its growth [186]. Whilst
wide angle X-ray diffraction (WAXD) is able to directly measure the onset of crystallization
[240], the use of a mooney-plot is able to give a useful representation of the stress-strain
curve that more clearly indicates the upturn in stress [265]; this upturn is associated with
the onset of SIC [186]. SIC is also found to delay rupture in elastomers, as evidenced by
the dramatic decrease in tensile strength when SIC is suppressed by testing at an elevated
temperature [79].
A similar upturn in stress at high strain can be observed in non-crystallizing elastomers,
such as PDMS. In this case the limited extensibility of short chains causes an upturn in
stress, whilst longer chains in the network help to delay rupture; in this way it can be described as a bimodal network [252]. It results in a less pronounced increase in the tangent
modulus than SIC in NR, however it is unaffected by temperature.
Whilst crystallization can have significant large strain effects in some elastomer materials, the direct effect of anisotropic reinforcements can also be evidenced. Whilst CB shows
little to no evidence of orientation at large strain, GE has been observed to orient by small
angle X-ray scattering (SAXS) [186] due to its anisotropic morphology. This orientation
can assist in the formation of crystallites, as demonstrated by CNTs oriented at large strain
in NR [14, 250], or in a breakdown of the electrical networks [14]. However it has been
demonstrated that, by careful knowledge of the mechanisms that form the networks, repeated loading of CNT reinforced epoxy can demonstrate improvements in mechanical and
electrical properties by creating a transversely isotropic network of CNTs [248].
3 Strain

amplification: The local strain of an elastomer around a rigid reinforcement often considerably
exceeds that of the macroscopic strain, due to the matrix bearing the entire deformation. The effect in filled
elastomers can be that phenomenon such as strain-induced crystallization occur at a reduced global strain
level.

2.1 Introduction

13

In vulcanized NR, such permanent orientations are not easily possible, however with the
use of unvulcanized NR it is thought that some permanent deformation of the reinforcement
could be achieved due to the viscous effects of the entangled polymer chains; and the resulting behaviour of the material could provide further understanding of the material networks.
Viscoelasticity
The viscoelastic properties of rubber materials can be advantageous (e.g. dissipation in
car tyres) or disadvantageous (e.g. the creep in plastic shopping bags when full). Therefore it is important to understand the viscoelastic effects, particularly as they are further
complicated by the addition of reinforcements [142].
The viscoelasticity of filled elastomers is often investigated under stress-relaxing conditions. An investigation of the effects of fillers has shown that in crystallizing elastomers,
i.e. NR, the fillers act as nucleation sites for crystallization resulting in an increase of the
relaxation rate due to a relaxation of the amorphous phases [205]. A similar increase in
relaxation rate with the addition of fillers in non-crystallizing materials has also been investigated and could be attributed to the strain amplification effect [49, 141], as well as a
loss of cross-linked network and breakdown of the molecular weight of the polymer during
processing, due to the presence of the fillers [43]. In unvulcanised specimens however, the
presence of fillers slow down the relaxation rate due to the increased viscosity and additional
entanglements provided by a large amount of CB (up to 46 wt%) [43].
Closely related to these relaxation effects are the dynamic moduli. When a viscoelastic
material is stretched at a given strain rate, the response of the material will not be instantaneous; in fact the stress lags behind the strain and allows the elastic and viscous quantities to
be separated mathematically, which are known as the storage and loss moduli respectively.
The ratio of these quantities, referred too as the loss tangent (4) , is representative of the
damping ability of a material.
The addition of reinforcing fillers to rubber has the effect of simultaneously increasing both the storage modulus and the loss tangent, although by different mechanisms. The
storage modulus is representative of the stiffness of a material at a dynamic strain (i.e. nonstatic) and is typically increased by the inclusion of reinforcements because of the stress
transfer between the rigid fillers and bulk matrix [16]. However, the reinforcements also increase friction, due to polymer/filler interactions as well as particle interlayer sliding [148].
This, accompanied by measurement of the glass transition temperature [173], can provide
information on the interfacial interactions of the material.
4 The

tan :=

loss tangent, also referred to as tan , is the ratio of the loss modulus to the storage modulus i.e.

G ()

G ()

14

Experimental Mechanical Characterisation

Fatigue
Rubbers are typically required to be a high performing materials that can withstand high
strain and have the ability to dissipate large amounts of energy efficiently. This makes it an
ideal material for use in tyres, seals, vibration isolators, belts and hoses, structural bearings,
boat engine impellers, impact dampers, footwear, among other applications. However, this
also means that in its life-time the material is expected to withstand repeated static and timevarying strains, making the fatigue life of vital importance. This can be especially difficult
in elastomer materials when effects such as strain-induced crystallization become important.
Whilst the nucleation and growth of cracks is of primary importance when considering
the life of rubber materials in service [166], and in fact nano-reinforcing materials have
been shown to stunt the growth of cracks, the prohibition or delay of this nucleation is
also advantageous. Crystallization is a known mechanism in NR that helps to delay fatigue
failure, particularly in reinforced elastomers where the amount of crystallization at a given
strain level is amplified [12, 146, 216]. In these cases however, the loading conditions are
particularly important due to their relation to crystallization. Non-relaxing load conditions
favour the formation of large crystallites [216], especially if the minimum strain is above
the crystallite melting strain ratio, which help to halt the growth of cracks [146]. Even at
moderate strains ( 30%), reinforced elastomers are able to accelerate the formation of SIC
around the highly strained crack tip and reduce its effect [259].
In contrast, non-crystallizing rubbers are typically not as fatigue resistance [11] although
similar mechanisms can still enhance the property. For instance, strain-induced anisotropy
helps to split the cracks and spread their energy; an effect that anisotropic inclusions can aid
[188]. Therefore, the amplitude of the applied strain and the minimum strain are important
in the same way as they are for NR [1]. Reinforcements can also roughen the surface
of cracks, causing crack pinning and deflection [19, 188]. However, without appropriate
dispersion and distribution, fillers can cause agglomerations which reduce the fatigue life
by causing cracks to propagate from stress concentrations [128].
To mitigate this risk, the proper inclusion of reinforcements is necessary. By distributing
them homogeneously and with a good dispersion, a strain amplification can occur whereby
a smaller strain is necessary to reach an equivalent stress. This reduces the growth of stressconcentrations into cracks by reducing the maximum stress the material operates at. As this
effect is directly related to the fatigue performance of a material, it is possible to predict the
long-term performance of a material through the use of relatively short-term testing such as
stress-ratcheting [263]. By cycling the material between fixed stress values and observing
the evolution of strain, it is possible to predict the fatigue performance [48] and even gain

2.1 Introduction

15

theoretical insight into the physics of the material [49].


Elastomeric Networks Created by Anisotropic Reinforcements
The formation, and subsequent breakdown, of filler networks is also an important area of
research. Whilst there has been a lot of work on the advantages of equi-axed nanoparticles
[34, 107], one of the biggest advantages of elongated fillers such as fibres and platelet, is
their ability to orient and form effective networks within the material that can enhance, for
example, the mechanical and electrical properties.
The improvement of these networks is typically achieved by superior dispersion and distribution of the fillers [265], however a better network can be formed by careful processing,
as Ref.[196] demonstrated by the formation of a web-like network structure that enhanced
the mechanical, electrical and thermal properties when compared to an even dispersion of
the same filler.
In fact, the positioning of fillers, coupled with their morphology, can create a tortuous
path that retards the advancement of, for example, cracks [188] or improves permeability
[204]. This tortuous path can be further supplemented by the flat elongated morphology of
GE platelets [264] (See Fig. 2.5).

Fig. 2.4 The tortuous path caused by aligned anisotropic reinforcements in a matrix medium.
The blue arrow shows the path of least resistance for, e.g. water molecules travelling through
a permeable material.

2.1.2

The Benefits of Oriented and Aligned Reinforcements

The effects of anisotropy upon material properties are extensively dealt within industry and
research: often this is incidental anisotropy produced as a result of manufacturing processes,
such as induced anisotropy in mylar sheets produced by roll milling [110], or the anisotropic
behaviour of NR induced by large strain deformation [237].
The reinforcement of materials by the addition of filler particles is typically applied to
improve the material properties, and often the use of anisotropic fillers is employed because

16

Experimental Mechanical Characterisation

of their increased interface. However the addition of these fibres can produce anisotropy,
even when the material is initially isotropic, due to the rotation of the fibres at large strain
[206].
By careful consideration of the spatial and orientational distribution of these reinforcements it is possible to tailor the anisotropic properties to give exceptional properties that
are not attainable in isotropic materials. The anisotropy can be attained in 3-dimensions,
giving a large design space and allowing optimisation of the constituent materials within the
composite. However, anisotropic materials can be highly non-linear and so it is important to
understand the behaviour, the factors that affect it and be able to predict it. It is also possible
to further utilise this anisotropy for other material properties such as the electrical [264],
magnetic [131] and permeability [220] properties.
Transversely Aligned Reinforcements
The orientation of reinforcing materials is a technique used extensively in nature, for
example the fibril angles of spruce tree branches are oriented to optimise the structure to the
loading encountered by the tree and can even adapt to changing loading conditions during
its growth. Due to the small size, and subsequently low moment of inertia of young branch
stems, they are unable to withstand high wind loads; therefore the fibrils of the spruce
are initially oriented away from the loading direction where they give low stiffness and
high toughness. As the tree branch grows, it creates stiffer deposits of material on the top
tension side of the branch, by gradually orienting the fibril angle until it is parallel to the
loading direction [59].
The advantages of orienting reinforcements in polymer systems have been demonstrated
by many researchers [52, 53, 154]. This is especially true for nano-dimensioned materials,
which can bring significant property improvements with small amounts of reinforcing material. For example, a 48% (415MPa to 615MPa) increase in stiffness has been obtained
by randomly dispersing CNTs in an epoxy matrix. However by aligning the CNTs, a 102%
(843MPa) increase in the stiffness can be obtained parallel to the fibres with only a small decease in the perpendicular direction (575 MPa) [248]. Similar increases in strength are also
observed, and a 74% increase in strength is obtained when orienting the fibres compared to
a random dispersion.
Platelets have the advantage that they can reinforce in two principle directions because
of their morphology, and control of both axes can be obtained using magnetic fields. In
Ref.[154] longitudinal alignment of alumina platelets has resulted in a 124% increase in
the flexural stiffness compared to the pristine epoxy, with a moderate increase in stiffness
when the platelets are oriented transversely. However, whilst a 10% increase in strength is

2.1 Introduction

17

also obtained when the platelets are oriented longitudinally to the loading direction, transverse alignment induces stress concentrations, fails to stop the propagation of cracks and
ultimately results in the onset of brittle fracture (See Fig. 2.5).

Fig. 2.5 The crack path for aligned anisotropic fillers. The blue arrow indicates the crack
path. (a) parallel to the loading direction, (b) perpendicular to the loading direction.
At this stage it is worth discussing the difficulties in experimental testing of anisotropic
materials, which has led to very little reliable experimental data being available in the literature; particularly for more complicated modes such as bi-axial and shear deformations [24].
A transversally isotropic lamina can be characterised to some degree by tensile tests of 0 ,
45 and 90 configurations [115]; however it is important to consider the large strain behaviour and the behaviour of different deformation gradients [24, 46]. This can be difficult
due to the oblique angle of anisotropy which is known to cause shear forces and bending
moments in standard grips, resulting in stress concentrations [45]. When the anisotropy is
weak, it may be appropriate to neglect the effects, however when the anisotropy between
principal directions is an order of magnitude different (such as in Ref.[256]) it is necessary
to use, e.g. oblique test fixtures, that clamp at the angle of anisotropy [256]. At large strain
this is not possible due to the rotation of the principal axes, and so rotating test fixtures are
needed.
Anisotropy has had beneficial effects to a number of other properties, for example, the
hardness of the material has been found to be highest when the platelets were oriented
vertically to the surface, compared to a very low hardness when the platelets were oriented
parallel or randomly to the surface [154]. This arrangement of the outer layer in a material
to give protective properties can be observed in mollusc shells.
The outer layer of mollusc shells consists of perpendicularly oriented calcite prisms that

18

Experimental Mechanical Characterisation

increase the wear resistance [215]. The wear of material consists of the development of
shear stresses at the surface that remove particles from this area, however particles perpendicularly aligned to the surface are harder to remove. Beneath this layer, the mollusc shell
is designed so that any cracks that do initiate are met by a layer of parallel layered aragonite
platelets that create a tortuous path and hinder crack propagation [215].
Polyurethane reinforced with magnetite coated alumina platelets have shown a 77%
increase in the wear volume when oriented vertically out of the surface, compared to longitudinally oriented platelets [52]. This increase in wear resistance is an important quality of
many materials, particularly those expected to be in service for extended periods, however
vertical alignment of the platelets is also accompanied by a reduced in-plane stiffness. In
order to achieve an optimised material that combines the best qualities of both materials, it
is possible to produce a layered material.
Reinforcements Distributed in a Layered Arrangement
Layered materials make it possible to combine the properties of multiple constituents,
and utilise the constituents in optimal configurations throughout a structure to strengthen
it. For example, teeth consist of a 3D multi-layer structure; primarily consisting of a hard
enamel shell built to withstand mastication loads and a dentine core that is very tough and
helps to stop cracks propagating [61].
The use of mismatched expansion ratios in material layers has been demonstrated to
create actuating capabilities, with the mode and degree of actuation controlled by the orientation of the constituent layers. Erb et al. [53] oriented super-paramagnetically activated
aluminium oxide platelets in a hydrogel, with the orientation of the platelets controlling
the swelling expansion ratios in water (because the in-plane swelling ratio is reduced with
respect to the out-of-plane direction). Multiple layers can be manufactured by adding successive material on top of the cured layer, and then applying a different magnetic field; the
result is that bending/twisting effects can be produced as one layer swells more than the
other.
Three-dimensional Spatial Distribution of the Reinforcements
Conventional composites typically consist of layered configurations of fibre reinforced
resin. However a difficulty with layered materials is in the stress concentrations that arise at
tapered sections of a structure. Tapering can consist of ply-drops which contain resin rich
areas; the structural and mechanical gradient at this point causes a stress concentration and
can initiatDe failure of the material [245].
Stress concentrations arise in structures when there are geometric or mechanical dis-

2.1 Introduction

19

crepancies between adjoining interfaces. A famous example is the De Havilland Comet,


which developed cracks due to the formation of stress concentrations at the window corners resulting in the catastrophic failure of two of its aircraft [47]. The human body also
has to deal with many mechanical gradients, for example, the interface between ligaments
and bone includes a large discrepancy in the moduli of around 2 magnitudes; to overcome
this challenge the body uses mechanical and structural gradients that gradually smooth the
transition of stress between the two constituents [253].
Libanori et al. [155] have shown that by combining micro-dimensioned and nanodimensioned platelets, it is possible to reduce the formation of stress concentrations within
a material. This hierarchical strategy allows the material to be sufficiently ductile, for the
inclusion of large concentrations of reinforcements. Similarly, a mechanical gradient can
be formed by the sequential layering of material with increasingly large stiffness so that the
through-thickness stiffness changes by as much as five magnitudes, whilst still being able
to achieve reversible stretching of up to 300% strain [153].
An alternative to layered structures is the spatial positioning and orientation of the reinforcement. Magnetically responsive reinforcements can be oriented in a 3D magnetic field,
with the alignment of the reinforcement controlled by the direction of the field lines. For
example the use of low strength permanent magnets has been used to vary the orientation of
platelets along a material, creating local stiffening effects that are visible as a wavy pattern
upon the swelling of the material [52]. However, by using a non-homogeneous magnetic
field it is also possible to control the position of reinforcements. Erb et al. [52] demonstrate
the possibilities of controlled positioning in the vicinity of a structural defect (e.g. a hole);
the reinforcing platelets aggregate towards the hole in a radial pattern to strengthen it and
prevent a stress concentration from forming. The consideration of reinforcing structurally
weak areas is a requisite of creating efficient structures, however this is usually achieved by
only considering geometrical changes (i.e. additional material) rather than by considering
more efficiently utilising the material constituents.

20

Experimental Mechanical Characterisation

2.2

The anisotropic effects of nano-inclusions

2.2.1

Natural Rubber Reinforced by Graphene Oxide

The addition of fillers into elastomers can add further improvements to the mechanical
[203], viscoelastic[186] and magnetic properties [29] of the material, among others[136,
149]. The improvement of these properties depends on a number of factors that include, but
is not limited to: particle morphology, size, aspect ratio, dispersion and surface structure
[65, 196]. Therefore, due to their large specific surface area, excellent physical properties
and increasing availability, there is currently significant interest engaged in the exploitation
of nano-dimensioned fillers (such as nanoclays [49], graphene [164, 186, 262], and carbon
nanotubes (CNT) [14]).
Since the discovery of single layer graphene [182], there has been increasing interest in
elastomer-graphene nanocomposites to improve on existing material solutions. Graphene
is nano-dimensioned, with an exceptionally large surface area due to its large aspect ratio
[39]. It also has very high in-plane stiffness, out-of-plane flexibility, high electrical and
thermal conductivity [147] and can be functionalized to further improve or customise the
properties; and increase the surface interaction with different polymer matrices [78]. Any
improvements in the properties of a nanocomposite material are often very dependent upon
the physiochemical interaction of the graphene and matrix [139], and so functionalization is
very important. Graphene Oxide is often used as a precursor material in the production of
graphene [265]; however the many oxygenated reaction sites on its structure (See Fig. 2.6)
can provide reaction sites for further functionalization [78], and are expected to increase
both the physical and chemical reactions with a matrix material. The result is expected to
be an increase in the bulk properties of an elastomer.

Fig. 2.6 Schematic structure of (a) graphene and (b) graphene oxide. Adapted from [150].
Among the most commonly used industrial elastomer materials is natural rubber (NR),
which is usually extracted in the form of a latex, consisting of a colloidal suspension of

2.2 The anisotropic effects of nano-inclusions

21

latex particles, non-rubber impurities and water [69]. Natural rubber is characterised by
its ability to deform to large strain, damping performance and durability; qualities that are
improved by cross-linking of the polymer chains through the vulcanization process (5) . Nonvulcanised NR shares many of the characteristics of vulcanised NR, in that the long polymer
chains entangle and create enhancements in the residual forces (van der waals) [241] and
additional functional groups at the end of polymer chains are able to interact with non-rubber
components to form a natural pseudo network [241].
NR demonstrates a characteristic upturn in the stress at high strain (typically 100%) due
to SIC, caused by the alignment of polymer chains at high strain to form crystalline regions
[42] and involves structural change, an increase in density and a release of latent heat [241]
(See Section 2.1).
The addition of morphologically anisotropic fillers (e.g. graphene and CNTs) provides
significant mechanical advantages to NR (see for example [186], [164] and [206]). Their
large surface area and aspect ratio are thought to accelerate the formation of a stiffening
filler network, compared to spherical particles [206].
In this section, the effects of GO nano-platelets upon the mechanical and viscoelastic
properties of unvulcanised NR are investigated; both of which are affected by the stress
transfer and dissipative properties of the fillers. At large strain, new insights into the material
behaviour are seen; where a possible rotation of the nanoplatelets occurs in the direction of
the applied load.

2.2.2

Materials characterization and experimental techniques

Materials Characterization
Natural Rubber (NR, 60% high ammonia, Chana Latex Co., Ltd., Thailand) Latex was obtained from a private source and used as received and without further modification. Fouriertransform (FTIR) spectra of graphite, graphene oxide, natural rubber and NR-GO nanocomposites are shown in Fig. 2.7c. The FTIR spectra of natural rubber showed signature peaks
at 2960 cm1 , 2920/2855 cm1 , 1447/1376 cm1 , 1086 cm1 and 838 cm1 corresponding
to cis-isoprene units [81].
Graphene Oxide (GO) was prepared from graphite according to the Hummers-Offeman
method described in [189]. TEM and AFM analysis of the GO precursor were performed
by air drying the aqueous GO solution on carbon coated copper grids and freshly cleaved
5 Vulcanisation

is the process of cross-linking the polymer chains in natural rubber and other similar polymers. Sulphur is predominantly used as the cross-linking agent but others can be used, such as peroxides. The
form and amount of additives has a large effect upon the material properties and can be used to customise the
properties [42]

22

Experimental Mechanical Characterisation

Fig. 2.7 (a) AFM images of the GO platelets reveal the delaminated sheets to be 1-1.4 nm
in thickness, consistent with completely exfoliated single layer GO; (b) TEM image of GO
platelet reveals ultra-thin flat sheet between 500 nm and 1 lm in size; (c) Fourier-transform
spectra of GO, graphite, natural rubber and NR-GO nanocomposites containing 1.00 wt%
GO; and (d) SEM images of fractured 1.00 wt% NR-GO specimens reveal homogeneous
dispersions of filler particles.

mica sheets respectively. Tapping mode AFM study showed the presence of delaminated
sheets that were 1-1.4 nm in thickness (Fig. 2.7a), which is consistent with the formation of
completely exfoliated single layer GO platelets. TEM images (Fig. 2.7b) indicate the size of
the largest platelets to be 500 nm to 1 m in length. FTIR analysis (Fig. 2.7c) of graphite and
GO exhibited features at 3446 cm1 , 2925/2852 cm1 and 1627 cm1 corresponding to OH, C-H and C=C stretching vibrations from absorbed water, alkyl groups at the edges and the
graphitic sp2 hybridised network. Additional peaks were present for the GO sample, with
peaks at 3378 cm1 , 1718 cm1 and 1225 cm1 indicating hydroxyl, carbonyl/carboxylic
acid and epoxide groups respectively [190].
The NR-GO nanocomposites were prepared by latex-mixing of the NR latex with an
aqueous dispersion of GO platelets. This involved the addition of 3 g of NR latex to a petri
dish, followed by 3ml of aqueous GO platelets (wherein the aqueous concentration of GO
is controlled to give the required wt%). The solution is then agitated on a vibrating plate
to achieve a homogeneous dispersion, and finally left to dry for 48 hours. The resulting

2.2 The anisotropic effects of nano-inclusions

23

nanocomposite showed no FTIR vibration peaks characteristic of GO, due to the strong
presence of symmetric and asymmetric stretching frequencies of isoprene units. However,
SEM images (Fig. 2.7d) indicated a homogeneous dispersion present throughout the fracture surface. Characterisation of the nanocomposites by Differential Scanning Calorimetry
(DSC) showed only a small increase in the glass transition temperature (Tg) of 0.7C
between concentrations of 0 wt% and 1.0wt%, indicating that the presence of GO lowers
the quantity of polymer that is fluidified at the Tg; a result previously reported in similar
nanocomposite systems [164, 186, 192, 206].
Nanocomposite concentrations of 0, 0.25, 0.50, 0.75, 1.0 wt% were produced and tested.
Testing Techniques
All tests were performed at room temperature, with rectangular specimens of dimensions
5.0 mm width, 1.0 mm thickness and approximately 17.5 mm length. The nominal stress is
determined as the ratio of the measured force to the original specimen cross-sectional area,
whilst the strain is determined as the ratio of the extension to the original distance between
the clamps.

Fig. 2.8 (a) Specimens of NR-GO used in experimental testing. Pure NR on the left, increasing in GO conc. up to 1.00 wt% on the right. (b) The petri dish arrangement from which
specimens are cut, showing a pure NR material.
Mechanical testing was performed with an Instron testing machine, unless otherwise
stated, on two batches of nanocomposite specimens.
The first set of specimens (batch one) are tested 6 weeks after synthesis to investigate
the effects of GO concentration on the stiffness, strength and strain-to-failure by applying
displacement control (100 mm min1 ) until rupture.
A second series of specimens from batch one are tested 2 weeks after synthesis, to
investigate the strain rate dependency of the nanocomposites. This involved cyclically load-

24

Experimental Mechanical Characterisation

ing/unloading the specimens to 30% strain; incrementally increasing the displacement rate
so that rates of 10 mm min1 , 50 mm min1 , 100 mm min1 , 150 mm min1 and 200 mm
min1 were applied.
A third set of specimens from batch one were used to investigate the relaxation and
fatigue properties of the nanocomposite. This involved stretching one set of specimens
to 50% strain, then holding them for a period of 10 minutes to observe the relaxation of
the applied force. An initial displacement rate of 100 mm min1 was applied. Another
set of specimens were stress-ratcheted between 0.3 MPa and 0.15 MPa for 200 cycles at a
displacement rate of 100mm min1 to observe the evolution of strain values between cycles.
Batch two was used for two separate series of experiments, tested at 2 and 6 weeks
after synthesis respectively, in order to examine the NR-GO networks and the effects of the
naturally occurring pseudo network [4]. The testing consisted of applying a stress of 1.4
MPa via displacement control (100 mm min1 ), before releasing the stress for 24 hours and
then again re-testing to 1.4 MPa. The stress value is chosen based on preliminary testing;
assuming the need to maximise disruption to the polymer networks, whilst ensuring the
specimen did not rupture.
Those specimens tested to high strain (> 400%) showed, upon retraction of the loading
force, a permanent deformation: when tested to 1.4 MPa, the permanent deformation for
pure NR was 30%; which decreased monotonically with the addition of GO to 10% at
a concentration of 1.0 wt%. The application of lower strains (i.e. 50%) resulted in the
complete recovery of the initial length with no residual deformation.

2.2.3

Results and discussion

Constant strain rate testing


The stress-strain results of uniaxial tensile testing, at constant strain rate, on specimens of
natural rubber reinforced with graphene oxide can be seen in Fig. 2.9. It displays the S
shape behaviour typical of rubber behaviour [241], displaying the upturn in stress at high
strain associated with SIC.
At moderate strain levels it is possible to fit the non-linear behaviour with the MooneyRivlin model [175, 208], often used for filled elastomers [13, 163, 226]. This allows the
small strain properties to be more easily determined. The Mooney-Rivlin relationship is
given by,
N11 = 2(C10 +C01 / )( 2 )

(2.1)

where N11 is the nominal stress parallel to the loading direction (i.e. force over reference

25

2.2 The anisotropic effects of nano-inclusions

2.5
2

0.5
0.00
0.25
0.50
0.75
1.00

wt%
wt%
wt%
wt%
wt%

1.5
1
0.5
0
0

(a)
100

200

300

400

500

600

Nominal Stress (MPa)

Nominal Stress (MPa)

0.4
0.3

0.00 wt%
0.25 wt%
0.50 wt%
0.75 wt%
1.00 wt%
Model

0.2
0.1
0
0

(b)
5

10

15

20

25

30

Strain (%)

Strain (%)

Fig. 2.9 Uniaxial tensile test results of NR-GO specimens (a) The stress-strain curve up to
failure. (b) The small-strain fitting of the Mooney-Rivlin model (Eq. 2.1) up to 30% strain.
area), is the elongation ratio (i.e. deformed length over original length) and the parameters
C10 and C01 are phenomenologically determined material constants determined by comparison to the experimental data. The Youngs modulus of the nanocomposites can then be
related to the model parameters by E = ( / ) 1 = 6(c10 + c01 ).
The fitting at low strain can be seen in Fig. 2.9b, showing the accuracy of the fitting up
to 30% strain.
Table 2.1 NR-GO nanocomposite tensile properties at concentrations between 0.0 wt% and
1.0 wt%.
GO (wt%)
0.00
0.25
0.50
0.75
1.00

Strain-to-failure (%)
598 (39)
546 (33)
516 (03)
501 (24)
473 (23)

Strength (MPa)
2.34 (0.17)
2.30 (0.18)
2.87 (0.18)
3.12 (0.36)
3.13 (0.24)

Youngs Modulus (MPa)


1.58 (0.10)
1.85 (0.28)
2.09 (0.24)
2.18 (0.37)
2.33 (0.15)

The results in table 2.1 show that with the addition of GO: the stiffness increases, the
strength increases and the strain-to-failure decreases.
The significant effects seem to be at low strain (indicated by the increase in the Youngs
Modulus) which have a knock-on effect to the large strain results; at large strain the nanocomposites behave very similar, although there is a slight increase in the tangent modulus
(Fig. 2.9a). Two reinforcing mechanisms are thought to act on the small strain response:
reinforcement by the filler [13] and the reinforcement from the polymer network of intercalating polymer chains; which itself consists of end-linking pseudo-networks and entanglements [186].

26

Experimental Mechanical Characterisation

At large strain, for increasing concentrations of GO, Fig. 2.9a shows a decrease of the
strain at which the upturn in stress occurs. This upturn in stress has been associated with
SIC in filled and unfilled natural rubber [186]; and the decrease of the strain level at which
it occurs is due to the presence of the filler and trapped rubber increasing the crystallisation
rate for a given level of macroscopic strain [196].
To further understand these effects, and additionally the effects of the naturally occurring pseudo network, supplementary tests were performed at 2 and 6 weeks after synthesis.
Specifically this will allow the effects of the cross-linking density to be investigated, which
is known to increase with time, in relation to the concentration of GO. It will also allow the
effects of large strain deformation upon the NR-GO networks to be investigated, and their
resulting effect upon the SIC, by subjecting each specimen to a non-destructive stress level
of 1.4 MPa before resting it for 24 hours and again testing to 1.4 MPa.
The results of this testing, that can be seen in Fig. 2.10 and Table 2.2, show that for
all specimens on the second cycle; both the Youngs Modulus and the strain at which the
upturn in stress occurs (calculated from a Mooney plot, see Fig. 2.10) are decreased. The
Mooney plot allows the upturn in stress (defined as the point at which the tangent of the
modulus begins to increase), and considers the reduced stress as a function of the inverse
.
stretch, where the reduced stress is defined by force
1
2

Table 2.2 Youngs modulus, E, calculated from based on the fitting of the Mooney Rivlin
formulation (Eq. 2.1) to the stress-strain response up to 30% strain, alongside an estimate
of the strain value at which the upturn in stress occurs (associated with the onset of SIC
[37]), which is estimated from a Mooney plot. The testing consisted of the first uniaxially
stretching the specimen up to 1.4MPa and then resting it, followed by again testing to 1.4
MPa after 24 hours; resulting in two sets of results.
GO (wt%)
0.00
0.25
0.50
0.75
1.00

First cycle
E (MPa)
upturn (%)
1.03 (0.03) 211 (4.9)
1.44 (0.04) 177 (10.6)
1.69 (0.01) 164 (10.1)
1.78 (0.09) 149 (15.6)
1.89 (0.13) 158 (6.9)

Second cycle
E (MPa)
upturn (%)
0.79 (0.09) 141 (20.9)
0.99 (0.10) 133 (5.0)
1.00 (0.11) 107 (21.4)
1.09 (0.10) 108 (13.3)
1.08 (0.03) 116 (12.0)

It is well understood that both entanglements and the pseudo network play an important
role in the Youngs modulus of unvulcanised NR [4]; therefore the fact that the Youngs
modulus decreases on the second cycle suggests there is a breakdown in many of the polymer networks and a reduction in the number of entanglements, that is not recovered in 24
hours. The percentage reduction in the Youngs modulus also increases with higher concentrations of GO due to the presence of more filler-matrix interactions that can be disrupted.

27

2.2 The anisotropic effects of nano-inclusions

0.6
0.4
0.2

0.6

1st cycle
2nd cycle
1.25

2.5

3.75

Stretch,

(d)

0.4
0.3
1st cycle
2nd cycle

0.2
0.25

0.8
0.6
0.4
0.2

0.6

0.5

0.5

0.75

Inverse Stretch, 1/

1st cycle
2nd cycle

0
0

1.25

2.5

3.75

Stretch,

1
0.8
0.6
0.4
0.2

0.6

0.5
0.4
0.3

1.2

1st cycle
2nd cycle
0.25

0.5

0.75

Inverse Stretch, 1/

1st cycle
2nd cycle

0
0

(e)

0.2
1

Nominal Stress (MPa)

0.8

(c)

1.2

Reduced Stress (MPa)

Nominal Stress (MPa)

0
0

Reduced Stress (MPa)

(b)

Reduced Stress (MPa)

Nominal Stress (MPa)

(a)
1.2

1.25

2.5

(f)

0.5
0.4
0.3
1st cycle
2nd cycle

0.2
1

3.75

Stretch,

0.25

0.5

0.75

Inverse Stretch, 1/

Fig. 2.10 Uniaxial tensile testing results of cycling specimens up to 1.4 MPa, in between
24 hour rests. (a)&(d) 0.00 wt%, (b)&(e) 0.25 wt%, (c)&(f) 0.50 wt%. (a)-(c) Stress-strain
curves. (d)-(f) Mooney plots.
The decreased strain level at which the upturn in stress occurs on the second cycle is
a peculiarity that is not adequately explained by residual crystallisation. It is anticipated
that a small amount of reconfiguration will occur in the sample, especially as it is unvulcanised natural rubber (i.e. no permeant cross-linking exists) and so the presence of oriented
regions upon full retraction of the load in the first cycle is possible[4, 240], however the
amplification of this effect with increased GO concentration, i.e. a decrease in the strain of
the upturn with increased GO wt% (See Fig. 2.10), suggests the possibility of some permanent orientation of the GO platelets towards the direction of the applied load during the first
cycle.
The rotation of platelet-like reinforcements has previously been observed (see e.g. [186,
206]), and indeed the permanent orientation of CNTs after large strain cyclic loading has

28

Experimental Mechanical Characterisation

been observed [229]. This would potentially increase the effectiveness of the reinforcement
(due to GO in-plane stiffness being much higher than the bending stiffness [229]), and have
significant effects upon the large strain performance, as we observe here.
The permanently cross-linked network in Vulcanized Rubber prevents polymer chain
movement, and would therefore return the platelet to its original orientation. However, nonvulcanised NR does not have this permanent network and so small configurational changes
could occur internally, even if there are no obvious visible signs. The result would be
increased reinforcement in the direction of the applied load, which would account for the
further effects that are observed on the second load cycle with increasing GO concentration.

Variable strain rate testing


The specimens were subjected to uniaxial tension loading at varying strain rates. The results
at 50 mm min1 for three different concentrations (0.00, 0.25 and 0.50 wt%) can be seen in
Fig. 2.11, with the strain and stress evolutions in time shown in Fig. 2.11a and Fig. 2.11b
respectively.
Two major effects of increasing the concentration of filler can be observed from Fig. 2.11c:
an increase in the Youngs modulus and an increase in the area within the stress-strain
curves; the latter hysteresis indicating an increase in the dissipated mechanical energy. The
increase in the Youngs modulus can be attributed to the action of entanglements and interactions between the rigid filler and polymer chains; hence, as the GO concentration increases,
the short strain response should closely resemble the effect of an increasingly cross-linked
network [43]. At high strain the platelets could slide across the matrix, and the large interface of the GO platelets means increased polymer/filler friction, resulting in the increasing
hysteresis as the GO concentration increases.
To discern between these two effects, i.e. the increases in Youngs modulus and hysteresis, the value of the dissipation, D(6) , and the ratio between the dissipated and stored
energies, D/S( 7) , are calculated by quantifying the area within the stress-strain curve (see
Fig. 2.12). The results, assuming a loading ramp duration of tR /2 and an entire cycle of tR ,
6A

non-viscous material would have loading and un-loading paths that would coincide, indicating that
al the mechanical energy provided to deform the specimen was completely recovered and so the dissipated
energy, D, would be zero.
7 The ratio between the dissipated and stored energies, D/S, is related to tan ; obtained from dynamic
mechanical analysis (DMA).

2.2 The anisotropic effects of nano-inclusions

29

Fig. 2.11 Loading\unloading cyclic test at 50 mm min1 (strain rate of 5.55 102 ). (a)
Strain versus time; (b) stress versus time for 0.25 wt.% against the constitutive model in
Eq. 2.1; (c) stress versus strain at 50 mm\min. The model for 0.00 wt.%, 0.25 wt.% and
0.50 wt.% are represented by the solid, dashed and dash-dot lines respectively.
can be estimated by

Z tR

D=

(t) (t)dt;

0
Z tR /2

S=

(t) (t)dt

(2.2.a)
(2.2.b)

The dissipative contribution of the NR specimen (0.00 wt%) has been normalised, so as
to highlight the contribution of the filler. For all concentrations of GO, the dissipated energy
increases monotonically with the applied displacement rate of the loading ramp: from 1.5
at 50 mm min1 to 2.3 at 200 mm min1 for the 0.25 wt% specimen and from 2.2 at 50 mm
min1 to 4.1 at 200 mm min1 for the highest GO concentration (1.00 wt%). However, it
should be noted that the highest dissipation is observed in the 0.75 wt% specimen and does
not increase for the 1.00 wt% specimen; this saturation effect has also been observed for the
Youngs modulus (see Table. 2.1).
Contrary to the dissipative properties, the ratio D/S monotonically decreases with loading rate but increases with filler content. It is well known that low strain rates favour the
dissipative aspects of material behaviour, whilst higher strain rates favour the elastic properties of materials [142]; however its relationship with filler content is less easy to predict
due to its dependence on a number of factors such as the interfacial strength between the

30

Experimental Mechanical Characterisation

polymer and filler.


Fig. 2.12 can give an indication of the different effects of the platelets upon the high
and low strain rate responses: at low strain rate the ratio, D/S, increases with increased GO
concentration; indicating the predominant effect of the platelets upon the dissipative energy
due to friction between the polymer chains and GO. However, as the strain rate is increased,
the elastic energy increases faster at a given concentration of filler due to the difficultly in
disentangling the polymer chains. The notable feature is that even at higher strain rates, the
GO platelets increase the ratio D/S; indicating that the platelets have a predominant effect
upon the dissipative properties of the material.

D/D

NR

(a)

Normalised Dissipated Energy


5
4
3

0.25 wt%
0.50 wt%
0.75 wt%
1.00 wt%

2
1

50

(b)
0.2
D/S

0.15
0.1

100
150
Velocity (mm min1)
Dissipated Energy / Stored Energy

200

0.00 wt%
0.25 wt%
0.50 wt%
0.75 wt%
1.00 wt%

0.05
50

100
150
Velocity (mm min1)

200

Fig. 2.12 (a) Normalised dissipated energy D during cyclic tests estimated through Eq. 2.2.b
normalised with respect to dissipation of the natural rubber sample (i.e. 0.00 wt.%). (b)
Ratio between the dissipated and stored energies during the cyclic test (estimated through
Eq. 2.2.b).
The effect of the GO platelet fillers can be further understood by observing the loading
ramps at different strain rates (see Fig. 2.13). The results show that a higher stress is reached
at 30% strain as the concentration of filler is increased, and independently as the strain rate
is increased; both phenomenon well understood. However the sensitivity to the strain rate is
also effected by the filler concentration, which can be indicated by the maximum stress at the
end of each cycle (Fig. 2.13): for NR the stress increases from 0.204 MPa at 50 mm min1
to 0.225 MPa at 200 mm min1 , a 10.2% increase, for the 0.50 wt% specimen the stress
increases from 0.332 MPa at 50 mm min1 to 0.370 MPa at 200 mm min1 . This increased
sensitivity of the strain rate with the addition of filler has been observed in other elastomer
systems [203], and is caused by the entanglements between polymer chain networks; an

31

2.2 The anisotropic effects of nano-inclusions

effect that GO and immobilised rubber may contribute to by creating effective cross-links.
0.4

Nominal Stress (MPa)

0.35

50 mm/min
100 mm/min
200 mm/min

0.3
0.25
0.2
0.15
0.1
0.05
0
0

10

15

20

25

30

Strain (%)

Fig. 2.13 Comparisons of the stress-strain plots at different cross-head speeds. Tested at 50,
100 and 200 mm min-1 for specimens v = 0.00 wt.%(continuous line), 0.25 wt.% (dashed
line) and 0.50 wt.% (dash-dotted line).

Relaxation
It has been seen that the addition of GO nanoplatelets to NR increases the stiffness of the
nanocomposite by providing a stiff network for the transfer of stress; a well known phenomenon in reinforced materials [142, 205]. However, prolonged loading of the composite
causes the network of entanglements and end-links to slowly breakdown over time [241];
resulting in the phenomena known as creep and relaxation. Here we report the results of
relaxation testing (see Fig. 2.14a), which involves loading the specimen to 50% strain at a
loading rate of 100 mm min1 , and holding it for 10 minutes.
The results show that at short times, the increased stiffness of the material, provided by
the presence of the stiff GO reinforcements, increases the maximum stress that is reached at
50% strain. At longer times, as the stress is unloaded, the rate of relaxation increases with
GO concentration; a result that has been observed in unfilled [239] and filled elastomers
[205], and even elastomers at much higher concentrations [141].
When the applied initial strain is large (i.e. >200%), any change in the relaxation times
can be associated to the formation of crystallites; which are themselves amplified by the
presence of fillers and immobilised rubber [239]. However, when the applied deformation
is lower (as in this case), the rate of crystallisation is known to be too low to justify this
increase in the relaxation rate. In Ref.[43] a similar increase was observed in swollen carbon
black filled synthetic rubber, and is thought to be due to a breakdown in the molecular

Experimental Mechanical Characterisation

0.5
0.4

0.00wt.%
0.25wt.%
0.50wt.%
0.75wt.%
1.00wt.%

(a)

0.3
0.2
0.1
0 1
10

10

10

Time (s)

10

Normalised stress, /max

Nominal stress (MPa)

32

(b)

0.9
0.8
0.7
0.6

0.00wt.%
0.25wt.%
0.50wt.%
0.75wt.%
1.00wt.%
1

10

10

Time (s)

Fig. 2.14 (a) Stress-time plot of the relaxation tests for the different GO concentrations. The
specimens were loaded at 100 mm min1 up to 50% strain, then held fixed for 600 seconds.
(b) The stress-time plot, without the initial loading ramp, normalised against the maximum
stress.
weight of the polymer during processing, produced by the presence of the filler. However
the procedure used in this study, i.e. "latex-mixing", is unlikely to produce any significant
breakdown of the polymer chains.
The large surface area of the GO nanoplatelets is likely to have two effects relevant to
the relaxation rate: firstly to reduce the number of entanglements within the rubber as the
chains lie flat against the platelet, and secondly to increase the maximum stress; of which
both effects would increase the relaxation rate with the addition of GO. To discern between
these two effects, the normalised unloading curves are shown in Fig. 2.14b. The results
indicate that, after normalising the stress, the relaxation rate still increases with increasing
GO content; suggesting that the polymer chains are less entangled, with more freedom to
reconfigure.
Fatigue
Fatigue in materials is a result of repeated cyclic loading; propagating damage to weaken
the material and often causing failure to occur at a value much lower than the strength of the
material. However, the addition of reinforcements has been shown both to increase these
effects by creating stress concentrations [128] and to decrease these effects by strengthening
polymer networks [49], especially in nano-dimensioned materials [262].
Fatigue is considered a long-term material behaviour, however it has been shown that
relatively short-term stress-ratcheting between fixed values of stress can be used to give
an indication of the fatigue resistance of a material [49].

33

2.2 The anisotropic effects of nano-inclusions

Strain (%)

60
40
20

(a)
0
0

10
20
Time (s)

Stress (MPa)

The testing involved cycling the specimens between 0.3 MPa and 0.15 MPa for 200 cycles at a displacement rate of 100 mm min1 ; a rapid evolution in the strain values between
cycles gives an indication of fatigue failure. Fig. 2.15 indicates the significant difference in
the fatigue resistance of the material with the addition of 1.00 wt% of GO; this is likely to
be due to the increased stiffness of the material because of the stiff network imparted on the
material by the presence of GO, meaning that the specimen is cycled between lower values
of strain (see Fig. 2.15a & 2.15a). The result is that the polymer chains of the NR within the
material are lower, meaning less breakdown of the network and a higher fatigue resistance.

(c)
0
0

10
20
Time (s)

30

Stress (MPa)

Strain (%)

30

10

0.2
0.1

(b)
0
0

30

20

0.3

50
100
Strain (%)

0.3
0.2
0.1

(d)
0
0

50
100
Strain (%)

Fig. 2.15 Stress-ratetching of NR-GO specimens. (a) The first 30 seconds of the NR specimen strain cycle. (b) NR stress-strain cycle up to failure. (c) The first 30 seconds of the
NR-GO specimen (1.00.wt%) strain cycle. (d) The NRGO (1.00wt.%) stress-strain cycle up
to failure.
By calculating the maximum strain of each cycle, it is possible to observe these effects
based on the concentration of GO; this is shown in Fig. 2.16. An increase in the fatigue
resistance is shown as the GO concentration increases up to 0.75wt%; the fatigue resistance
at 1.00wt% is slightly decreased, likely due to a similar saturation effect to that observed
for the Youngs Modulus. Similar reductions in the reinforcement effectiveness have been
observed in [49] for larger concentrations of nanoclay blends. There is a significant impact
due to the inclusion of even 0.25wt% GO compared to the pure NR specimen; showing the
increased fatigue resistance provided by a small amount of nano fillers due to their large
surface area and resulting strong internal network of polymer chains and fillers.

34

Experimental Mechanical Characterisation


200

Strain (%)

150

0.00
0.25
0.50
0.75
1.00

wt%
wt%
wt%
wt%
wt%

100

50

0
0

50

100

150

200

No. of cycles
Fig. 2.16 The evolution of the maximum strain reached at the peak of each stress cycle
during stress-ratcheting between 0.15 MPa and 0.30 MPa of specimens of NR-GO with
concentrations of GO between 0 wt% and 1.00 wt%.

2.2.4

Concluding Remarks

Graphene oxide reinforced natural rubber elastomers have been produced by latex-mixing
in a simple process that requires no additional heat or chemical input. Characterisation of
the specimens confirms that homogeneous specimens are produced, containing a regular
size distribution of oxidised graphene of single layer thickness.
Observations show that the introduction of the nano reinforcements increases the stiffness and strength of the material, indicating an increased stress transfer through the material.
The decrease of the strain-to-failure with increasing concentration indicates the stiffness increase is more significant than the strength. The increase in strength suggests that the GO is
well dispersed, as confirmed by the mechanical tests carried out on specimens cut along different directions of the same batch. The significant increase in the Youngs Modulus, even
for low filler content (e.g., +55% @ 0.75 wt%) illustrates the reinforcing effect, and suggests the presence of rubber trapped within the filler network. This is in part confirmed by
Tg measurements that showed only a marginal increase with increasing GO concentration.
The effect of the strain rate in experimental testing is known to be a relevant factor, and
is shown to be particularly relevant for nanocomposites, not only in increasing the Youngs
Modulus but also in the dissipative properties of the material. The ratio between dissipated
and stored energies has been calculated in cyclic tests and is shown to increase as the GO
concentration increases; presumably due to the increased friction caused by the presence of
high aspect ratio GO platelets. However, as the strain rate increases the ratio between the

2.2 The anisotropic effects of nano-inclusions

35

dissipated and stored energy decreases for a given GO concentration; due to the reduced
time given for the mechanism to take place.
Relaxation tests also show that the stress reduces more quickly with the addition of GO
because of strain amplification effects and a reduction in the number of entanglements. The
strain amplification effects also help to improve the short cycle fatigue testing behaviour as
the GO concentration is increased.
Interestingly, orientation of the GO platelets is thought to take place at high strain, which
subsequent testing on the material suggests is permanent to some degree. This is evidenced
by the reduction of the strain level at which the upturn in the stress-strain curves occurs when
the same specimen is subsequently retested after 24 hours rest; and effect that increases at
higher GO concentrations.
This effect would require further investigation which could be carried out by Small Angle Neutron Scattering on pristine and stretched samples, however the underlying behaviour
suggests the interesting properties of aligned anisotropic reinforcements in an elastomer that
is deformed to large strain.

36

Experimental Mechanical Characterisation

2.3
2.3.1

The anisotropic effects of micro-inclusions


PDMS reinforced by Carbon Fibres

The properties of composite materials can be improved by orientation of the reinforcements,


giving better utilisation of the constituent materials by strengthening in positions where they
can provide critical reinforcement. In conventional composites these challenges are often
met by careful orientation of the composite layers, for example anisotropy can help in aeroelastic applications [122, 140]. In elastomer materials the consequences of this orientation
effect, because of the large strain potential, have effects upon both the large and small strain
response. At large strain additional non-linear effects have been observed (Section 2.2 and
Ref.[226]) in discontinuous filler reinforced elastomers. Due to the potential complexities of
designs possible with 3D spatial reinforcement, it is important to understand this behaviour,
its effects upon the material behaviour and to understand the mechanisms underpinning its
behaviour, plus be able to model the behaviour.
The challenges in controlled orientation of materials have been considered using a number of different manufacturing methods, for example, shear forces [223], ultrasonic waves
[219], electrical currents [167] and magnetic alignment [31, 52, 131, 134, 197]. Some of
the advantages of magnetic alignment include:
The magnetic field provides a contactless volume force that orients the magnetically
anisotropic materials along field lines into 3D spatial configurations [52]
Permanent magnets and electromagnets are readily available and produce strong enough
fields to orient the reinforcements. In fact, surprisingly weak field strengths can orient
reinforcements when the size and shape is carefully considered [52]
Compared to electric fields, magnetic fields do not produce currents and are not sensitive to surface charge and pH [207].
There are found to be many factors affecting the orientation of anisotropic fillers in
a viscous solution. Firstly it is important to consider the magnetic field, in particular its
strength [134], field direction [52] and homogeneity [225]. Secondly, it is important that the
reinforcement morphology and magnetic functionalization compliment the magnetic field
[52]. By controlling these factors, it is possible to predict the distribution and configuration
of the reinforcing fibres within the host matrix [41], and hence provide spatial reinforcement
that can have delicate consequences upon the mechanical properties.
In this section, we take inspiration from some of natures solutions and apply these concepts to the large strain behaviour of elastomer materials reinforced with discontinuous

2.3 The anisotropic effects of micro-inclusions

37

fibres. The focus of this investigation is understanding the behaviour of the material when
the short fibres are oriented in a transversely isotropic configuration. It is expected that the
orientation of the discontinuous fibres will have consequences upon both the small strain behaviour, as has been shown in such a material previously, however the rotation of the fibres
at large strain is also expected to contribute to additional non-linear effects. An understanding of these effects, and the many other factors that control the mechanical behaviour of
such a material is critical to fully utilising its potential as a novel and innovate solution to
elastomeric reinforcement.
Motivated by the potential benefits of such a material, a transversely isotropic constitutive model is compared to the experimental data in order to gain an understanding of the
behaviour. Its performance at understanding the large strain behaviour is considered.

2.3.2

Material Preparation and Experimental Techniques

Material Preparation
PDMS (Sylgard 184) elastomer is used without further modification and reinforced with
nickel coated carbon fibres (Ni/C 40/60 wt.% chopped to 0.25 mm, diameter 4.8m) purchased from Marktek Inc. The PDMS fibre reinforced elastomers are made by direct mixing
of the two compounds under the influence of a magnetic field (in a process described later in
section 4.2); which includes strategies to achieve a homogeneous dispersion of oriented fibres with few agglomerations at a concentration of 6 wt% NiC fibres. This is later confirmed
by micrographs of the fractured surfaces (Fig. 2.17b).

Fig. 2.17 (a) The neodymium magnetic setup for aligning magnetic fibres in viscous PDMS
solution, and (d) The dispersion of fibres through the thickness of the specimen, observed
by SEM on a fractured specimen.
Once cured, the specimens are cut into dumbbell specimen shapes ready for mechanical

38

Experimental Mechanical Characterisation

Fig. 2.18 The dumbbell test specimen according to ASTM D1708, and an inset showing the
alignment of the fibres within a small central sample of the specimen (Highlighted fibres
shown for illustrative purposes only).
testing according to ASTM D1708 (see Fig 2.18). The resulting specimens have distributions in the fibre lengths and orientations that can be seen in Fig. 2.19; this is expected to
influence the mechanical properties [66], hence they are documented in order to make comparisons between the experimental results and theoretical/numerical methods, it also allows
the repeatability of the results to be assessed.
To characterise the fibre angle distribution and to aid in comparisons, the following
distribution function is applied:
ebcos(2 )


( ) = R
/2
bcos(2 ) d
e
/2

(2.3)

where b is a coefficient to be determined. This function is found to better describe the experimental results than expressions considered previously [63] and will be used later (Chapter 3.4) to understand the effects of fibre angle distribution the numerical environment. The
fitting of this function for a sample is shown in Fig. 2.19a.
Meanwhile, a normal distribution function is found to describe the length distribution of
the fibres better than alternative expressions presented in the literature to [63].
1
(L d)2
exp(
)
h(L) =
(2c2 )
( 2 c)

(2.4)

where c and d are coefficients to determine, and the fitting is shown in Fig. 2.19b.
The interfacial properties between the fibre and matrix are also of importance to the

39

Probability Density

1.5

Distribution
Distribution Function

0.5

(a)
0
90

45

45

Fibre Angle, ( )

90

Probability Density

2.3 The anisotropic effects of micro-inclusions

Distribution
Distribution Function
6

(b)
0
0

0.1

0.2

0.3

0.4

0.5

Fibre Length, L (mm)

Fig. 2.19 The fibre distributions from a specimen aligned between the four Neodymium N52
magnets. (a) The fibre orientation distribution, fitted to Eq. 2.3 (b=4.3), (b) The fibre length
distribution, fitted to Eq. 2.4 (c=23.0, d=1.59). The averaged values of 6 specimens are b=
4.5 0.57, c=25.7 4.9 and d=1.61 0.15.
overall material properties [262]. It is generally assumed in numerical and constitutive
models that the interface is ideal, i.e. that there is perfect stress transfer between the fibre and
the matrix. The result is that the composite is optimally strengthened by the stiff reinforcing
fibres, however in reality this is rarely the case and efforts are undertaken to improve the
interfacial properties [260]. Due to the lack of functional reactive groups on either the
PDMS or nickel-coated carbon fibres the interface is unlikely to be optimal, illustrated by
the clean fibre-pull out observed on fractured specimen surfaces (Fig. 2.20a); although some
trapped rubber is observed in-between fibre bundles that have agglomerated (Fig. 2.20b).

Fig. 2.20 Interface of the fractured PDMS specimens. (a) A poor interface showing fibres
without matrix attached [96]. (b) Trapped rubber between bundles of fibres.

40

Experimental Mechanical Characterisation

Experimental Techniques
All mechanical testing was performed at room temperature, according to ASTM D1708
on specimens of approximate thickness 0.5 mm. The nominal stress is determined as the
ratio of the measured force to the original specimen cross-sectional area, whilst the strain
is the ratio of the deformation between two points in the gauge length, compared to the
original distance. This is calculated by digital image correlation (DIC), by spray painting
the specimen with a very light speckle pattern and choosing distinct points either side of the
mid-point of the specimen at an approximate distance from each other of 10mm. Due to
the very small size of the specimens, this distance is kept consistent to ensure the resolution
does not affect the results, or that the end-tabs do not interfere with the strain field in this
region. Testing is performed on an Instron testing machine, applying displacement control
(10 mm/min) until failure to 6 specimens for each angle tested (i.e. 0 , 15 , 30 , 45 , 60 ,
75 , 90 ).
Microscope images of each specimen are taken prior to mechanical testing using a Carl
Zeiss Jenavert optical microscope; the post-processing of these images involved analysing a
60 mm x 40 mm section of each petri dish. Image software is used to create an image layer
onto which the fibres are drawn; this layer is then imported to the MATLAB (ver.2012b)
Image Processing Toolbox and information on the fibre position, orientation and size can
be determined. Fibres that interact/overlap are added to a separate image layer so that they
can be treated as two distinct fibres. The fibres are selected based on them in being in clear
focus and completely intact (i.e. not half cut from the image). Further microscopy is undertaken on identical specimens during uniaxial mechanical testing up to 60% strain to analyse
the rotation of fibres in-situ and therefore verify the applicability of the assumed deformation gradient used in constitutive modelling. SEM (JSM-IT300) images of the fractured
specimen surfaces are taken to analyse the interface between the matrix and fibres, and the
dispersion of fibres.

2.3.3

Results and Discussion

The results of the uniaxial tensile testing are shown in Table 2.3, and indicate a significant
correlation between the angle of alignment and the elastic modulus; a maximum is observed
when the reinforcement is aligned parallel to the loading direction (i.e. 0 ), with a minimum
modulus at 60 . The variation of the modulus in this manner can be explained by making
comparisons to a transversely isotropic hyperelastic model, which will be discussed later.
The strength and strain-to-failure are also similarly correlated; the strength is minimised
at 60 due to the reinforcement reorienting its body rather than transferring stress, which

41

2.3 The anisotropic effects of micro-inclusions

results in the strain-to-failure being at its maximum. Conversely, when the reinforcement is
oriented towards the loading direction it efficiently transfers stress and restricts the deformation of the specimen.
Table 2.3 The average mechanical results of specimens tested to failure, standard deviations
shown in brackets. Indicating, the Youngs modulus, E, calculated from based on the fitting
of the Transversely Isotropic constitutive model (Eq. 2.10) to the stress-strain response up
to 30% strain, the strain-to-failure and the strength of the specimens.
0 ( )
0
15
30
45
60
75
90

E (MPa)
2.10 (0.35)
1.90 (0.37)
1.56 (0.29)
0.71 (0.28)
0.71 (0.13)
0.86 (0.01)
0.83 (0.27)

Strain-to-failure (%)
93 (2.4)
96 (11.6)
88 (15.6)
113 (9.4)
118 (5.6)
108 (5.8)
113 (6.6)

Strength (MPa)
0.65 (0.05)
0.71 (0.11)
0.70 (0.13)
0.65 (0.22)
0.55 (0.10)
0.58 (0.06)
0.60 (0.13)

The nominal stress-strain behaviour of PDMS displays a weak S shaped curve (See
Appendix A.1) due to the limited extensibility of the elastomer chains (in comparison to the
S shaped stress-strain curve of Natural Rubber, e.g. see Fig. 2.9a). However due to (1)
the variability of test results that continue to diverge at high strains, (2) the applicability of
moderate strains to many elastomer applications and (3) the increased difficulty in identifying a constitutive behaviour at large strains, particularly when additional non-linearities
are present due to limited chain extensibility, comparisons between the specimens are made
only up to 60% strain.
The results of the large strain tensile tests are shown in Fig. A.1 for selected specimens
of each orientation up to 60% strain, and include confidence intervals calculated from the
average stress-strain curve of all specimens of each reinforcement angle. The stress-strain
curves highlight the different stress transfer mechanisms that take place for different orientations of the reinforcements. When the fibres are oriented towards the loading direction
(0 -30 ) they are able to carry the majority of the load and a high modulus is achieved; on
the contrary, low moduli correspond to configurations where the fibres are almost orthogonal
to the loading (60 -90 ).
It is interesting to note the effects of the reinforcement reorientation during deformation
and its effects upon the stress, which can be evidenced when comparing the reinforcement
at 0 and 30 in Fig. A.1a. The initial modulus of 0 is higher, but at = 1.35 the rotation
of the fibre towards the loading direction causes the 30 tangent modulus to decrease less
and the stress to rise above the other. A similar occurrence is not observed at higher angles

Experimental Mechanical Characterisation

0.6

00
30
60
75

0.5
0.4
0.3
0.2

(a)

0.1
0
1

1.1

1.2

1.3

1.4

1.5

1.6

Nominal Stress (MPa)

Nominal Stress (MPa)

42

0.6

15
45
90

0.5
0.4
0.3
0.2

(b)

0.1
0
1

1.1

Stretch,

1.2

1.3

1.4

1.5

1.6

Stretch,

Fig. 2.21 Large strain tensile experimental results of magnetically aligned fibre reinforced
PDMS, up to 60% strain. (a) Reinforcement angles 0 , 30 , 60 and 75 , (b) Reinforcement
angles 15 , 45 and 90 . The 90% confidence bars, calculated from 6 tested specimens at
each angle, are shown at strain intervals for each specimen.
(e.g. between the 60 and 75 specimens), however this is most likely due to the relatively
moderate change in modulus that a rotation of the fibres would cause at these angles (see
Fig. 2.23).
The behaviour described is highly non-linear, especially at large strain, however by comparing the behaviour with a transversally isotropic hyperelastic model, it will allow further
understanding of the material behaviour and allow the constitutive behaviour to be applied
to further modelling.
The strain energy, M , of the model has the following form,
M (I1 , I4 ) =

[(I1 3) + g(I4 )] p(J 1)


2

(2.5)

with

1
g(I4 ) = ( I4 1)2 ( I4 + 2)
I4

(2.6)

with the stress-strain relationship derived under the assumption of a uniaxial deformation gradient, i.e.

0
0
1

FM = 0 2
0 ,
1
0
0 2

(2.7)

This form of strain energy is essentially the Neo-Hookean model [241] with an addi-

2.3 The anisotropic effects of micro-inclusions

43

tional contribution to account for the transversely isotropic distribution of fibres, represented
by the dimensionless parameter, . I1 and I4 represent the 1st and 4th strain invariants, defined in section 3.1.1. A more detailed description to the model can be found in section 3.2.
However, the nominal stress can be derived as:

2
1

N11 = 2 +
cos (0 )

(2.8)

with the auxiliary function, , given by:


(1 + )2 (1 + )(2 + )
+

=
2 2
2

(1 + )2 (2 + ))
2 3

(2.9)

Comparisons of the constitutive model to the experimental data, assuming uniaxial tension loading, can be seen in Fig. 2.22. At strains above 30% the model begins to struggle
to capture the highly non-linear behaviour observed in the experimental data. Therefore a
more complex form of the constitutive parameters 1 , 2 and 4 may be considered, such
as those seen in [165] but with such complicated models it is difficult to gain an understanding of the underlying mechanics of the material behaviour. Instead, the model here is applied
with the limitations understood, and attempts will be made to understand the reasons for the
difficulty in defining a constitutive model. Even still, it will be shown that even with deformations below 30% strain it is possible to observe significant non-linear behaviour which is
highly dependent upon the reinforcement direction.
The accurate fitting at small strain suggests the applicability of the model to small deformations, and so the expression for the effective elastic modulus E11 is obtained (Details
in Section 3.2),
E11 (0 ) =

3
(16 + 5 + 8 cos (20 ) + 3 cos (40 ))
16

(2.10)

As expected the effective elastic modulus depends on the fibre orientation. In particular
E11 has a minimum for 0 = 65.9 ; in this configuration the fibre provides the least amount
of resistance to the uniaxial forces imposed upon it. On the contrary, for 0 = 0 or 0 = 90 ,
the effective modulus is a local maxima. The maximum at 0 = 0 being significantly larger
than at 0 = 90 . However, the limiting value at 0 = 90 is interesting as it indicates
the potential importance of the lateral constraint of the reinforcement in an incompressible
system, a phenomenon that is experimentally observed (See Fig. 2.23).

Experimental Mechanical Characterisation

0.6

00
30
60
75

0.5
0.4
0.3
0.2

(a)

0.1
0
1

1.1

1.2

1.3

1.4

1.5

Nominal Stress (MPa)

Nominal Stress (MPa)

44

15
45
90

0.6
0.5
0.4
0.3
0.2

(b)

0.1
0
1

1.6

1.1

1.2

Stretch,

1.3

1.4

1.5

1.6

Stretch,

Fig. 2.22 Large strain tensile experimental results compared to the transversally isotropic
constitutive model, Eq. 2.5, up to 60% strain. Markers indicate the experimental results,
lines indicate the constitutive model fitting. (a) Reinforcement angles 0 , 30 , 60 and 75 ,
(b) Reinforcement angles 15 , 45 and 90 . The 90% confidence bars, calculated from 6
tested specimens at each angle, are shown at strain intervals for each specimen.

Initial Modulus (MPa)

2.5
2
1.5
1
0.5
0

20

40

60

80

Reinforcement Angle, 0 (o )

Fig. 2.23 Small strain behaviour of the material, averaged from all specimens and showing
the standard deviation. The model (Eq. 2.10) fitting is shown to fit the behaviour well.
( = 0.252, = 1.85).

The stress-strain relationship is derived under the assumption of a uniaxial deformation


gradient. To analyse the applicability of the applied uniaxial tensile loading in the constitutive model, the rotation of the fibres is observed, in-situ, during tensile testing at strains
values up to 55% (see Fig. 2.24). This has allowed the rotation of individual fibres from the
centre of each specimen to be captured and compared to rotations predicted by the model,
which is obtained by using Nansons formula (see, e.g., Ref.[184]). 0 is the direction of the

2.3 The anisotropic effects of micro-inclusions

45

fibre in the undeformed configuration, thus its direction is given by N = (cos(0 ), sin(0 ), 0),
whilst the direction normal to the platelet is A = ( sin(0 ), cos(0 ), 0). The current normal vector a is a = FT A1 FT A , that in view of the imposed deformation gives the
current orientation
of the fibre,
h
i , in terms of the original angle, 0 , and the stretch, , i.e.,
= tan1 3/2 tan(0 ) .
The results of the fitting with this formula are shown in Fig. 2.25a for five specimens
with initial angles 0 , 15 , 45 , 75 and 90 . A remarkably close agreement between the
rotation predicted by the model and the actual rotation of the fibres is apparent up to 55%
of deformation for fibres observed in the centre of the specimens, which suggests that discontinuous reinforcements are well approximated as continuous fibre reinforcements in this
context.

Fig. 2.24 Experimental micrographs of a reinforcing fibre during uniaxial tensile tests at
different strain levels.

Fig. 2.25 The experimental rotation of the fibres compared tohthe rotations predicted
by the
i
1
3/2
model, obtained through Nansons formula, i.e. = tan

tan(0 ) .

46

Experimental Mechanical Characterisation

However, as shown in Fig. 2.22, there are significant discrepancies in the experimental
data at large strain that can not be described by the model and suggest that the angle of
isotropy may be significant. There are a number of possibilities that could help to explain
the deviation from the assumed transversally isotropic model:
Whilst the rotation of the fibre reinforcements helps to validate the plane stress assumption of the model, Fig. 2.26 shows that the lateral constraint of the reinforcement may cause
discrepancies as the reinforcement angle increased (i.e. 90 ). This has been analysed
by digital image correlation (DIC) on each specimen, that tracks the deformation and calculates the stretch components parallel and perpendicular to the applied force. The results
in Fig. 2.26 show that for specimens with the reinforcement aligned towards the loading
direction (i.e. 0 - 30 ), the specimens are well described by the deformation gradient (2.7)
8 . As the reinforcement angle increases towards 90 , there is a gradual reduction of in
2
comparison to 1 ; which is further assessed in the following.

Lateral Stretch, 2

1
0.95
0.9
0.85
0.8
1

00
15
30
45
60
75
90

1.2
1.4
Longitudinal Stretch, 1

1.6

Fig. 2.26 The lateral stretch, 2 vs longitudinal stretch, 1 of selected specimens of each
reinforcement angle, compared to the theoretical plane stress (dashed line) prediction, i.e.
2 = 10.5
This demonstrates that the lateral constraint of the reinforcement has an effect upon the
deformation and in fact at angles between 0 and 90 a shearing of the sample occurs due
to a shift in the stress transfer of the reinforcement (See Fig. 2.27b). A discussion of this
effect in experimental testing of transversely isotropic materials in the literature is included
in section 2.1.2, and discusses the difficulty in these experimental tests; particularly at large
strain. However the effects are dependent upon the angle of isotropy and the strength of the
transversely isotropic effect (i.e. the value of in Eq. (2.5)) and so a quantification of these
8 This

confirms the well known results in non-linear elasticity that simple tension implies (2.7) only if the
fibres are aligned in the direction of the load.

2.3 The anisotropic effects of micro-inclusions

47

effects is important in the range of interest.

Fig. 2.27 (a) The COMSOL Multiphysics (ver. 5.0, 2014) dumbbell numerical model, compared to the idealised constitutive behaviour, for angles 0 , 30 , 60 &90 (=0.252 MPa &
=1.85). Dashed lines refer to the idealised constitutive behaviour, markers refer to the
behaviour of a dumbbell according to ASTM D1708. (b-c) The deformation of the 45
dumbbell numerical model, indicating the lateral misalignment between the top and bottom of the gauge length, at 20% strain for parametric values (b) =0.252 MPa & =1.85
(Lateral misalignment = 0.31mm) and (c) =0.252 MPa & =18.5 (Lateral misalignment =
0.58mm). The parameter x refers to the lateral misalignment in millimetres, compared to
the undeformed configuration.
In order to assess this inhomogeneity, the constitutive behaviour is compared the behaviour of a dumbbell specimen with clamped tab boundary conditions (according to ASTM
D1708) modelled in the commercial FE software COMSOL Multiphysics (ver. 5.0, 2014).
The results, shown in Fig. 2.27a, confirm that the 0 specimen matches the constitutive behaviour, as no shear deformation is developed. However, at angles greater than 0 there
is an S-shaped deformation that occurs due to the development of shear, indicated by a
lateral misalignment between the top and bottom of the gauge length (See Fig. 2.27b&c).
The result on the stress-strain relationship is that the material is effectively softened at low
0 due to the shear modulus being considerably lower than the Youngs modulus (See the
15 and 30 specimens in Fig. 2.27a, compared to the predicted constitutive behaviour).
This illustrates the effect of the boundary conditions during the experimental testing. If
the degree of transverse isotropy were increased (i.e. a greater , relative to ) or the gauge
length were decreased, then the deviation from the constitutive behaviour would become
more pronounced.
The degree to which this S-shape develops depends on the ratio of the In-plane and
shear moduli. The shear modulus has a maximum at 45 degrees, whilst the Youngs modulus

48

Experimental Mechanical Characterisation

has a minimum at 65.9 degrees (See Section 3.2.2 for details). The consequence is that,
although the shear moduli at 30 degrees and 60 degrees are identical; there is a much more
evident S-shape for the 30 specimen because the Youngs modulus is higher and forces the
specimen to shear along the reinforcement direction (see Fig. 2.27b and Fig. 2.27c).
At larger angles however, the behaviour results in the dumbbell specimen being initially
stiffer (See e.g. the 60 specimens in Fig. 2.27a). At higher strains, due to the rotation of the
angle of isotropy, the constitutive behaviour eventually increases above these stresses. The
switch in these behaviours (i.e. when the initial modulus of the constitutive and experimental
models is identical) is found to occur at = 56.8 , although it depends on the relationship
between the shear modulus and the tensile modulus.
Whist it is shown that the deformation field causes some discrepancies in experimental
testing of transversally isotropic materials, the behaviour is marginal for the experimental
specimens (See Fig. 2.27b). However, if the strength of the transversely isotropic effect were
to increase (i.e. an increase in ) this effect would become prominent (See Fig. 2.27c). At
such a point, it would become important to consider better clamping and test fixtures that
allow the angle of isotropy to be tracked. A discussion of test fixtures for transversely
isotropic materials is given in section 2.1.2.
These effects do not explain the discrepancies at 0 or the stress increase at 30 , that are
observed in the experimental tests, whereby it increases above that of the 0 specimen at
= 1.35 (See Fig. A.1). In fact, the induced shear response would be expected to cause a
softening of the material at 30 .
Indeed, the increase in stress at large strain for the 30 specimen, compared to 0 , suggest the possible influences of snubbing friction (9) , which is the friction that results from
the fibre being pulled from the matrix at an oblique angle to the loading direction, and is
independent of the fibre-matrix interfacial properties. The evidence of fibre pull-out from
the matrix in Fig. 2.20 supports this idea and the phenomenon has been observed and numerically accounted for previously [63, 251] and can have the effect of increasing the stress
transfer as the angle of the reinforcement is oriented away from the loading direction.
Additionally, the upturn in stress (due to the limited extensibility of the PDMS elastomer
chains) could also be affected by the rotation of the fibres (as has been seen for SIC in NR
[226]), causing the strain at which this behaviour occurs to vary. Such effects would generally require an extension of the constitutive model to include additional parameters; a logical
extension to the constitutive behaviour would be to a transversally isotropic Mooney-Rivlin
model (i.e. containing one additional parameter, dependent on the invariant I2 ) however it
is shown in Ref.[40] that it is unable to satisfactorily improve the model fitting.
9 Snubbing

friction is analogous to belt friction in pulley systems.

2.3 The anisotropic effects of micro-inclusions

49

Further additional parameters would likely yield better fitting of the constitutive equations, however their addition complicates the understanding of the parameters and is beyond
the scope of this research, and in fact, many models are mostly motivated by improved constitutive fitting rather than a desire to directly model the mechanics of materials.

2.3.4

Concluding Remarks

PDMS reinforced with aligned nickel-coated short carbon fibres have been produced by
magnetic alignment during the pre-cure phase. Characterisation of the specimens confirms
that a homogeneous distribution is achieved with the degree of alignment easily assessed
by optical microscope; additionally allowing the fibre position and length distribution to be
recorded.
Observations show that the introduction of aligned fibres modifies the mechanical properties, and is dependent upon the alignment angle in comparison to the direction of the
applied load. Two observed stiffening effects are observed at 0 and 90 , termed longitudinal and lateral stiffening respectively. Similar, albeit weaker, correlations are observed in
the strength and strain-to-failure which show that a minimum of the strength occurs at 60
due to the reinforcement realigning rather than transferring the stress. This also causes the
strain-to-failure to be a maximum at this point.
The assumption of plane stress deformation is reasonable, although the lateral constraint
of the reinforcement can cause difficulties when the reinforcement is laterally oriented. Even
so, the rotation of the fibres within the elastomer is well described up to 60% strain.
A neo-hookean transversely isotropic model can fit the experimental data up to 30%
strain. Discrepancies at larger strains are likely primarily caused by the weak interface that
is shown to exist between the PDMS and reinforcing fibres.

Chapter 3
Numerical Modelling
3.1

Introduction

Long continuous reinforcements are ubiquitously found in natural materials such as softtissues and plants, and the mimicry of this behaviour is a major challenge in the optimisation of engineering design. Glass or carbon-fibre reinforced epoxy are examples of artificial
materials that try to mimic the behaviour of a natural composite and to exploit the load carrying capability offered by the continuous reinforcement. On the contrary, the mechanical
properties of composites are decreased when the reinforcements are discontinuous but they
do give a number of advantages such as increased ductility, formability and process-ability.
The understanding and prediction of this material behaviour is a key requirement in
engineering design, however many challenges exist; particularly in understanding the large
strain behaviour of materials. The anisotropy of the reinforcement has little effect on the
material behaviour at the macroscale if a homogeneous and isotropic dispersion is achieved,
but by deliberately controlling the orientation of the reinforcement it is possible to exploit
the anisotropic behaviour. The behaviour of anisotropically reinforced elastomers can be
highly non-linear, as seen in section 2.3, due to a number of factors that include:
The material heterogeneity contrast of the constituents.
The morphology and shape of the fillers [157].
The initial orientation and subsequent rotation of fillers during deformation.
The variable length and orientation of the fillers [64].
The condition of the filler coating and the subsequent interfacial strength [65].

52

Numerical Modelling
The distribution and dispersion of the fibres [201].

A fundamental problem with particle reinforced composites is to predict their overall behaviour in terms of the mechanical properties of the constituents, particularly the anisotropic
response at large strain. Most papers in the literature have focused on macroscopically
isotropic composites with a random or nearly random distribution of reinforcements. At
small strains, the Guth-Gold equation is used to predict the dependence of linear moduli on
filler/matrix mechanical properties for spherical and rod shaped filler particles [206]. Halpin
and Tsai developed a semi-analytical approach based on Hills "self-consistent" method to
predict the elastic modulus of composites reinforced with a variety of filler geometries ranging from spheres to long fibres (see, e.g. Ref. [91]). Whilst these methods have been successfully applied for a variety of different materials, Finite element analysis is required for
more advanced morphologies and distributions [157].
At large strain, the analytical approach is made difficult by the presence of geometrical
and constitutive nonlinearities. The pioneering works of Hill [102] and Hill & Rice [103]
have shown that it is possible to find an effective strain energy function able to describe
the overall behaviour of the composite. This has led to the development of a number of
approaches [85]. In fact, efforts to understand many of the effects that control the large
strain performance of reinforced materials can be found in the literature [104, 160]. In
particular, Gasser et al. [72, 187] developed a constitutive model to look at the effects of
fibre angle distribution on the large strain behaviour of arterial walls.
The benefit of such a model is the relatively small computational cost compared to alternative methods such as the finite element method and molecular dynamics. However it is
difficult to develop a model that can account for even relatively simple configurations, and
often only particular boundary value problems are considered [158]. More general settings
require the use of the finite element method, and accompanying numerical homogenisation of a representative volume element. Recently, numerical homogenisation was used by
[86], [176] and [88] to study the large strain behaviour of Incompressible Particle Reinforced neo-Hookean composites (IPRNCs). In particular Guo et al. [88] simulated the 3D
reinforcement of a representative volume element (RVE) with multiple spherical-shaped inclusions. The authors have shown that when both the filler and the matrix are neo-Hookean
the overall response is still neo-Hookean and the apparent shear modulus can be predicted
through the Halpin-Tsai equation.
In this chapter, motivated by the need to understand the highly non-linear behaviour of
anisotropic particle reinforced elastomers, we use a numerical homogenisation approach to
investigate the mechanical behaviour of the composite at large strain under general boundary
conditions. The numerical model is a simplified representation of the reinforced elastomers

3.1 Introduction

53

introduced in section 2.2 and section 2.3, and allows the assumptions of the constitutive
model can be explored. In this way, the validity of the constitutive equation to describe the
experimental behaviour can be assessed. The numerical results show that the orientation
of the platelet induces an anisotropic response that can be adequately described by a generalised neo-Hookean model in which the stiffness depends on the fibre stretch, when both
the reinforcement and the matrix are assumed to behave like incompressible neo-Hookean
solids.
The combination of a constitutive model, a finite element (FE) model and the experimental results of section 2.3, allow for greater understanding of the material behaviour
of anisotropic dis-continuous fibre reinforced elastomers and indicate to what extent the
transversally isotropic neo-Hookean constitutive model can be applied, and give an assessment of the degree to which a simplified FE model can describe the non-linear composite
behaviour.

3.1.1

Literature Review

The Elastic Response of Reinforced Composites


The understanding and prediction of the behaviour of engineering materials is a prerequisite
of design and in the development of new materials. To address this, many different models
have been proposed that vary in complexity in the assumptions of the constituents, the loading conditions and the information extracted. The essential goal is to define a relationship
between the stress and strain, which for a linear elastic material can be related by an average
stiffness tensor for a composite, ie. = C.
One of the fundamental approaches which has been adapted in later models is the equivalent inclusion theory of Eshelby [55], which considers the effect of an ellipsoidal inclusion
in an infinite elastic body. The theory considers an initially homogeneous stress-free body,
from which an ellipsoidal inclusion undergoes a transformation. If it were a separate body
to the infinite medium then it would undergo a uniform strain T , however, due to the bond
between the inclusion and medium, the composite actually develops a strain field, C . The
relationship between the two strain tensors is then given by c = E Eshelby T where E Eshelby
is known as the Eshelby tensor, which depends on the inclusion aspect ratio and matrix
elastic properties.
A second scenario is considered where the inclusion undergoes no transformation, but
has a different stiffness. By subjecting both bodies to a uniform strain, Eshelby was able to
find a strain concentration tensor to relate the average strain in the composite to the strain
in the inclusion. The tensor, A, can then be used to relate the stiffness of the constituents to

54

Numerical Modelling

the average stiffness of the composite, i.e. E = Em + v f (E f Em )AEshelby , where AEshelby


is the strain-concentration tensor, Em and E f are the stiffnesses of the matrix and fibres,
respectively, and v f is the fibre volume fraction. A more detailed derivation can be found
in the literature [108, 180]. However the theory only considers dilute composite materials
(low concentrations) and no particle interactions.
Mori-Tanaka [177] developed a model from Eshelbys theory that considered the effect
of multiple identical inclusions, whose derivation is methodically explained by Benveniste
[8]. Whilst the equivalent inclusion theory calculated a tensor to relate the strain in the
inclusion to that of the composite, i.e. f = AEshelby , Mori-Tanaka assume that multiple
fibres would relate to the average strain of the matrix instead, i.e. f = AEshelby m . The
result is an alternative extension of the Eshelby strain-concentration tensor,
AMT = AEshelby [(1 v f )I + v f AEshelby ]1 .

(3.1)

This theory is commonly used in the literature [58, 98], and has been further developed
to consider effects such as fibre distribution [35].
Alternatively to the equivalent inclusion theories, a self-consistent method has been developed, originally by Hill [101] and Budiansky [25]. In this scheme, the inclusion is assumed to be embedded in an infinite medium that has the average properties of the composite. In this way, the transversely isotropic properties tensor of short-fibre reinforced
composites can be calculated by an iterative process [145]. A generalised self-consistent
model has also been developed , in which Hermans [99] considered that both the fibre and
matrix are embedded in an infinite medium. From this work, the Halpin-Tsai equations were
developed [91] by finding a common form for expressing the moduli of composites, i.e.
1 + v f
(E f /Em ) 1
E
=
where, =
.
Em
1 v f
(E f /Em ) + 1

(3.2)

In this form, E represents one of a number of different moduli and is chosen based
on the particular moduli being considered. The Halpin-Tsai equations are commonly used
to estimate the mechanical properties of composites based upon knowledge of the material
constituent properties (See e.g. [58, 108]) however the many experimental variables in
testing mean the accuracy varies, so an alternative method is to set bounds upon the material
properties.
The first models of this type are the Voigt and Reuss bounds, which assume the fibre
and matrix have the same uniform strain or stress, respectively. The Voigt bound is then
found by minimising the potential energy, and calculates an upper bound to the stiffness.

55

3.1 Introduction

Alternatively the Reuss bound finds the maximum of the energy, and calculates a lower
bound of the composite stiffness. The results assume an isotropic composite, with very
wide bounds; but by giving a reference material either infinite or zero stiffness, the upper
and lower bounds can be tightened [97] and extended to anisotropic materials [246].
Bounding methods are useful because of the difficulty in predicting the material properties of composites due to the uncertainty in the material properties caused by, for example,
the dispersion of fibres and distributions in their lengths. However they give wide bounds
that are difficult to design from. In consideration of the effects of fibre length and angle distribution, Fu and Lauke [64, 67] developed a theory based on the laminate analogy approach
(LAA), which used distribution functions to represent the variability in the microstructure
of the fibres and its effect upon the elastic response. The method can be used to consider
many variables within a composite such as particle size, interface and vol% [65], and shows
reasonable agreement to experiments when the fibres are well aligned [119].

Constitutive Modelling of the Large Strain Response of Composites


However, in order to describe the behaviour of an elastomer material it is important to consider its non-linear behaviour. Moreover it is important, in general, that any model should be
capable of describing a material when subject to any feasible deformation gradient. For example Gumbrell et al. [82] obtained significantly different results, inconsistent with theory,
under equi-biaxial extension, although tension data fitted the model well. The difficulty in
defining a constitutive theory for generalised strain, that is based on the material molecular
and microstructural composition, is evident from the comparative number of phenomenologically based models that are derived based on observations of their stress-strain behaviour
from an entirely mathematical point of view. From this context, it is important to introduce
the strain invariants that are used in the majority of phenomenological models.
To quantify the deformation of a body, it is convenient to consider the deformation
gradient, F, which is the derivative of each component of the deformation vector, x, with
respect to its undeformed counterpart, X, so that,

x1
X1

xi
x
= 2
F=
X j X1

x3
X1

x1
X2
x2
X2
x3
X2

x1
X3

x2

X3

x3
.
X3

(3.3)

The eigenvalues of the deformation gradient are known as the principal stretches, given

56

Numerical Modelling

by 1 , 2 and 3 . Rivlin [208] considered that even powered functions of these three principal stretch ratios would satisfy the constraints of a material strain energy density function, ,
that was assumed isotropic and incompressible. An alternative notation is made by choosing
these principal stretches to represent the basic variants of the right and left stretch tensors
(U and V, respectively). However, the calculation of U and V is generally mathematically
inconvenient.
However, the square of the stretch tensors provide the right and left Cauchy-Green deformation tensors (i.e. C and B, respectively),

C FT F = U2
B FF T = V 2

(3.4)

which are easier to calculate, and from this it is possible to calculate the three invariants of
an isotropic material, which are independent of the choice of coordinate axes and are called
strain invariants,

I1 = tr(C) = 12 + 22 + 32
I2 = tr(C1 ) = 22 32 + 32 12 + 12 22

(3.5)

I3 = det(C) = det(B) = 12 22 32
where tr and det indicate the trace and determinant of a matrix, respectively. For an incompressible material, I3 is constant and equal to 1. As the majority of elastomers are
incompressible, in the following only incompressible constitutive models are considered.
One of the simplest theories to describe the behaviour of elastomers is the neo-Hookean
model, which is equivalent to the statistical theory of networks, Eq. 3.6. The statistical
theory of networks is a physically-based model that considers the effect of a number of longchain molecules linked together at discrete points (e.g. vulcanisation) to form a 3D network
and provides a good approximation of the behaviour of rubber under different types of strain.
The strain energy density, i.e. the work in deforming the elastomer, can be expressed as,
1
= NkT (I1 3)
2

(3.6)

where, N is the number of network chains per unit volume, k is the Boltzmann constant
and T is the absolute temperature. The physical basis of the model has led to its critical
assessment, and the range of its validity explored. Whilst the theory is useful for a first

57

3.1 Introduction

approximation of the behaviour of a rubber material within a general strain context, at large
strains the statistical network theory underestimates the stresses due to the upturn in stress
caused by SIC and chain extensibility [241].
An extension of this theory that considers the dependence of the behaviour on the invariant I2 as well, was developed by Mooney [175] and commonly referred to as the MooneyRivlin equation due to the work of Rivlin [208] in expanding the theory. It is derived under
the assumption that Hookes law is obeyed in simple shear, which is known to be true up to
moderate strains of around 100% [82].
= C10 (I1 3) +C01 (I2 3)

(3.7)

When considering only simple extension, the Mooney-Rivlin equation gives a better
approximation of the non-linear behaviour than the neo-Hookean model, however the theory
is unable to satisfactorily improve upon the neo-Hookean model when additional strain types
are considered [241].
More complex even powered functions of i can be expressed in terms of the three
variants, and so Rivlin [208] showed that the strain energy function of an incompressible
isotropic elastic material may be expressed as the sum of a series of terms,

Ci j (I1 3)i (I2 3) j

(3.8)

i=0, j=0

For i=1 and j=0 we get the neo-Hookean model, which is the simplest form of the series.
i=0, j=1 can be used, but this part of the series has no physical basis for rubber on its own
but can be added to form the Mooney-Rivlin equation.
However, the difficulty of these models to describe the large strain behaviour (such as
the upturn in stress), without the addition of many parameters, has led to a number of different elastomer material models being developed. A few notable models that describe the
behaviour up to relatively large strain, for general strain, will be presented.
Rivlin and Saunders [209] proposed that the dependency on I2 should be a function
decided by experimental testing, rather than a constant as in the Mooney-Rivlin equation,
i.e. C01 = f (I2 3). From this Gent and Thomas [74] proposed the following form of strain
energy,
= C1 (I1 3) + k ln(I2 /3)

(3.9)

that has a non-linear dependence on I2 and requires only two parameters to be fitted.
By considering the limited extensibility of polymer chains and the resulting observed

58

Numerical Modelling

non-linear behaviour, Hart and Smith [95] suggested a non-gaussian 3 parameter formula to
account for the rapid increase in stress at high strain.



k2

2
(i)
= G exp k1 (I1 3) (ii)
=G
I1
I2
I2

(3.10)

The use of the invariants is only for mathematical convenience. Therefore, Ogden [183]
ignored the restriction to even powers of extension ratios and wrote a power series for an
incompressible rubber that is qualitatively similar to the Mooney-Rivlin model but is able
to model the strain-hardening effect at large strain more effectively.
=
n

n N
( + 2N + 3N 3)
n 1

(3.11)

It has been shown that, for simple extension and pure shear, 4 parameters can be used
to match experimental data, but when additionally considering equi-biaxial extension 6 parameters were required to match the data up to 600% strain [241].
Arruda and Boyce [6] proposed a model that assumes a unit sphere with a distribution of
chains in eight directions along the vertices of a cube, and due to the central junction of the
chains and the symmetry that results, the deformation is related to the first invariant. The
integration of the stain energy is difficult to compute, although it can be presented in the
polynomial form as,


1
1 2
11
(I1 ) = G (I1 3) +
(I1 9) +
(I13 27)
2
2
20N
1050N
 (3.12)
19
519
4
5
+
(I 81)+
(I 243)...
7000N 3 1
673750N 4 1
where N is the number of rigid links composing a single chain. The model is physically
based, as the parameter G is the same as in the statistical theory of networks, i.e. G = nkT ,
and requires only 2 parameters to fit to experimental data.
Many more constitutive equations are available that can describe isotropic materials,
that are either mathematically motivated (see e.g. the Gent [75] and Alexander [3] models)
or physically motivated (see e.g. the Yeoh [261], extended-tube [121] and van der Waals
[124] models). A comparison of these methods, and others, can be found in the literature
[21, 117, 165]. However, in general, it is necessary to include a large number of parameters
to describe the full stress-strain curve of an elastomer and so much of the physical interpretation is lost. Up to moderate strain levels ( 60%) the neo-Hookean model is generally
considered a good choice, due to the good fitting of the physically motivated function with

59

3.1 Introduction
only one parameter [165].

Constitutive Modelling of the Large Strain Response of Transversally Isotropic Materials


When the elastomer material is anisotropic there is generally a requirement for more complicated models, especially when considering the application of complex strain. A special
case of anisotropy is found when the properties of the material are aligned to 3 orthogonal
directions in the material; such a material is labelled orthotropic. If two of these directions
have equal properties then the material can be considered as transversely isotropic; such a
material might be an aligned fibre reinforced elastomer (See Fig. 3.1).

Fig. 3.1 The transverse isotropy of a fibre. Indicating the plane of isotropy orthogonal to the
principal fibre direction.
The transversely isotropic assumption leads to a significant simplification of the constitutive relationship, although the ability to describe a material successfully in an experimental
context can still be challenging. A strategy to simplifying the derivation of a stress-strain
relation is to have an additional term in the strain energy function that is decoupled from
the isotropic contribution, i.e. (C) = Isotropic + Anisotropic . Within this strain energy
function two additional invariants can be added,

I4 = A CA
I5 = A C2 A

(3.13)

where A is the axis of transverse isotropy in the reference configuration, i.e., the initial orientation of the platelet shown in Fig. 3.1. The dot product, , is defined as A B ni=1 Ai Bi .
A is a vector in the direction of the transverse isotropy in the reference configuration,
indicating the initial orientation of the reinforcement shown in Fig. 3.1, I4 is related to the

60

Numerical Modelling

square of the stretch in the direction of A, i.e. f2ibre , whilst I5 has no direct physical meaning
but can be related to fibre/matrix shear interaction [191].
A resulting form of the strain energy, assuming decoupling of the isotropic and anisotropic
components, can be given as,
(I1 , I4 , I5 ) = Isotropic (I1 , I2 , I3 ) + Anisotropic (I4 , I5 )

(3.14)

and the Cauchy stress tensor for an incompressible material is given by [45],
= pI + 21 B 22 B1 + 24 a a + 25 (a Ba + Ba a)

(3.15)

where p is the lagrange multiplier to account for incompressibility, i / Ii , and a


FA. is used to indicate the tensor product defined by (a b)i j = ai b j . However due to
the computational difficulty in such an equation, especially in the dependence on I5 , the
form of the strain energy is typically reduced to depend on only I1 and I4 for mathematically
simplification. A common model of this form is the so-called standard reinforcing model,


c
2
(I1 , I4 ) = I1 3 + (I4 1)
2

(3.16)

where c is that of the standard neo-Hookean model, and is a non-dimensional material


constant to measure the effectiveness of the reinforcement, and used in [45, 181].
It is important to carefully consider the whole deformation response from the constitutive model, for instance, at large values of the material can show a non-monotonic relationship between the stretch and stress (i.e. for a compressive uniaxial load at > 4.961
[199]). Hence, Guo et al. [87] modified the standard reinforcing model in order that it gives
monotonic behaviour in compression.
Similar models that depend on the fibre stretch, I4 , for the anisotropy material behaviour
have been developed (See e.g. [105, 106]), however it is also possible to construct a strain
energy dependent on I4 that considers the effects of fibre/matrix interaction [85, 249]. An
equivalent form of the standard reinforcing model has also been investigated by Merodio &
Ogden that depends on I5 , rather than I4 due to the dependence of I5 on the shear behaviour
[171],


c
2
(I1 , I4 ) = I1 3 + (I5 1)
2

(3.17)

however no significant advantage of this strain energy function is evident, and similar nonmonotonic behaviour can be observed in compression due to the finite strain energy in compression.

3.1 Introduction

61

The difficulty to develop a mathematically simple constitutive equation that adequately


describes the anisotropic behaviour under general loading conditions has led to many different models being developed. One of the most common formulations is that developed
by Holzapfel et al. [104], who developed a phenomenological transversally isotropic constitutive model to describe the mechanical response of arterial tissue, commonly referred
to as the Gasser-Ogden-Holzapfel model. Whilst the underlying behaviour of arterial walls
is similar to that of fibre-reinforced elastomers, the double layered material means that the
model considers the effects of two fibre families, given by the vectors a1 and a2 . In this way
there are additional available invariants, where I4 is the strain invariant in the direction a1 ,
and I6 is the equivalent invariant in the a2 direction.

n 
o

k1
c
2
exp
k
(I

1)
= iso (I1 ) + aniso (I4 , I6 ) = (I1 3) +

1
2 i

2
2k2 i=4,6

(3.18)

Here c is a parameter of the neo-Hookean type, k1 is similarly a stress-like material


parameter in the fibre direction, whilst k2 is a dimensionless parameter. The exponential
form of the strain energy density is motivated by the behaviour of collagen fibrils, which
show a strong stiffening effect at high strain, analogous to strain-induced crystallization.
Despite being intended to describe the behaviour of arterial walls, the model is found
to describe many transversely isotropic materials, and the extension to this theory, that considers fibre angle distributions (Ref. [72]), has been implemented in the commercial FE
software ABAQUS (ver. 6.12). A further extension of this theory, that is intended to give
a more accurate representation of the fibre distribution by including higher order terms, has
been presented in Ref. [187], whilst other models that include the effects of fibre angle
distribution can be included [32, 33].
The accuracy of a model usually depends on an inherent knowledge of the underlying
constituents of a material which can be particularly difficult for phenomenologically motivated models. Guo et al. [87] attempted to make some quantitive parameters in terms of the
homogeneity contrast and vol% of the constituents, although in general a homogenisation
scheme is required. Homogenisation schemes at small strain have been discussed, however at large strain additional geometric and constitutive non-linearities exist that make the
constitutive approach difficult.
However, it is shown by the pioneering work of Hill [102] and Hill & Rice [103] that an
effective strain energy function can be found to describe the overall constitutive behaviour
of a composite material of more than one constituent, i.e. the Hill-Mandel condition. Upper
and lower bounds on the strain energy, of the Voigt and Reuss types, were later found by

62

Numerical Modelling

Ogden [185] and Ponte Castaeda [195].


Various research has since focused on particular boundary value problems, for instance,
Lopez-Pamies et al. [158] used an iterative homogenization technique to show that for
two-phase neo-Hookean composites with particulate microstructure the overall behaviour
is neo-Hookean for axisymmetric loading conditions. In Ref. [161], the constituents of a
composite are assumed to take a layered configuration and shows the effect of varying the
angle of transverse isotropy on the large strain behaviour. An advantage of a homogenisation
scheme can be the ability to relate the microstructure of a material directly to the constitutive
behaviour (i.e. size, shape, orientation and mechanical properties of the constituents). Lopez
et al. used the homogenisation theory of Ref. [159], to develop a constitutive homogenised
model of oriented reinforcing fibres in a hyperelastic matrix. The model is able to relate the
strain energy function of the overall composite to the fibre volume fraction, fibre angle and
heterogeneity contrast of the constituents,

( I4 + 2)( I4 1)2

(I1 , I4 , c0 , i ) = (I1 3) +
2
2
I4

(3.19)

where and are both functions of the volume and the properties of the constituents.
The attraction of such methods is the relation of the constituent properties to the overall composite behaviour, whilst phenomenological models are usually designed firstly for
mathematical convenience. The majority of constitutive models are developed for biological
tissues (See e.g. [104, 105, 249]), and then subsequently find applicability in similar material systems. However, particular elements of the material behaviour are often different and
require more specific models to be developed, and homogenisation schemes are difficult to
develop and generally only applicable to a specific range of material behaviour and loading.
An alternative approach is to develop a constitutive model directly from the experimental
behaviour [24]. Whilst generally a number of parameters are needed for the models of this
type found in the literature (See e.g. [111] which fits 6 parameters to describe the nylon fibre
reinforced rubber compound used in an industrial v-belt), care must be taken to ascribe the
parameters to any deformation or loading condition not explicitly considered in the original
experimental data.
The difficulty to obtain experimental data for isotropic rubber has already been discussed, and obtaining reliable data for an anisotropic material is only made more difficult.
It is said that a transversely isotropic laminate can be characterised by an experimental
tension test of the material when the fibre is aligned at 0 , 45 and 90 from the tensile direction [115], however it is inevitable that, for example, the compressive, bi-axial and shear
responses of an anisotropic material will behave very differently than that predicted by just
an examination of the tensile performance. Limited experimental data is therefore available

3.1 Introduction

63

to validate any constitutive equation (See e.g. Table 1 of Ref.[24]) due to the difficulties in
accurate experimental testing, some of which is discussed in section 2.3. To assist in the
accumulation of experimental data, it is possible to supplement it with analyses of materials
in the finite element environment.
Identification of the constitutive behaviour from the mechanical response of an RVE
The homogenisation of materials in a finite element environment, when considering the microstructure of the constituents, typically consists of a unit-cell or representative volume
element (RVE) of the material (Fig. 3.2). A unit-cell is a model that represents some periodic feature of the microstructure, whilst an RVE is the smallest representative volume of
microstructure information that can represent the overall behaviour of the macroscopic material [73]. Whilst the terms may be interchangeable in some instances, the unit-cell would
be unable to capture the effects of, for example, fibre dispersion and distribution.
Simplicity in the design of the RVE is advantageous, both in reducing the computational
cost and in simplifying parametric studies of a model; however it is important that it can
capture the realistic behaviour of the material (within the limits required of the model). A
consideration of randomness and variance is usually found to be important in the behaviour
of materials.

Fig. 3.2 The smallest section of the material microstructure that can model the properties
of the overall composite is called the representative volume element. An assumption of
periodicity in the microstructure allows a further simplification to a unit-cell.
The behaviour of many materials can often be predicted by a consideration of a single
reinforcement [36, 255], particularly when low concentrations of filler are considered, so

64

Numerical Modelling

that the effects of fibre-fibre interactions are minimised. However, when the interaction of
randomly distributed reinforcements becomes significant, as a result of processing or due to
an increase in concentration, either multiple RVEs or an RVE with multiple reinforcements
are required. Sheng et al. [221] showed that the effects of particle interaction played a
role in the efficiency of platelet reinforcements in a 2D plane strain RVE, resulting in a
reduced modulus when a platelet is shielded from the strain field by another platelet. This
can mean that the effective aspect ratio of the reinforcements reduces, due to stacking of the
reinforcements [98].
It has been shown that a platelet reinforced model predicts a slightly stiffer material than
when assuming 2D plane stress [98]; although the plane stress assumption may contribute
to this effect due to the lateral stiffening expected of a platelet reinforcement, that a plane
strain assumption would consider. The small strain effects of alignment and reinforcement
angle are also of particular interest to anisotropic materials [116, 247], as are the effects
of reinforcement geometry [247]. At large strain, however, the effects can become more
important; especially as the reinforcement mechanism can change during the deformation.
Kouznetsova et al. [137] showed that the difference between a single void RVE and a
multiple void RVE is minimal at small strain for an elasto-visco-plastic matrix. However, as
the strain is increased, the results begin to deviate slightly due to development of additional
plasticity between the voids. Similar results are found in Ref. [113] using the hyperelastic Gent model [75] to model CB reinforced rubber at large strain, which show that the
boundary conditions become very important at large strain, and multiple reinforcements are
required to ensure the RVE doesnt overestimate or underestimate the stress.
The results obtained from finite element analysis (FEA) can often be useful to explore
the mechanisms of a composite material, and can be used as a method to efficiently explore
the design space of the constituents in a controlled environment, especially when the loading
conditions and constituents of experiments are difficult to apply and control. However, they
also assist in the validation of less computationally expensive constitutive models. Guo et
al. [83, 84] investigated the effects of circular voids on the mechanical properties of an RVE
subject to different loading conditions, and were able to report the ability of a transversally
isotropic constitutive relationship to fit the data. Similarly, it has been shown that when a
filler and matrix are both neo-Hookean, the overall response is also neo-Hookean and the
apparent shear modulus can be predicted by the Halpin-Tsai equation.
The case of a fibre or platelet reinforcement is expected to differ significantly to that of
a spheroid reinforcement, mainly due to the effects orientation and rotation that will affect
the performance of the composite during deformation. Efforts to describe the behaviour of
fibre reinforced composites has shown that a transversely isotropic model can fit the nu-

3.1 Introduction

65

merical simulations well in general loading conditions up to 20-30% strain, and that the
transversely isotropic model can be related to the constituents of the RVE [44] or that the
Halpin-Tsai can be fitted well to the small strain results, and used to predict some features
of the large strain behaviour [86]. Both studies considered a unit-cell to represent the material behaviour, however the effects of fibre orientation, rotation during deformation and
dispersion/distribution are not possible to explain from such a model; in this case an RVE is
required.
A 3-dimensional multiple fibre reinforced slightly compressible RVE, developed in Ref.[77],
shows the effect of a random distribution of aligned fibres on the transversely isotropic behaviour of the model, which is compared to hyperelastic constitutive relationships. The
stress-stress relationship is investigated up to 20% strain, and is able to use the periodic
boundary conditions to model multiple reinforcement angles with one RVE model. A parametric study of the reinforcement angle is undertaken for an RVE with randomly dispersed,
perfectly aligned fibres of length/width ratio 20 and a heterogeneity contrast between the
matrix and reinforcement of 25. The volume content of the fibres is varied between 10 and
25 vol%, and shows that a transversely isotropic form of the strain energy can represent the
material behaviour well up to 10-20% strain.

66

3.2
3.2.1

Numerical Modelling

Constitutive Modelling
Introduction

The determination of a constitutive model to describe the behaviour of different materials


is an active area of research [165], particularly when the micro-structure presents interesting features to the overall behaviour (See, e.g., the mechanical and stability response of
fibre-reinforced elastomers [160]). However, experimental testing of materials can often be
challenging [256] and so the constitutive models are compared to FE simulations to determine the limitations of the models in different loading conditions [86].

Fig. 3.3 A fibre reinforced thin elastomer specimen (t << L) can be simplified to a multiple
fibre RVE, assuming plane stress in the through thickness direction. In addition, a platelet
reinforced elastomer can be assumed to have plane strain in the through thickness direction,
if the thickness is much greater than the length, i.e. t >> L. In both cases, a further simplification to a single reinforcement RVE can be made if an assumption of periodicity and
dilute inclusions is made.
In this section, a transversally isotropic model is introduced that is expected to describe
the behaviour of composites with aligned, reinforcing filler particles. The deformation of an

3.2 Constitutive Modelling

67

elastomer reinforced by an aligned fibre (reinforced in 1-dimension) is expected to be modelled by a simplification of the constitute behaviour to plane stress, i.e. the stress through the
thickness is assumed to be negligible and equal to zero due to the very thin specimen thickness and lack of lateral constraint, Fig. 3.3. Such an equation has shown that is can successfully model the behaviour of a fibre reinforced PDMS elastomer up to 30% strain, subject to
aligned reinforcement in the range 0 = {0 , 90 }. At higher strains, above 60-100% strain,
additional non-linear effects, such as strain-induced crystallisation and inextensibility of the
polymer chains, make identification difficult. A neo-Hookean based transversally isotropic
model should be able to describe the behaviour up to moderate strains (<60%), see e.g.
[165], but the inability of the model to describe the behaviour in the strain range 30%-60%
suggests that the aligned dis-continuous reinforcements cause non-linearities that are not
described well by a transversally isotropic material model that assumes the reinforcement is
continuous.
In contrast to fibre reinforcements, the morphology of a platelet reinforces a material
in 2-dimensions and therefore provides a lateral constraint (See Fig. 3.3); this means the
through thickness stress is not negligible. In addition, if the platelets are very long in the
through thickness direction, and the specimen thickness is much greater than the length,
i.e. t >> L, then a plane strain assumption can be enforced. A constitutive model is also
presented for this microstructure.
A real composite would behave somewhere between these two extremes.
These models will later be compared to FE simulations to determine the applicability
of the constitutive model to describe the microstore under consideration, i.e. that of an stiff
aligned dis-continuous reinforcement in an elastomer matrix.

3.2.2

The Transversally Isotropic Neo-Hookean Model

When the reinforcing particles have uniform orientation distribution, the overall behaviour
of the composite is isotropic and the corresponding stress-strain relationship at the macroscale
would follow the behaviour of the constituents [160]. On the contrary, when the composite
is reinforced with anisotropic inclusions oriented in a preferred direction, it is reasonable to
expect that their orientation has an effect on the overall mechanical response. The intent of
our analysis is to show in what limit of the aspect ratio of the platelet and the size of the RVE
the overall composite can be actually described as transversally isotropic. These models are
usually employed to describe the behaviour of materials reinforced with continuous fibres
that span the entire length of the body. When the inclusions are discontinuous, one has to
assess the predicting capabilities of the models against numerical results.
In the following it is assumed that the platelet oriented in a preferred direction cause the

68

Numerical Modelling

constitutive response at the macroscale to be transversely isotropic. To model this effect,


the set of invariants pertinent to incompressible materials introduced earlier are used, with
the added subscript M henceforth referring to the macroscopic quantity,

I1 = tr(CM ) ,

1
I2 = tr(CM
)

2
I4 = A CM A , I5 = A CM
A

(3.20)

In the most general case, the strain energy function of an incompressible transversely
isotropic material depends on all the invariants defined in (3.20): two isotropic, I1 and I2
and an additional two, I4 and I5 used to account for anisotropic effects. However, we focus
the attention on the subclass of materials whose strain energy depends only on the two
invariants I1 and I4 . This choice is motivated by the behaviour of the constituents, whose
constitutive response has been assumed to depend only on I1 m (see Eq. 3.74). The additional
dependence on I4 allows the effect of the stretch in the direction of the platelet to be weighed.
Such a form of the strain energy has been largely and successfully used in the literature to
model soft-tissues, gels and reinforced elastomers (see, e.g., [40, 45]).
Under this assumption, the macroscopic Cauchy stress, M , is given by,
M = p I + 2 1 BM + 2 4 a a ,

(3.21)

where J is the determinant of the deformation gradient and where i = M / Ii , i = {1, 4}


and is derived from the following strain energy density function, M ,
M (I1 , I4 ) =

[I1 3 + g(I4 )] p(J 1)


2

(3.22)

with

1
g(I4 ) = ( I4 1)2 ( I4 + 2)
I4

(3.23)

The model (3.22) was introduced in [44] and used in [86] and [160]. The strain energy
function, , is the sum of two different contributions: the energy associated with the overall
stretch of the composite, i.e., I1 3, plus the energy associated with the stretch along the
fibre direction, i.e., g(I4 ). This decomposition is based on the assumption that the strain

energy stored in the fibres depends only on the fibres elongation, given by I4 (see [44]).
The isotropic neo-Hookean model is recovered for = 0. Note that, in general, with this
formulation the constitutive parameter loses its significance as the shear modulus of the
composite.

69

3.2 Constitutive Modelling

Additionally it is well known that the effective shear modulus of the composite can be
estimated from the mechanical properties of the constituents in terms of shear moduli of the
filler, (p) , of the matrix (r) and the corresponding volume fractions (p) and (r) (such
that (p) = 1 (r) ). Among several formulations available in the literature, the HalpinTsai equation has been successfully used to predict the in-plane and transverse shear moduli
of composites reinforced with unidirectional inclusions [91]. Therefore when the fibre is
aligned in the direction E1 , i.e., 0 = 0 , Halpin-Tsai equation gives the following relationship between the composite shear modulus, c (0 ) = , and those of the constituents

f + 1 + (r) ( f 1)
f + 1 (r) ( f 1)

(3.24)

where f = (p) / (r) is the so-called heterogeneity contrast. Transversally isotropic materials require an additional parameter, defined as , to capture the effect of the alignment
angle, 0 (See Fig. 3.4). Formulations to describe in terms of the constituents are available in the literature for continuous reinforcement, see e.g. Ref.[160] and Ref.[86], and an
assessment of these formulations will be made in the context of dis-continuous platelet-like
reinforcements.

Fig. 3.4 A schematic representation of the RVE in the cartesian coordinates plane E1 ,E2 .

70

Numerical Modelling

Plane Strain
The plane strain deformation in the E1 , E2 plane of a Cartesian coordinate system has the
following form,

F11 F12 0

FM = F21 F22 0 ,
0
0 1

(3.25)

or equivalently in terms of the left-Cauchy Green strain tensor,

B11 B12 0

BM = B12 B22 0 ,
0
0 1

(3.26)

2 = 1 is the unit stretch normal to the plane. Here B = F 2 + F 2 ,


where B33 = F33
11
11
12
2
2
B12 = F11 F21 + F22 F12 and B22 = F22 + F21 .

Since the material is incompressible, the following relationship between the strain components holds,
B11 B22 B212 = 1 ,

(3.27)

that shows that only two components of the deformation can be controlled independently
when an incompressible material is constrained to a plane strain deformation.
If 0 is the orientation of the axis of isotropy in the reference configuration, i.e., the
platelet orientation, (see Fig. 3.4), then A = (cos(0 ), sin(0 ), 0)T while its spatial counterpart is a = (a1 , a2 , 0)T where,
a1 = F11 cos(0 ) + F12 sin(0 ) ,
a2 = F21 cos(0 ) + F22 sin(0 ) .

(3.28)

In view of Eqs. (3.25) and (3.28), I1 and I4 specialize to,


I1 = B11 + B22 + 1
I4 = a21 + a22
and from (3.21) the components of the Cauchy stress immediately follows as

(3.29)

71

3.2 Constitutive Modelling

11 = p + 2 1 B11 + 2 4 a21

(3.30)

22 = p + 2 1 B22 + 2 4 a22

(3.31)

33 = p + 2 1

(3.32)

12 = 2 1 B12 + 2 4 a1 a2

(3.33)

As expected, the only component independent of the pressure field, p, is 12 . However,


by subtracting the stress in the E3 direction, 33 , from (3.30) and (3.31) one obtains the
following stress differences,

11 33 = 2 1 (B11 1) + 2 4 a21

(3.34)

22 33 = 2 1 (B22 1) + 2 4 a22

(3.35)

that are independent of p and together with (3.33) are used to compare analytical and numerical results.
Incidentally, the chosen form of M , i.e., M (I1 , I4 ), represents the largest class of
transversally isotropic materials that can be fully identified through plane strain deformations. I.e. The previous equations involve two unknown functions, 1 and 4 , and two
independent components of the deformation, B11 and B12 . In a displacement controlled test,
one can vary them to characterise completely the macroscopic strain energy function M .
Indeed, if M were dependent also on I2 or I5 , additional unknowns would have appeared
in (3.30)-(3.32) causing them to be not identifiable uniquely in plane strain.
The following plane strain deformation is here considered,

1 0
1 0
0 0

FM = 0 1 0 = 0 1 0 0 1 0 ,
0
0
1
0 0 1
0
0 1

(3.36)

which is a uniaxial extension, , accompanied by a simple shear .


Given (3.36), it is easy to calculate the invariants
I1 = 2 + 2 + 2 + 1
I4 = ( sin (0 ) + cos (0 ))2 + 2 sin2 (0 ) .

(3.37)

The direction normal to the platelet in the reference configuration is N = ( sin(0 ), cos(0 ), 0)T
1 T
(see Fig. 3.4), whereas the current normal vector is n = FT
M N FM N; in view of (3.25)

72

Numerical Modelling

the current orientation of the platelet in terms of the original angle 0 and the deformation
components and is given by,



1 cos(0 )
1
= tan
.
2
cos(0 ) sin(0 )

(3.38)

eM = M /(2) in terms of the


The expressions of the non-dimensional Cauchy stress,
deformation components , and the orientation of the platelet 0 , can then be obtained
from (3.21),

2
e11
e33 = 2 1 + sin (0 ) + 2 cos (0 )

(3.39)

1 + 2
1 + sin2 (0 )
2

(3.40)

e22
e33 =


e12 = + sin (0 ) sin (0 ) + 2 cos (0 )

(3.41)

where the auxiliary function, , is defined by = 2 / , with



3/2
= ( 2 + 1) sin2 (0 ) + 2 sin(20 ) + 4 cos2 (0 )

(3.42)

In section 3.5.4, the cauchy stresses (3.39)-(3.41) are compared with the FE simulations
of the RVE.
To assure the compatibility with small strain linear elasticity it is useful to recover the
expression of the effective shear modulus c from (3.41). This can be done by assuming
that the the RVE is deformed in simple shear with no overimposed stretch, i.e., = 1 and
= 0, in which case c is given by c = 12 / for 0, that is,


3
c (0 ) = 1 + (1 cos(40 )) .
8

(3.43)

and as expected, depends on the fibre orientation. In particular c has a maximum for
0 = 45 in which case the platelet is aligned along the principal direction of the deformation
and is either stretched or compressed; as such it provides the maximum resistance to external
forces that in turn results in the maximum effective shear modulus. On the contrary for
0 = 0 or 0 = 90 , the effective modulus is minimum and equal to .
These two limit values can be used to identify the constitutive parameters and from
the numerical results of the small strain analysis. Through Eq. 3.43 these are given by,

= c (0 ),

4
=
3


c (45 )
1 .
c (0 )

(3.44)

73

3.2 Constitutive Modelling

which show that, if a composite is described by the strain energy function (3.22), full characterization of the nonlinear response requires only two small strain tests: a shear test with
the fibres oriented at 45 and another one with the fibres at 0 . As a result, and are
uniquely determined.
Plane Stress
For the plane stress deformation in the E1 , E2 plane of a Cartesian coordinate system, the
macroscopic deformation gradient has the following form,

F11 F12 0

FM = F21 F22 0 ,
0
0 F33

(3.45)

or equivalently in terms of the left-Cauchy Green strain tensor,

B11 B12 0

BM = B12 B22 0 ,
0
0 B33

(3.46)

Since the material is incompressible, the following relationship between the strain components holds,
B33 (B11 B22 B212 ) = 1 ,

(3.47)

that shows that only two components of the deformation can be controlled independently
when an incompressible material is constrained to a plane stress deformation (As B11 , B22
and B33 are coupled).
If 0 is the orientation of the axis of transverse isotropy in the reference configuration, i.e., the fibre orientation, (see Fig.3.4), then A = (cos(0 ), sin(0 ), 0)T while its spatial
counterpart is a = (a1 , a2 , 0)T , where
a1 = F11 cos(0 ) + F12 sin(0 ) ,
a2 = F21 cos(0 ) + F22 sin(0 ) .

(3.48)

In view of Eqs. (3.45) and (3.48), I1 and I4 are given by,


I1 = B11 + B22 + B33
I4 = a21 + a22
and from (3.21) the components of the Cauchy stress immediately follows as,

(3.49)

74

Numerical Modelling

11 = p + 2 1 B11 + 2 4 a21

(3.50)

22 = p + 2 1 B22 + 2 4 a22

(3.51)

33 = p + 2 1 B33

(3.52)

12 = 2 1 B12 + 2 4 a1 a2

(3.53)

As expected, the only component independent of the pressure field is 12 , although the
plane stress assumption means that 33 = 0 and so,

p = 2 1 B33

(3.54)

Which leads to the following results

11 = 2 1 B11 + 2 4 a21 2 1 B33

(3.55)

22 = 2 1 B22 + 2 4 a22 2 1 B33

(3.56)

12 = 2 1 B12 + 2 4 a1 a2

(3.57)

33 = 0

(3.58)

The previous equations involve two unknown functions, 1 and 4 , and two independent components of the deformation, B11 and B12 . As with plane strain, in a displacement
controlled test, one can vary them to characterise completely the macroscopic strain energy function M . The chosen form of M , i.e., M (I1 , I4 ), represents the largest class of
transversally isotropic materials that can be fully identified through plane stress deformations, and in fact, if M were dependent also on I2 or I5 , additional unknowns would have
appeared in (3.55)-(3.58) causing them not to be uniquely identifiable in plane stress.
Considering a plane stress deformation given by,

1
0
0
2
0
1 0
1
1

FM = 0 2
0 ,
0 = 0 1 0 0 2
1
1
0 0 1
0
0 2
0
0
2

(3.59)

which is a uniaxial extension accompanied by a simple shear , and given (3.59), it is

75

3.2 Constitutive Modelling


easy to calculate the invariants,
2
+ 2 + 1


2
I4 =
sin (0 ) + cos (0 ) + 1 sin2 (0 )

I1 =

(3.60)

eM = M /() in
that gives the following expression of the non-dimensional Cauchy stress,
terms of the deformation components , , , and the initial orientation of the fibre, 0 ,


sin (0 ) 2
1
e11 = + cos (0 ) +

1/2
1 + 2 1
e
22 =
+ 1 sin2 (0 )



sin (0 )
1/2
1/2
e12 = +
sin (0 ) cos (0 ) +
1/2
2

(3.61)
(3.62)
(3.63)

where the auxiliary function is defined as,


(1 + )2 (1 + )(2 + )
+

=
2 2
2

(1 + )2 (2 + ))
2 3

(3.64)

within which the parameter is,


sin (0 ) 2
sin2 (0 )
=
+ ( cos (0 ) +
)

1/2


1/2
(3.65)

In section 3.4.5, the cauchy stresses (3.61)-(3.63) are compared with the FE simulations
of the RVE, whilst in section 2.3, comparisons have been made to experimental results.
To assure the compatibility with small strain linear elasticity it is useful to recover the
expression of the effective shear modulus c from (3.63). This can be done by assuming
that the the RVE is deformed in simple shear with no overimposed stretch, i.e., = 1 and
= 0, in which case c is given by c = 12 / for 0, that is,


3
c (0 ) = 1 + (1 cos(40 )) .
8

(3.66)

which is the same form as for plane strain i.e. it depends on the fibre orientation, c has a
maximum for 0 = 45 and a minimum for 0 = 0 and 0 = 90 .

76

Numerical Modelling

Similarly, these two limiting values can be used to identify the constitutive parameters
and from the numerical results of the small strain analysis. Through Eq. 3.66 these are
given by

= c (0 ),

4
=
3


c (45 )
1 .
c (0 )

(3.67)

which show that, if a composite is described by the strain energy function (3.22), full characterization of the nonlinear response requires only two small strain tests: a shear test with
the fibres oriented at 45 and another one with the fibres at 0 . As a result, and are
uniquely determined.
Additionally, in order to make comparisons to the small strain elasticity of uniaxial
extension experimental results, it is useful to recover an expression for the effective elastic
modulus, E11 , from (3.61). This can be done by calculating the derivative of the stress, and
taking the value at which there is no imposed stretch or shear, i.e., = 0, in which case E11
is given by E11 = 11 / for 1, that is,
E11 (0 ) =

3
(16 + 5 + 8 cos (20 ) + 3 cos (40 ))
16

(3.68)

As expected the effective elastic modulus depends on the fibre orientation. In particular
E11 has a minimum for 0 = 65.9 ; in this configuration the fibre provides the least amount
of resistance to the uniaxial forces imposed upon it.
On the contrary for 0 = 0 or 0 = 90 , the effective modulus is a local maxima. The
maximum at 0 = 0 being significantly larger than at 0 = 90 . However, the limiting value
at 0 = 90 is interesting as it indicates the potential importance of the lateral constraint of
the reinforcement in an incompressible system, an observation that has been experimentally
observed and will be discussed in section 2.3.
Finally, it is useful to derive the expression of the fibre orientation in terms of the deformation components and . The direction normal to the fibre in the reference configuration is N = ( sin(0 ), cos(0 ), 0)T (see Fig. 3.4), whereas the current normal vector is
n = FMT N1 FMT N; in view of (3.45) the current orientation of the fibre, , in terms of
the original angle, 0 , and the deformation components and is given by:
"
= tan1

#
1/2 sin(0 )
.
cos(0 )

(3.69)

The previous equation allows comparisons to be made between the orientation of the
fibre as predicted. As it is shown in the next section, the mismatch between the numerical
results and the analytical model increases for increasing values of the stretch . This effect

3.2 Constitutive Modelling

77

is thought to cause the mismatch between the stress values obtained by ABAQUS and the
ones from (3.61)-(3.63).

78

Numerical Modelling

3.3
3.3.1

Numerical Homogenisation
Introduction

The ability of a hyperelastic model, such as the neo-Hookean or Mooney-Rivlin models,


to capture the behaviour of a material up to moderate strains (30-100% strain) has been
demonstrated previously. On the contrary, a transversally isotropic hyperelastic models
are usually employed to describe the behaviour of materials reinforced with continuous
fibres that span the entire length of the body and so it is important to assess the predicting
capabilities of the models against numerical results when using discontinuous fibres. The
intent of our analysis is to show the extent to which such a model accurately describes the
behaviour of both the RVE and experimental results.
In order to investigate the applicability of constitutive models, it is important to validate
the model with respect to a comprehensive set of experimental data. In light of the difficulty
to experimentally test anisotropic materials (See e.g. the experimental difficulties in section 2.3 and Ref.[256]), an alternative is to simulate the material behaviour in the context of
a representative volume element (RVE).
Three RVE configurations are considered in determining the behaviour of an elastomer
material reinforced by aligned fibres up to moderate strains (i.e. <60%): (1) A 3D, multiple
reinforcement RVE as a benchmark for the real material behaviour, (2) A 2D, multiple
reinforcement RVE, and (3) A 2D, single reinforcement RVE. The 2D RVEs are investigated
to determine their ability to describe the behaviour of the elastomer composite at a reduced
computational cost.
The constituents of the RVE (i.e. the matrix, and the reinforcement) are both modelled as
neo-Hookean materials with a heterogeneity contrast in the neo-Hookean parameter, C10 , i.e.
(p) / (r) . The justification for assuming the constituents behave as neo-Hookean materials
is based on the knowledge that:
An elastomer matrix (e.g. NR or PDMS) can be described as neo-Hookean up to
moderate strains (<100%) [165].
When the heterogeneity contrast between the constituents is large, the fibre is expected
to be approximated well by a neo-Hookean model.
The neo-Hookean model requires one parameter to be defined (i.e. C10 ), which simplifies any parametric study.
The addition of a moderate amount of compressible filler particles (<10%) to an
nearly-incompressible elastomer is not expected to increase the compressibility much

3.3 Numerical Homogenisation

79

[24].
It has been shown that a composite with a matrix reinforced by spherical fillers, both
neo-Hookean, can be successfully described by a neo-Hookean constitutive equation
[88].
It has further been shown that when the reinforcement is a fibre, the overall behaviour
can be described by a transversely isotropic constitutive equation [86].
Despite the research on transversely isotropic materials, including both constitutive and
numerical studies, there are still many unanswered questions that an RVE of the type discussed in the literature is unable to answer:
The models available in the literature only consider reinforcement configurations 0
and 90 , and therefore fail to consider the effects of the reinforcement angle and the
subsequent rotation of the reinforcement during deformation.
The effects of fibre angle distribution have been investigated in a constitutive context
[72, 187], however a constitutive model makes a number of assumptions (e.g. no fibre
interaction) that could change the behaviour. The effects of fibre angle distribution
could significantly change the behaviour of the composite.
It is difficult to assess the mechanical performance of transversely isotropic materials
in certain loading conditions (See e.g. [256]), hence there is a limited amount of
experimental data available from which to validate constitutive models. An RVE can
be used to assess the effects of different loading conditions that are difficult to assess
experimentally (e.g. The compression of a specimen, reinforced by obliquely aligned
reinforcements).
Furthermore, the extent to which a simplified single reinforcement RVE (Fig. 3.4) can
model the behaviour of an elastomer, reinforced by aligned particles, is unclear. The benefit
of such a model would be the simplification of a parametric study to investigate effects
of, e.g., the particle aspect ratio, filler vol% and constituent heterogeneity contrast. This
would also be performed at a greatly reduced computational cost. The model could also
be used as the basis for further investigation into the effects of the matrix/reinforcement
interface, failure criteria and an investigation of the large strain effects with more advanced
constitutive behaviour.

80

3.3.2

Numerical Modelling

Homogenization Procedure

The macroscopic behaviour of the composite can be determined from the mechanical response of a Representative Volume Element (RVE). The accurateness of the solution depends critically on the ratio between the size of the RVE and inclusion. In general, for reinforced composites, it can be assumed that the microstructure characteristics are such that
the dimensions of the filler are on a completely different length scale to the macroscopic
dimensions of the body.
In the following, we denote xm as the current position of a microscopic particle, and Xm
as the position in the reference configuration of the RVE, that is assumed to be stress free.
The subscript "M" refers to a macroscopic quantity, while the subscript "m" will denote a
microscopic quantity. The motion is a one to one mapping (Xm ,t), that assigns to each
point, Xm , belonging to the reference configuration the position xm at time t, i.e. xm =
(Xm ,t). As such, the deformation gradient Fm is defined through Fm = / Xm . If both
the constituents are hyperelastic, one can postulate the existence of a strain energy function
m that depends upon the deformation gradient Fm , the position Xm , i.e, m (Fm , Xm ), and
satisfies objectivity and material symmetry.
The homogenisation procedure allows previous microstructural quantities to be related
to the macroscopic response of the body in terms of average deformation and stress. These
are defined by,
1
1
F dV =
xm N dS,
FM =
|0 | 0
|0 | 0
Z
Z
1
1
PM =
P dV =
tm X dS,
|0 | 0
|0 | 0
Z

(3.70)

where 0 is the volume occupied by the RVE in the reference configuration, |0 | its meaR
sure, i.e., 0 dV , xm and tm represent the position and traction on the boundary of the RVE
and N denotes the outward normal to the boundary of the RVE.
One can introduce the macroscopic strain energy function, M , that is in relation with
the average stress and deformation,
1
M (FM ) :=
|0 |

m (Fm ) dV = PM FM ,

(3.71)

also known as the Hill-Mandel condition. Here and henceforth indicates the inner product.
Considering the form of the strain energy given in (3.22), The previous equation implies
that for an incompressible material, i.e. (3.22), the first-Piola Kirchhoff stress is calculated

81

3.3 Numerical Homogenisation


as,
PM =

M (FM )
.
FM

(3.72)

The volume average of the macroscopic quantities does not hold for the Cauchy stress
(see, e.g., [138]); however once the first-Piola Kirchhoff stress is calculated the usual relationship between the stress measures means the macroscopic Cauchy stress could be defined
as,
M = p I +

M (FM ) T
FM .
FM

(3.73)

The problem of defining the constitutive response of the homogenised hyperelastic material is hence reduced through (3.72) or (3.73) to finding the macroscopic strain energy
function, M . Unfortunately, the derivation of an analytical expression is often challenging even for very simple microstructures due to geometrical and constitutive nonlinearities.
Therefore, to find the homogenised model one has to use the finite element solution obtained
by applying the proper average deformation gradient, FM , to the RVE.

3.3.3

Description of the RVE

In the following, we consider either a 3D or 2D RVE model, made transversely isotropic


by the addition of stiff discontinuous reinforcements. Either a plane strain or plane stress
deformation can be applied to the 2D RVE; in this way the reinforcement can be considered
as either a fibre reinforcement or platelet reinforcement, respectively.
Owing to the geometry of the specimens and reinforcements, it is possible to reduce the
full 3D model to a 2D one by enforcing a plane stress/strain deformation. As such, four
numerical models are considered in this study. To model the behaviour of an elastomer
block reinforced by platelets (See Fig. 3.5a): A 3D multiple platelet RVE (3D-MPRVE) is
reduced to a 2D single platelet RVE (2D-SPRVE) under the assumption of plane strain. To
model the behaviour of a thin fibre reinforced elastomer specimen, a 2D multiple fibre RVE
(2D-MFRVE) is adopted under a plane stress assumption (Fig. 3.5b). This model is reduced
to a 2D single fibre RVE (2D-SFRVE), again under the assumption of plane stress.
The reduction of the models to a single reinforcement model (i.e. 2D-SPRVE or 2DSFRVE) is likely to be valid under the hypothesis of dilute inclusions (low wt%) and that the
overall behaviour of both composites (i.e. fibre and platelet reinforced) can be inferred from
a single reinforcement model with significant advantages in terms of computational cost
and accuracy of the solution, however the degree to which this simplification is applicable

82

Numerical Modelling

is unclear.

Fig. 3.5 A schematic representation of the different models examined: from the full 3D
RVE to the single platelet one. In each case, the model is simplified further. (a) The platelet
reinforced elastomer block, (b) The fibre reinforced thin elastomer specimen.

Both the matrix and the filler are assumed to behave like incompressible neo-Hookean
materials with the strain energy functions given by,
()
(I1 m 3) p() (Jm 1) ,
(3.74)
2
where = {r, p} indicates the matrix and the filler, respectively. Here, , is the shear modulus of the constituents and I1 m = tr(Fm FTm ) and p() is the Lagrange multiplier associated
with the incompressibility constraint.
()

m (I1 m ) =

3.3 Numerical Homogenisation

83

For the numerical simulations, the shear modulus of the matrix was chosen as (r) =
2 MPa, and the heterogeneity contrast between the constituents, i.e. (p) / (r) , in the range
(p) / (r) = {1, 10, 100, 1000}; p() is determined from the boundary conditions.
Parametric simulations were run with the FE software ABAQUS (ver. 6.12) for different values of the constituents parameters and orientation of the reinforcement 0 [0, 90].
In particular, seven different orientations were considered for the 3D and 2D multiple reinforcements model (0 = 0 , 15 , 30 , 45 , 60 , 75 , 90 ). For the 2D single reinforcement
model, the reinforcement orientation was investigated at 5 intervals, between 0 and 90 ,
giving a total of 19 unit-cells.
The single reinforcement model is generated through a combination of ABAQUS input
file and python script, which allows the orientation of the platelet to be accurately controlled
and varied. In this case, the mesh was automatically generated through the Abaqus input
file. In contrast, the 3D multiple reinforcement RVE and 2D multiple reinforcement RVE are
meshed in the ABAQUS CAE environment. The multiple reinforcements are positioned and
oriented in a semi-random manner, by firstly assigning an orientation to each reinforcement
in correspondence to a prescribed distribution (eq. 3.75). The reinforcements are positioned
by initially considering a uniform distribution, and then applying a random displacement,
to the reinforcement (i.e. x < < x), where x is the distance between the reinforcements. The procedure is repeated until a configuration is achieved without reinforcements
overlapping each other, or the boundary of the RVE.
The number of elements was established through a convergence check and it was seen
that a good accuracy/convergence was reached with about 200,000 elements for the 3D
model, 26,000 elements for 2D multiple platelets and 4000 elements for the single reinforcement RVE. The *Hyperelastic, neo hooke model in ABAQUS was used for both
the matrix and the platelet with bulk modulus k = 104 , that makes the material nearly incompressible. For all simulations a *Static analysis with nonlinear geometry was carried
out.
The affine boundary conditions were enforced on the RVE through the Abaqus *DISP
subroutine. The average stress PM was computed from the resulting forces on the boundary
through (3.70). The effective incremental moduli are obtained by constructing the numerical
tangent through a linear perturbation analysis after each step of the nonlinear analysis.
For all simulations the ratio between the height and the length of the RVE is kept fixed
and equal to 1, i.e., a square RVE. The ratio t between length and thickness of the fibre is
t = 40, unless otherwise stated, and corresponds to a volume fraction of 1.2 vol%. The overall geometry and assumptions in the numerical model are comparable to the experimental
materials in section 2.3 and is also typical of the values used in nanocomposites [226].

84

3.4
3.4.1

Numerical Modelling

Results of the FE Analysis for Fibre Reinforced Elastomers


Introduction

The addition of anisotropically aligned reinforcements can significantly increase the properties of composites [52], with the directions of alignment being chosen to produce a tailored
response [40]. When fibre reinforcements have only one referred plane of orientation, the
composite can be described as transversally isotropic. At large strain, the effects of the
anisotropic reinforcement and rotation are expected to complicate the response of the material, however the behaviour of nickel-coated short carbon fibre reinforced PDMS has been
shown to be described well by a transversally isotropic hyperelastic incompressible model
up to approximately 30%. At strains greater than 30%, the model is unable to capture the
highly non-linear behaviour of the material.
The intent of our analysis is to develop a numerical model to assist in understanding the
behaviour of a transversally isotropic composite that is reinforced by dis-continuous fibres.
The models are usually employed to describe the behaviour of materials reinforced with
continuous fibres that span the entire length of the body, however, when the inclusions are
discontinuous and dispersed, one has to assess the predicting capabilities of the models.
When the reinforcements are oriented in a 3 dimensional arrangement it is expected that
this behaviour would require a full 3D analysis and a consideration of the angular and spatial distribution of the reinforcements. The computational cost of such a model is large,
and restricts the possibility to parametrically explore the design space of the composite.
However, when the fibres are restricted to an orientation within a single plane and the specimen thickness is very thin, as in our case, (Fig. 3.6), it is expected that the model can
be simplified to a 2D plane stress model. In fact, it is further proposed that the behaviour
may be described by a single reinforcement. The extent to which a 2D-SFRVE can model
the transversally isotropic material requires assessment, however its implementation would
greatly reduce the computational cost of the numerical analysis, reduce the variables in a
parametric study and allow a basis to consider other microstructural features, e.g. into the
effects of the matrix/reinforcement interface.
Never the less, it is known that the distribution of fibres, and the distribution of fibre
angles, can have a significant effect upon the behaviour of the composite. In order to investigate these effects, a 2D RVE consisting of 100 randomly distributed fibres (2D-MFRVE)
is created with the distribution of the fibre orientation carefully controlled to give a bespoke
distribution.

3.4 Results of the FE Analysis for Fibre Reinforced Elastomers

85

Fig. 3.6 A thin elastomer (length much greater than the thickness) with fibres restricted to a
single plane (E1 , E2 ) can be modelled as a 2D plane stress structure.

3.4.2

Numerical Methods

The numerical study consists of assessing the behaviour of a 2D-SFRVE (Fig. 3.7b) and
comparing it to the transversally isotropic constitutive model, Eq. 3.22. The model consists
of 4000 CPS4R elements, subject to plane stress deformation. A single reinforcement represents the behaviour of a dis-continuous fibre, which has an aspect ratio of 40, i.e. AR = L/t ,
a volume of 0.8% and a shear modulus of (r) = 2000 MPa, therefore giving a heterogeneity
contrast between the reinforcement and matrix of, (p) / (r) = 1000. The reinforcement is
aligned at 5 intervals between 0 = 0 and 0 = 90 . The RVE is compared to the transversally isotropic hyperelastic constitutive model up to 60% strain in tension and shear; in such
a way the RVE is fully characterised under the plane stress assumption for an incompressible
material.
To assess the ability of the 2D-SFRVE to model aligned reinforcements, a 2D-MFRVE
is developed (Fig. 3.7a); consisting of 100 aligned reinforcements of equivalent properties,
i.e. AR = 40, vol% = 0.8 and heterogeneity contrast, (p) / (r) = 1000. The reinforcements
are then aligned at the required angles, 0 . The number of elements required is increased
to 26,000, using the same plane stress elements, i.e. CPS4R. The reinforcement of the
platelets is considered at 15 intervals, at angles 0 = {0 , 15 , 30 , 45 , 60 , 75 , 90 }.
An additional parametric study is undertaken on the 2D-MFRVE, to investigate the effects of fibre angle distribution. The angle is varied in accordance with the distribution given
by,

86

Numerical Modelling

Fig. 3.7 The two RVE configurations in the undeformed configurations. (a) The multiple
aligned fibre RVE model for 0 = 30 (2D-MFRVE). The inset shows the mesh surrounding
a reinforcement. (b) The single fibre reinforced RVE model for 0 = 30 (2D-SFRVE). The
inset shows the mesh around the inclusion edge.

1
 ebcos(2 )
( ) = R 

bcos(2 ) d
0 e

(3.75)

By varying the parameter b at values b = {2, 5, 8} a set of MFRVEs with varying angle
distribution are created (See Fig. 3.8). The values are chosen based on the experimental
results of section 2.3, in which b = 4.5. The results of the five configurations, i.e. b = 2,
b = 5, b = 8, the perfectly aligned 2D-MFRVE and the 2D-SFRVE, are compared and the
ability of the hyperelastic model to capture the behaviour is assessed in each case, for each
angle, in uniaxial tension up to 60% strain. A total of 10 numerical specimens are created for
each MFRVE configuration, each with a randomly created distribution of fibre alignments
defined by Eq. 3.75, in order to quantify the resulting variability of fibre position and angle.

3.4.3

Results of the SFRVE FE Analysis

The reinforcement is expected to increase the overall stiffness of the RVE; whilst the transversely isotropic properties that result from the oriented fibres are expected to have significant effects upon the large strain behaviour due to the initial orientation of the fibre and its
subsequent rotation as the RVE is stretched. However, the dis-continuous morphology of the
reinforcement may result in discrepancies to the hyperelastic model due to its assumption
of a continuous, inherent, reinforcement.
The results of the simple tension tests are shown in Fig. 3.9 in the range [1, 1.6] for
different initial orientations of the reinforcement, 0 = {0 , 25 , 45 , 65 , 90 } and hetero-

87

3.4 Results of the FE Analysis for Fibre Reinforced Elastomers

Probability density

1.5
b=2
b=5
b=8
Model Fit

1
0.5
0
0.5
0
0.5
0
90

60

30

30

60

90

Reinforcement angle, 0

Fig. 3.8 The distribution of the fibre angles, as defined by the parameter b in Eq. 3.75.
The images show an arrangement whereby the fibres are aligned around the 0 = 0 configuration. The fibre distributions of the 2D-MFRVE are shown for b = 5 and the aligned
configuration.

0.5

Cauchy Stress, 11 (MP a)

Cauchy Stress, 11 (MP a)

geneity contrast, f = (p) / (r) = 1000 ( (r) = 2 MPa). The maximum stretch is chosen to
reflect the ability of a neo-Hookean material to model a rubber up to strains of at least 50%,
and in general are unlikely to be subjected to much larger deformations in most applications.

0
25
45
65
90

1.5

(a)
0.6

0.7

0.8

Stretch,

0.9

0
25
45
65
90

0
1

(b)
1.1

1.2

1.3

1.4

1.5

1.6

Stretch,

Fig. 3.9 Cauchy stress 11 against stretch for simple tension in terms of reinforcement
orientations 0 = {0 , 25 , 45 , 65 , 90 }. In all cases, (r) = 2 MPa, f = 1000, AR = 40
and vol = 1.2%.

The effects of the initial orientation of the fibre 0 can be seen from Fig. 3.9, in which
the 0 orientation has the higher stiffnesses, whereas the 65 RVE displays the lowest initial
stiffness, with the 90 RVE only marginally stiffer initially. For the 65 configuration the
stress applied on the boundary is transferred by the fibre predominantly by shear forces and
hence the lower tensile modulus. At larger tensile stretches, the 65 RVE becomes stiffer

88

Numerical Modelling

due to the reorientation of the fibres towards the loading direction (See Fig. 3.10a). Such a
rotation does not occur for the 90 RVE, as is seen in Fig. 3.10b, due to no shear stresses
being formed that rotate the fibre. The 0 RVE in compression has a much larger stress
response, partially due to the boundary effects of the specimen that is unable to rotate and
results in an instability at large strain; hence a similar instability is not seen in e.g. the 25
RVE.
This result illustrates the two distinct types of stiffening that the model shows at small
strain: longitudinal stiffening due to the stress transfer from the matrix to the fibre (most
prominent when the fibre is oriented parallel towards the loading axis i.e. 0 = 0 ), and
lateral stiffening due to the Poissons effect (when the fibre is oriented perpendicular to the
loading direction i.e. 0 = 90 ). At large strain the lateral stiffening effect is seemingly
reduced and the contribution of the longitudinal stiffening becomes more important to the
overall behaviour of the RVE. The fact that the stiffness at 0 = 65 is higher at larger tensile
stretch than 0 = 90 (shown clearly in Fig. 3.10a) shows that the longitudinal stiffening
effect is stronger at high stretches; and that the rotation of the fibre during stretching will
have a significant effect upon the large strain behaviour of the material.
The evolution of the fibre orientation in terms of the overall stretch, , has been plotted
in Fig. 3.10b against Eq. 3.69. When the fibres span the entire RVE, Eq. 3.69 gives the exact
expression of the current reinforcement angle; however in the present case, the movement
of the fibre is not constrained to the boundary that causes Eq. 3.69 and can not be assumed
equivalent. However, the error in comparing the predicted rotation and actual RVE rotation
is limited, with larger error for the 25 angle. Apparently, due to the geometry of the RVE,
at 0 and 90 no reorientation of the fibre takes place. On the contrary, when the angle, 0 >
and 0 < 90, the fibre aligns itself to the direction of the applied load, i.e., the 1-axis. For
instance, the fibre initially oriented at 65 reaches the orientation of 40 at = 1.6 that
corresponds to a rotation of 25 . Interestingly, the rotation of the fibre will add a non-linear
effect. For instance, for an initial fibre angle of 35 the stiffness contribution of the fibre
will increase with deformation due to the change in fibre angle. At an initial angle of 75 ,
the rotation of the fibre would be expected to decrease the stiffness of the RVE as it rotates
towards the softest response of 65.9 ; further rotation would then subsequently increase the
stiffness as it rotates towards the loading direction.
The RVE subjected to a simple shear deformation, i.e. Eq. 3.59 with = 1, shows a similar dependence upon the initial orientation and subsequent orientation. The corresponding
results are shown in Fig. 3.11. With these boundary conditions, the RVE has the stiffest response when the fibre is orientated at 45 ; in this configuration the reinforcement is aligned
along the principal direction of deformation and is either stretched or compressed and thus

89

Tangent Modulus (MPa)

20

Reinforcement Angle, ( )

3.4 Results of the FE Analysis for Fibre Reinforced Elastomers

0
25
45
65
90

15

10

5
1

(a)
1.1

1.2

1.3

1.4

1.5

1.6

0
25
45
65
90
Model

80
60
40
20

(b)
0
0.6

0.8

1.2

1.4

1.6

Stretch,

Stretch,

Fig. 3.10 (a) The tangent modulus during tensile loading of initial reinforcement angles
0 = {0 , 25 , 45 , 65 , 90 }. (b) The fibre change of orientation during simple tension
for different initial configuration of the reinforcement 0 = {0 , 25 , 45 , 65 , 90 } plotted against the predicted behaviour, Eq. 3.69. In all cases the parameters of the RVE are:
AR=40, vol%=1.2, (r) = 2 MPa and f = 1000.

Cauchy Stress, 12 (MP a)

Cauchy Stress, 12 (MP a)

provides the maximum resistance. In this latter configuration the platelet is compressed for
< 0 which in turn causes numerical problems, hence the results are only shown up to
= 0.42.
0
25
45
65
90

0.5

1.5
0.6

(a)
0.5

0.4

0.3

0.2

Shear Strain,

0.1

1.5

0
25
45
65
90

0.5

0
0

(b)
0.1

0.2

0.3

0.4

0.5

0.6

Shear Strain,

Fig. 3.11 Cauchy stress, 12 , against shear, , for simple tension in terms of reinforcement
orientations 0 = {0 , 25 , 45 , 65 , 90 }. In all cases, (r) = 2 MPa, f = 1000, AR = 40
and vol = 1.2%.

The effects of the fibre on the overall composite response are minimal for 0 = 0 due
to the fact that fibres parallel to E1 are unstretched under simple shear and behave approximately like the bulk matrix; therefore the thinner the platelet, the lower its influence on the

90

Numerical Modelling

composite response. At 90 the reinforcement effect is also minimal at small strain, due to
the fact that the height of the RVE does not change and the fibre initially rotates around its
position. However, as the shear strain increases the rotation of the fibre causes the RVE to
stiffen compared to the responses of the matrix behaviour and that of the 0 RVE. These
rotation effects can be seen by observing the tangent modulus, as shown in Fig. 3.12a. For
larger stretches, the tangent modulus at 90 becomes larger than 0 = 0 and for > 0.4
it even overcomes the 0 = 25 configuration. This effect can be justified by looking at
the evolution of the orientation in terms of shown in Fig. 3.12b. The platelet initially at
0 maintains its orientation and therefore has the lowest shear modulus. On the contrary,
the platelet initially at 90 rotates and aligns with the principal direction of deformation,
furnishing a stiffening response.

Tangent Modulus (MPa)

3
2.8

0
25
45
65
90

2.6
2.4
2.2
2
1.8
0

(a)
0.1

0.2

0.3

0.4

Shear Strain,

0.5

0.6

Reinforcement Angle, ( )

The effect of the re-orientation for the 45 RVE, is that the specimen becomes softer
than the 65 one for > 0.33, due to the 65 RVE rotating towards 45 whilst the 45 RVE
rotates towards the softer configuration (i.e. 0 ) and in fact has rotated by about 30 at 60%
strain. For all angles Figure 3.12b gives a close fit to the actual orientation of the platelet
computed through numerical simulations, with a slight deviation for the 45 specimen due
to the divergence of the solution.
0
25
45
65
90
Model

120
100
80
60
40
20

(b)

0
0.6

0.4

0.2

0.2

0.4

0.6

Shear Strain,

Fig. 3.12 (a) The tangent modulus during simple shear loading of initial reinforcement angles 0 = {0 , 25 , 45 , 65 , 90 }. (b) The fibre change of orientation during simple tension
for different initial configuration of the reinforcement 0 = {0 , 25 , 45 , 65 , 90 } plotted against the predicted behaviour, Eq. 3.69. In all cases the parameters of the RVE are:
AR=40, vol%=1.2, (r) = 2 MPa and f = 1000.

Complimentary to these results is the small strain behaviour shown in Fig. 3.13, demonstrating clearly the predominant longitudinal stiffening of the RVE at 0 < 60 and as well
as the complimentary lateral stiffening at 0 > 60 . The minima at 60 could be utilised

91

3.4 Results of the FE Analysis for Fibre Reinforced Elastomers

Tensile Modulus (MPa)

8.5

Abaqus Model
Theory

8
7.5
7
6.5
6
5.5

15

30

45

60

75

90

Reinforcement angle,

Fig. 3.13 The change in initial modulus during simple tension for different initial configuration of the reinforcement 0 = {0 90 }, (r) = 2 MPa and f = 1000 against Eq. 3.69

in the design of small strain applications, and could be further considered in respect to the
stiffening/softening effects achieved as the reinforcement orients at large strain. Interestingly, the theoretical hyperelastic model is fitted to the behaviour (i.e. Eq. 3.68) and shows
a minimum stiffness configuration of 65.9 as opposed to the mimimum of around 60 for
the RVE; this suggested the dis-continuous fibre reinforcement transfers the stress differently to that assumed by the inherent transverse isotropy of the hyperelastic model. The
RVE also demonstrates a more pronounced transversely isotropic effect, as shown by the
larger stiffness increases at 0 and 90 compared to the softest configuration. In particular,
Eq. 3.68 shows very little lateral stiffening effect.

3.4.4

Results of the MFRVE FE Analysis

A comparison between the results of the 2D-SFRVE and the 2D-MFRVE shows the very
close agreement between the RVEs and validate the choice of a single reinforcement to
model the reinforcing behaviour of a fibre reinforced hyperelastic material up to 60% strain
in uniaxial tension (Fig. 3.14). However, in order to observe the effects of distribution in the
fibre angles, it is required to consider multiple reinforcements.
Figures 3.15 (a-d) display the results of the MFRVE with varying distributions of the
fibre angle, and show that the overall behaviour remains consistent, i.e. there are no additional non-linear effects at the macroscale caused by the multiple reinforcements. It is
observed that, although the distribution has an effect upon the macro-response of the RVE,
the response is dependent upon the primary angle of the reinforcement. For instance, the
0 RVE shows an expected increased in the stiffness with increased alignment of the fibres

92

Cauchy Stress, 11 (MPa)

Numerical Modelling

0 - SFRVE
30 - SFRVE
60 - SFRVE
90 - SFRVE
0 - MFRVE
30 - MFRVE
60 - MFRVE
90 - MFRVE

4.5
4
3.5
3
2.5
2
1.5
1
0.5
0
1

1.1

1.2

1.3

1.4

1.5

1.6

Stretch ()

Fig. 3.14 A comparison between the uniaxial cauchy stress of the multiple and single fibre
2D plane stress RVE models up to 60% strain.

(The inset of Fig. 3.15a shows a stress increase of 0.3 MPa at = 1.6). However, the
60 RVE decreases in stiffness as the alignment of the fibre increases. A explanation can
be found by a consideration of Fig. 3.13 and the relationship of the reinforcement configuration to the modulus, i.e. Eq. 3.68; the 0 configuration is the stiffest response and so
any deviation of the reinforcement angle will soften the response. However, any deviation
of the softest configuration, 60 , results in an increase of the stiffness. Incidentally, the
30 configuration lies on a slope between the stiffest and softest configurations and so the
angle distribution has had a minimal effect upon the tensile response (0.075 MPa stress
difference at = 1.6).
The effect of the fibre angle distribution is also evident on the small strain response.
Fig. 3.16a indicates that, as the alignment increases, the transverse isotropy of the RVE also
increases; as illustrated by the increased difference at 0 and 90 when compared to the
softest RVE response. This softest configuration was found to be approximately 60 for the
SFRVE, however Fig. 3.16a suggests this may change as the distribution increases. This
is due to the stiffening effect that mis-aligned fibres would have in stiffening the response,
especially when they are aligned towards 0 , which is greatest when b = 2 (See Fig. 3.8).
The effects of the fibre alignment upon the individual small strain responses of each
angle can be seen in Fig. 3.16b, and show a prominent effect of the alignment on the 0
RVE compared to the other configurations due to the significant decrease in stiffening effect
offered by misaligned angles compared to a perfectly aligned fibre in this configurations
(See Fig. 3.13).

93

b=2
b=5
b=8
Aligned

= 0
0
1

1.1

1.2

1.3

1.4

1.5

1.6

3.8

ne
ig
Al

3.75

b=

ne

Al
ig

b=

b=

4.2

4.4

b=2
b=5
b=8
Aligned

= 30
0
1

1.1

1.2

b=2
b=5
b=8
Aligned

1
0.5

= 60

0
1

1.1

1.2

1.3

1.4

1.6

1.5

1.38

1.37

8
ig
ne
d

b=

1.5

1.36

Al

Al

1.5

1.5

(d)

1.39

b=

8
ig
ne
d

b=

b=

3.1

1.4

11 (MPa) @ 30% strain

2.5

3.2

(c)

b=

2.5

1.3

Stretch,
Cauchy Stress, 11 (MPa)

11 (MPa) @ 60% strain


3.3

b=

Cauchy Stress, 11 (MPa)

Stretch,
3

(b)

11 (MPa) @ 60% strain


3.85

b=

b=

(a)

Cauchy Stress, 11 (MPa)

11 (MPa) @ 60% strain

b=

Cauchy Stress, 11 (MPa)

3.4 Results of the FE Analysis for Fibre Reinforced Elastomers

b=2
b=5
b=8
Aligned

1
0.5

1.6

= 90
0
1

1.1

Stretch,

1.2

1.3

1.4

1.5

1.6

Stretch,

Initial Modulus (MPa)

8.5

b=2
b=5
b=8
Aligned
SFRVE

8
7.5
7
6.5
6

(a)
0

15

30

45

60

75

Reinforcement Angle, 0 ( )

Initial Modulus (MPa)

Fig. 3.15 The variation of the stress-strain curves with varying amounts of fibre distribution
for different angles. (a) 0 , (b) 30 , (c) 60 , (d) 90 .

=0
=30
=60
=90

6
Decreasing Alignment

(b)

90
b=2

b=5

b=8

Aligned

SFRVE

Fig. 3.16 The small strain effects of a distribution of fibres in the RVE. (a) The reduction in
transverse isotropy is shown by the reduction in the stiffness difference between longitudinal
and lateral directions, (b) Show the effect of increased alignment on the stiffness for each
angle.

94

Numerical Modelling

3.4.5

Discussion

Due to the geometry of the reinforcement, and the material properties of the constituents,
the composite is assumed to behave as a transversally isotropic material that is closely modelled by an incompressible hyperelastic transversally isotropic constitutive model, such as
Eq. 3.22. The model has shown it is effective is describing the behaviour of nickel-coated
short carbon fibres in a PDMS matrix up to about 30% strain (Section 2.3), however a
neo-Hookean type model can typically describe isotropic reinforced elastomers up to much
larger strains (<100%) [165], and even an extension of the current model to a Mooney-Rivlin
type transversally isotropic model has shown minimal improvement in the fitting [40].
The underlying assumption of these models is that the transverse isotropy is from continuous fibres that span the length of the composite. Hence, rather than consider more complicated constitutive models, it is deemed important to explore the mechanics of a transversally
isotropic composite when it is reinforced by discontinuous fibres, and test the validity of this
assumption.
The strain energy, Eq. 3.22, is linear with respect to both I1 3 and g(I4 ) and means
that a plane is formed within these axes which should be matched by the numerical RVE
results. To validate this hypothesis, the strain energy, , is compared to the one computed
by ABAQUS for the 2D-SFRVE: for all angles, Eq. 3.22 is fitted against the strain energy
data for simple shear and simple tension by using the Matlab lsqnonlin function (nonlinear
least square method). As a result, a set of values of the constitutive parameters and are
identified for each RVE (i.e. one set of parameters to describe all reinforcement angles for
an RVE of equivalent material properties).
The results for 0 = 25 , 45 , 90 are shown in Fig. 3.17. The coloured plane is Eq. 3.22,
that is linear with respect to both I1 3 and g(I4 ), which is defined in Eq. 3.23. The
ABAQUS results are shown by dotted red lines for simple tension/compression and by dotted magenta lines for simple shear. The accurateness of the fitting confirms that the overall
behaviour of the RVE can be modelled with the constitutive equation (3.22) and hence that
the assumption of transverse isotropy is acceptable for the considered platelet geometry.
Note that all the numerical points lie on the model plane; the same results were obtained for
all the angles.
The corresponding values of and were
= 2.018 MPa,

= 0.165,

To further assess the capabilities of the model, the fitting of the strain energy function is
shown in Fig. 3.18 in the vs. strain plane for different angles {0 , 45 , 75 }. The match

3.4 Results of the FE Analysis for Fibre Reinforced Elastomers

95

Fig. 3.17 The transverse isotropy of the model is shown by the close fitting to Eq. 3.22 for
different values of the heterogeneity contrast (a) f = 10, (b) f = 100 (c) and f = 1000.
The other values are 0 = 35 , (r) = 2 MPa, AR = 40 and = 0.8%. The planes show
the constitutive model fitted to the data. The plotted dots show the numerical values of the
strain energy function as computed by Abaqus for simple tension/compression (red dots)
and simple shear (magenta dots).

of the model with the data is excellent for both simple tension/compression and simple
shear. In Fig.3.18 (b), (d) and (f), the constitutive parameters and identified from the
fitting of the strain energy are used to predict through, Eqs. (3.61)-(3.63), the corresponding
stress-strain curves. For the 45 and 75 RVEs up to 60% strain, the model matches the
numerical results very closely for both uniaxial and shear loading conditions. The 0 RVE
matches the model very well in all loading conditions up to 20% strain, however in uniaxial
compression the model and the numerical curves start to diverge due to instabilities used
by the compression of the reinforcement. In all other loading conditions, the 0 RVE fits
the model remarkably well. Simiar instabilities are caused in uniaxial tension for a 90
RVE configuration, again due to the instability of the reinforcement under compression.
Similar compression of the reinforcement occurs for the 45 RVE in compressive shear,
however the geometry of the reinforcement does not have this confining effect; however
larger reinforcement geometries may expect this to be encountered.

96

Numerical Modelling

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 0
1

0.5

0
60

11 , 12 (MPa)

1.5

20

20

40

2
1
0
1

(a)
40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

(b)
2
60

60

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 45
1

0.5

0
60

20

20

40

1
0

(d)
40

20

20

40

60

Strain (%)

0.5

11 , 12 (MPa)

Strain Energy,

60

2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 75

40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

20

(c)
40

Strain (%)

11 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

0
60

(f)

(e)
40

20

20

Strain (%)

40

60

2
60

40

20

20

40

60

Strain (%)

Fig. 3.18 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 0 , (c)-(d) 0 = 45 and (e)-(f) 0 = 75 . The other values were
(r) = 2 MPa and f = 1000. The remaining angles can are found in the Appendix B.1

3.5 Results of the FE Analysis for Platelet Reinforced Elastomers

3.5
3.5.1

97

Results of the FE Analysis for Platelet Reinforced Elastomers


Introduction

When platelet-like reinforcements, such as Graphene, have uniform orientational distribution, the overall behaviour of the composite is isotropic and the corresponding stress-strain
relationship at the macroscale would follow the behaviour of the constituents [160]. However, when the anisotropic reinforcements are oriented in a preferred direction, it is reasonable to expect that their orientation has an effect on the overall mechanical response. When
the reinforcements are suspended within a 2D single plane, the behaviour can be described
as transversally isotropic.
The intent of our analysis is to show in what limit the large strain behaviour of the
overall composite can be accurately described as transversally isotropic. The models are
usually employed to describe the behaviour of materials reinforced with continuous fibres
that span the entire length of the body, however, when the inclusions are discontinuous and
reinforce in 2 directions, one has to assess the predicting capabilities of the models against
numerical results.
Attempts in the literature to describe the behaviour have focused on fibre reinforced
materials [44, 86]; even so, the models do not consider the effects of the reinforcement
rotation at large strain, which is expected to effect the behaviour and to depend on the
loading conditions. It may be expected that a consideration of this behaviour would require
a full 3D analysis and a consideration of the random distribution of the reinforcements (in
orientation and position). The computational cost of such a model is large, and restricts the
possibility to parametrically explore the design space of the composite. However, when the
platelets are restricted to an orientation within a single plane (Fig. 3.19), it is expected that
the model can be significantly simplified whilst still allowing much of the behaviour to the
explained and for a parametric study to be possible.
It has been shown that a greatly simplified 2D-SFRVE (single fibre RVE) can model
the behaviour of a fibre reinforced composite (Section 3.5). It is expected that a similar
model can be used to describe a reinforcement by platelet-like reinforcements, although
plane strain is applied to account for the platelets oriented in a block elastomer (Fig. 3.19).
The extent to which a 2D-SPRVE can model the transversally isotropic material requires
assessment, however its implementation would greatly reduce the computational cost of the
numerical analysis and would provide a basis for further investigation, e.g. into the effects
of the matrix/reinforcement interface, aspect ratio and vol%.
For small deformations, the mechanical behaviour of a neo-Hookean material is comparable to a material with linear elastic behaviour. Thus it is possible to consider an effective

98

Numerical Modelling

Fig. 3.19 Long Platelets oriented in a single plane (E1 , E2 ) of a block elastomer, wherein the
thickness is much greater than the length, can be modelled by a 2D plane strain assumption.

stiffness of the composite, and use it to estimate the strain energy for the shear deformation and extend the procedure to large deformations. This effective stiffness has been
explored extensively and various formulations are available, including the semi-analytical
Halpin-Tsai (H-T) equations that are found to fit experiments well (Eq. 3.76). However, the
2D-SPRVE allows a uniquely controlled environment that is not possible in experimental
testing; as such, comparisons can be made between the predictions of the SPRVE and the
H-T equations.
In [160] and [86] the Halpin-Tsai (H-T) equations [91] are used to predict the value of
in terms of the properties of the constituents. However, in those papers the reinforcements
were assumed to span the entire RVE (i.e. long fibres) and hence the morphology of the
reinforcement was not taken into account. When the filler has a platelet-like shape, the
following expression of the H-T equation can be used (See e.g., [10, 226]):

= (r)

2 f 2( + 1) + f
+ f (1 ) 2

(3.76)

where the additional parameter is introduced to take into account the aspect ratio of the
platelet. This equation is useful in predicting the properties of aligned reinforcement composites, although does not predict the effects on the transverse isotropy of the material, given
by . For what concerns the constitutive parameter , in [160] the following expression has

3.5 Results of the FE Analysis for Platelet Reinforced Elastomers

99

been derived,
=

(1 ) ( f 1)2
.
1 + ( f 1) + f

(3.77)

The application of such an equation would be useful in extending the usefulness of


the H-T equations to the full set of orientational configurations, i.e. 0 = 0 to 0 = 90 .
However its usability to dis-continuous reinforcing platelet-like reinforcements is unknown.
It is important for a semi-empirical model, such as the H-T equations, to validate the design
space considered; otherwise the predictive capabilities of the model are limited. As such, a
parametric study shows the fitting of the small strain response to Eqs. (3.76) and (3.77).

3.5.2

Numerical Methods

The numerical study consists of initially capturing the behaviour of a 3D MPRVE (multiple
platelet RVE) and comparing it to the transversally isotropic constitutive model, Eq. 3.22.
The model consists of 200,000 C3D8R elements, subject to plane strain deformation, and
contains 30 aligned and randomly distributed platelets of aspect ratio 40, i.e. AR = L/t , of
total volume 1.2% and shear modulus (r) = 2000 MPa, therefore giving a heterogeneity
contrast between the reinforcement and matrix of, (p) / (r) = 1000. The reinforcement of
the platelets is considered for the angles 0 = {0 , 30 , 60 , 90 }.

Fig. 3.20 The two RVE configurations in the undeformed configurations. (a) The 3D multiple aligned platelet RVE model for 0 = 30 (3D-MPRVE). The inset shows the mesh
surrounding a reinforcement. (b) The single platelet reinforced RVE model for 0 = 30
(2D-SPRVE). The inset shows the mesh around the inclusion edge.

The 2D-SPRVE is developed for comparison, and consists of a similar arrangement,


i.e. AR = 40, vol% = 1.2 and heterogeneity contrast, (p) / (r) = 1000. However the

100

Numerical Modelling

model consists of one reinforcement, oriented at a defined angle, 0 . The number of elements required is greatly reduced to 4000, using the plane strain elements CPE4RH. The
reinforcement of the platelets is considered at 5 intervals for angles between 0 = 0 and
0 = 90 .
In the case of both RVEs, the models are compared to the transversally isotropic hyperelastic constitutive model up to 60% strain in tension and shear; in such a way the RVE is
fully characterised under the plane strain assumption for an incompressible material.
An additional parametric study is undertaken on the 2D-SPRVE, to investigate the effects of filler volume, aspect ratio and heterogeneity contrast. The parameters are adjusted to
consider the volume in the range vol%= {0.4, 0.8, 1.2}, the aspect ratio in the range, AR=
{1, 10, 20, 40} and the heterogeneity contrast in the range (p) / (r) = {1, 10, 100, 1000}.
From this, comparisons are made to the Halpin-Tsai equations, and an alternative formulation is proposed that better fits the numerical data.

3.5.3

Results of the FE analysis

The inclusion of a stiff platelet reinforcement is expected to increase the overall stiffness of
the RVE; whilst the transversely isotropic nature of the material means that the results are
affected by the orientation of the platelet. For materials subject to large deformations, such
as rubber, this transversely isotropic behaviour is expected to have significant effects upon
the large strain behaviour due to the initial orientation of the platelet and its subsequent
rotation as the RVE is stretched.
Here we report the results of the large strain FE analysis carried out with ABAQUS.
All the results, unless otherwise stated, assume (r) = 2 MPa, heterogeneity contrast f =
(p) / (r) = 1000, aspect ratio of the platelet AR = 40 and volume fraction = 1.2%. The
maximum and minimum stretches were chosen by considering that composites are very
unlikely to be subjected to deformations larger than 50% for most practical applications.
The results of simple compression/tension tests for the 3D-MPRVE and the 2D-SPRVE
are shown in Figs. 3.21 (a)-(d) in the range [0.6, 1.6] for different initial orientations
of the reinforcement, 0 = {30 , 45 , 75 , 90 }. The results from the two numerical models
considered are remarkably close, although divergence for 0 = 45 in compressive shear
at large strain show boundary effects of the large single platelet. Similar discrepancies at
0 = 90 in tension are caused by convergence issues due to the compressive forces acting
on the platelet of the 2D-SPRVE; in those cases the results are only shown to = 1.3. In
shear, the RVEs generally show much closer agreement due to the reduction in converging
issues caused by the compressing of the reinforcement. The complete set of results for
0 = {0 , 15 , 30 , 45 , 60 , 75 , 90 } are available in Appendix B.3. The results suggest

101

3.5 Results of the FE Analysis for Platelet Reinforced Elastomers

2D-SPRVE Uniaxial
2D-SPRVE Shear
3D-MPRVE Uniaxial
3D-MPRVE Shear
Model Uniaxial
Model Shear

0 = 30

11 33 , 12 (MPa)

11 33 , 12 (MPa)

that the problem of a platelet-reinforced neo-Hookean matrix can be reduced to that of a


single-platelet plane strain numerical model, in certain circumstances.

2
60

(a)
40

20

20

40

60

2D-SPRVE Uniaxial
2D-SPRVE Shear
3D-MPRVE Uniaxial
3D-MPRVE Shear
Model Uniaxial
Model Shear

2
60

(b)
40

2D-SPRVE Uniaxial
2D-SPRVE Shear
3D-MPRVE Uniaxial
3D-MPRVE Shear
Model Uniaxial
Model Shear

0 = 75

2
60

(c)
40

20

20

Strain (%)

20

20

40

60

Strain (%)
11 33 , 12 (MPa)

11 33 , 12 (MPa)

Strain (%)
4

0 = 45

40

60

2D-SPRVE Uniaxial
2D-SPRVE Shear
3D-MPRVE Uniaxial
3D-MPRVE Shear
Model Uniaxial
Model Shear

0 = 90

2
60

(d)
40

20

20

40

60

Strain (%)

Fig. 3.21 Comparisons between the 3D-MPRVE and 2D-SPRVE for the cauchy stress difference (11 33 ) against strain for simple tension and compression, and the cauchy shear
stress (12 ) against the shear strain. Displayed for reinforcement orientations, (a) 30 , (b)
45 , (c) 75 , (d) 90 .

The results of simple compression/tension tests for the 2D-SPRVE are shown in Figs. 3.22
(a) & (b) in the range [0.6, 1.6] for different initial orientations of the reinforcement,
0 = {0 , 25 , 45 , 65 , 90 }. The effects of the initial orientation, 0 , are apparent from the
stress-strain curves and from the tangent modulus in Fig. 3.23a: 0 and 90 orientations confer to the RVE the higher stiffnesses whereas the 45 RVE has the initially softer response.
In this latter configuration the stress applied on the boundary of the RVE is transferred by
the platelet predominantly by shear forces and hence the tensile modulus at small strain is
very close to the one of the matrix, i.e., 6 MPa. These results are indeed in accordance with
the linear laminate theory that predicts identical stiffnesses at 0 and 90 (and minimum
stiffness at 45 ) (see, e.g., [90]). Interestingly, the two complementary angles 25 and 65

102

Numerical Modelling

have the same small strain longitudinal modulus but, as the stretch increases, the 65 one
becomes softer and by about = 1.22 is even softer than the 45 RVE.

0.5

0
25
45
65
90

0
25
45
65
90

11 33 (MP a)

11 33 (MP a)

3
2
1

(a)

1.5

(b)
0

0.6

0.7

0.8

Stretch,

0.9

1.1

1.2

1.3

1.4

1.5

1.6

Stretch,

Fig. 3.22 Cauchy stress difference against stretch for (a) simple compression and (b)
tension in terms of reinforcement orientations, 0 = {0 , 25 , 45 , 65 , 90 }.

These results illustrate the two distinct types of stiffening that the model shows: longitudinal stiffening due to the stress transfer from the matrix to the platelet (most prominent
when the platelet is oriented parallel towards the loading axis i.e. 0 = 0 ), and lateral stiffening due to the Poissons effect (when the platelet is oriented perpendicular to the loading
direction i.e. 0 = 90 ). It is seen that this lateral stiffening is partially driven by the geometry of the platelet in the 2D SPRVE, as the multiple platelet RVEs do not show such
prominent stiffening effects for 0 = 90 . The two stiffening effects also cause the divergence of the 25 and 65 configurations at large strain, due to the way the stress transfer
mechanism evolves. At 0 = 65 , the platelet rotates towards the softest configuration, i.e.
45 (See Fig 3.23b). However, the 25 specimen rotates towards the longitudinal reinforcing
direction.
For moderate strains the 90 in tension and the 0 RVEs in compression (2D SPRVE) are
actually much stiffer than the other configurations; this is due to the geometry of the problem
and by the geometrical constraint that the rigid platelet imposes on the movement of the RVE
edges. The stiffer response of 0 = 90 can be partially mitigated by the use of a smaller
reinforcement or multiple platelets that reduce the boundary effects of the reinforcement.
To further analyse these effects, the evolution of the platelet orientation in terms of the
overall stretch is shown for the 2D SPRVE in Fig. 3.23b against Eq. 3.38. When the
fibres span the entire RVE, Eq. 3.38 gives the exact expression of the current reinforcement
angle; however in the present case, the movement of the platelet is not constrained to the
boundary and hence Eq. 3.38 gives only approximate results. The error in terms of the

103

0
25
45
65
90

14
12
10
8

(a)
6

1.1

1.2

1.3

1.4

Stretch,

1.5

1.6

Reinforcement Angle, ( )

Tangent Modulus (MPa)

3.5 Results of the FE Analysis for Platelet Reinforced Elastomers

0
25
45
65
90
Model

80
60
40
20

(b)
0
0.6

0.8

1.2

1.4

1.6

Stretch,

Fig. 3.23 (a) Tensile tangent modulus against stretch in terms of reinforcement orientations, 0 = {0 , 25 , 45 , 65 , 90 }. (b) The platelet change of orientation for different initial
configurations of the reinforcement against Eq. 3.38. The absence of points for large tensile
and compressive strains both at 0 and 90 is caused by the premature failure of the FE
simulation.

actual orientation of the platelet is however limited, with larger errors occurring for the 25
and 65 angles at large strain. Apparently, due to the geometry of the RVE, at 0 and 90
no reorientation of the platelet takes place. On the contrary, when the angle, 0 > 0 and
0 < 90, the platelet aligns itself in the direction of the applied load, i.e., the 1-axis. For
instance, the platelet initially oriented at 65 reaches the orientation of 40 at = 1.6 that
corresponds to a rotation of 25 . Interestingly, during the deformation, the RVE assesses
the state of minimum stiffness that causes an effect in terms of the stress-strain curves, e.g.,
the tangent stiffness of the 65 RVE becomes softer than the 45 ones.
The same analysis was carried out for the RVE subjected to a simple shear deformation,
i.e., Eq. 3.36 with = 1; the corresponding results are shown in Figs. 3.24 (a) & (b). With
these boundary conditions, the RVE has the stiffest response when the platelet is orientated
at 45 (see Fig. 3.25a): in this configuration the reinforcement is aligned along the principal
direction of deformation and is either stretched or compressed providing the maximum resistance. In this latter configuration the platelet is compressed for < 0 which in turn causes
numerical problems, hence the results are only shown up to = 0.42. This is shown as a
divergence of the solution compared to the 3D-MPRVE (See Fig. 3.21 (b)), whilst the other
results show very close agreement.
The effects of the platelet on the overall composite response are minimal for 0 = 0 due
to the fact that fibres parallel to E1 are unstretched under simple shear; therefore the thinner
the platelet, the lower its influence on the composite response. At 90 the reinforcement

104

Numerical Modelling

Cauchy Stress, 12 (MP a)

Cauchy Stress, 12 (MP a)

effect is also minimal at small strain, due to the fact that the height of the RVE does not
change and the platelet initially rotates around its position. The effect of the re-orientation
that takes place at large strain, for the 45 RVE, is that the specimen becomes softer than
the 65 ones for > 0.33, due to the 0 = 65 configuration rotating towards 45 .
0
25
45
65
90

0.5

(a)

1.5
0.6

0.5

0.4

0.3

0.2

Shear Strain,

0.1

1.5

0
25
45
65
90

0.5

(b)
0
0

0.1

0.2

0.3

0.4

0.5

0.6

Shear Strain,

Fig. 3.24 Shear stress against (a) negative and (b) positive shear strain in terms of reinforcement orientations 0 = {0 , 25 , 45 , 65 , 90 }.

This rotation effect can be seen by observing the effective shear modulus, as shown
in Fig. 3.25a, from which it is seen that at both 0 and 90 the tangent modulus of the
overall composite is close to the shear modulus of the matrix, i.e., (r) = 2 MPa. For
larger stretches, the tangent modulus at 90 becomes larger than 0 = 0 and for > 0.4
it even overcomes the 0 = 25 configuration. This effect can be justified by looking at
the evolution of the orientation in terms of shown in Fig. 3.25b. The platelet initially at
0 maintains its orientation and therefore has the lowest shear modulus. On the contrary,
the platelet initially at 90 rotates and aligns with the principal direction of deformation,
furnishing an increasingly stiffer response. The 45 RVE rotates up to about 30 at 60%
strain. For all angles, Fig. 3.25b gives a close fit to the actual orientation of the platelet
computed through numerical simulations.

3.5.4

Discussion

Due to the geometry of the reinforcement, the overall composite response is assumed to
follow the transversely isotropic constitutive model (3.22) with the axis of isotropy given by
the orientation of the platelet. To validate this hypothesis, the strain energy, , is compared
to the one computed by ABAQUS for the 2D-SPRVE: for all angles, Eq. 3.22 is fitted against
the strain energy data for simple shear and simple tension by using the Matlab lsqnonlin

105

0
25
45
65
90

2.5

2
0

(a)
0.1

0.2

0.3

0.4

0.5

0.6

Reinforcement Angle, ( )

Tangent Modulus (MPa)

3.5 Results of the FE Analysis for Platelet Reinforced Elastomers


120

0
25
45
65
90
Model

100
80
60
40
20
0
0.6

(b)
0.4

0.2

0.2

0.4

0.6

Shear Strain,

Shear Strain,

Fig. 3.25 (a) Tangent shear modulus against shear strain in terms of reinforcement orientations, 0 = {0 , 25 , 45 , 65 , 90 }. (b) The platelet change of orientation during simple
shear for different initial configuration of the reinforcement against Eq. 3.38.

function (nonlinear least square method). As a result, a set of values of the constitutive
parameters and are identified for each RVE (i.e. one set of parameters to describe all
reinforcement angles for an RVE of equivalent material properties).
The results for 0 = 35 are shown in Fig. 3.26 for different values of the hetereogenity
contrast f = 10 (a), f = 100 (b) and f = 1000 (c). The coloured plane is Eq. 3.22, that is
linear with respect to both I1 3 and g(I4 ), which is defined in Eq. 3.23. The ABAQUS
results are shown by dotted red lines for simple tension/compression and by dotted magenta
lines for simple shear. The accurateness of the fitting confirms that the overall behaviour
of the 35 RVE can be modelled with the constitutive equation (3.22) and hence that the
assumption of transverse isotropy is acceptable for the considered platelet geometry. Note
that all the numerical points lie on the model plane; the same results were obtained for all
the angles.
The corresponding values of and were
10 = 2.019 MPa,
100 = 2.020 MPa,
1000 = 2.028 MPa,

10 = 0.08,
100 = 0.313,
1000 = 0.405,

(where the subscript refers to the heterogeneity mismatch) and indicates that the inclusion
rigidity has a predominant effect on the parameter rather than . Indeed, is a measure of
the transverse isotropy of the RVE with = 0 corresponding to the isotropic neo-Hookean
model; and its effects are reflected by the increasing slope with respect to the g(I4 ) axis of

106

Numerical Modelling

Fig. 3.26 The transverse isotropy of the model is shown by the close fitting to Eq. 3.22 for
different values of the heterogeneity contrast (a) f = 10, (b) f = 100 (c) and f = 1000.
The other values are 0 = 35 , (r) = 2 MPa, AR = 40 and = 1.2%. The planes show
the constitutive model fitted to the data. The plotted dots show the numerical values of the
strain energy function as computed by Abaqus for simple tension/compression (red dots)
and simple shear (magenta dots).

the plane in Fig. 3.26. On the contrary, the effect of the platelet stiffness on the equivalent
modulus, , is rather limited due to the low volume fraction considered, i.e., = 1.2%.
To further assess the capabilities of the model, the fitting of the strain energy function is
shown in Fig. 3.27 in the vs. strain plane for different angles {25 , 45 , 90 }. The match
of the model with the data is excellent for both simple tension/compression and simple
shear at both 25 and 45 . In Fig. 3.27 (b), (d) & (f), the constitutive parameters and
identified from the fitting of the strain energy are used to predict through, Eqs. (3.39)-(3.41),
the corresponding stress-strain curves. Up to 30% strain, the model matches the numerical
results with small errors for both uniaxial and shear loading conditions. When the strain is
larger, the model and the numerical curves start to diverge; a similar behaviour was observed
for the platelet orientation. In fact, at large strain the model does not accurately reproduce
the actual orientation of the platelet which in turn causes some error in the stress-strain
diagrams.
At 90 the RVE is much stiffer in tension than the one predicted by the constitutive
model. This feature was already observed at high strain (see Fig. 3.22) and it is caused
by the geometrical constraint that the rigid platelet imposes on the top and bottom bases;
a thought that is supported by the results of the multiple platelet RVEs. On the contrary,
simple shear does not have this confining effect and a closer match with the numerical
results is achieved.
Whilst the orientation of the platelet is important in determining the large strain results
of the RVE, the small strain response of the material also shows very clearly the effect of the
initial platelet orientation, 0 , on the effective shear modulus of the composite (Fig. 3.28)
and can help predict the large strain behaviour of the composite [87]. These results were

107

3.5 Results of the FE Analysis for Platelet Reinforced Elastomers


ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 25
1

0.5

0
60

(a)
40

20

20

40

11 33 , 12 (MPa)

1.5

3
2
1
0
1

(b)
2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 45
1

0.5

0
60

(c)
40

20

20

40

4
3

1
0
1

(d)

(e)
20

20

Strain (%)

40

60

11 33 , 12 (MPa)

Strain Energy,

0.5

40

60

40

20

20

40

60

Strain (%)

0
60

40

2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 90

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

Strain (%)
11 33 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

(f)
2
60

40

20

20

40

60

Strain (%)

Fig. 3.27 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 25 , (c)-(d) 0 = 45 and (e)-(f) 0 = 90 . The other values were
(r) = 2 MPa and f = 1000. The remaining angles can are found in the Appendix B.2

obtained by running a perturbative analysis on the undeformed configuration. For the considered boundary conditions and RVE geometry, the shear response of the platelet offers
the maximum resistance when oriented at 45 whereas at 0 and 90 it remains almost unstressed; this is mostly noticeable from the limited influence that the increase in the ratio
f has on the effective shear modulus at 0 = 0 with all the curves almost overlapping in
Fig. 3.28a (note that the case f = 1 corresponds to the homogeneous RVE). Note that, as
confirmed by Eq. 3.43 whose fitting is shown in the figure, angles reciprocal with respect to

108

Numerical Modelling

(a)

2.4

2.2

2
0

15

30

45

60

75

Reinforcement Angle, 0 ( )

90

AR = 1
AR = 20
AR = 40
Model

2.6

(b)

2.4

2.2

2
0

15

30

45

60

75

90

Reinforcement Angle, 0 ( )

Shear Modulus, c (MPa)

f=1
f = 10
f = 100
f = 1000
Model

2.6

Shear Modulus, c (MPa)

Shear Modulus, c (MPa)

45 give the same value of the composite shear modulus, e.g. c (30 ) = c (60 ).
vol = 0.4%
vol = 0.8%
vol = 1.2%
Model

2.6

(c)

2.4

2.2

2
0

15

30

45

60

75

90

Reinforcement Angle, 0 ( )

Fig. 3.28 Effective shear modulus c in terms of platelet orientations 0 for various heterogeneity contrast {1, 10, 100, 1000} against Eq. 3.43. The other parameters were (r) =
2 MPa, = 1.2% and AR = 40. Note that the case f = 1 corresponds to the homogeneous
solution.

The accurate fitting of Eq. 3.22 in Figs. 3.27 and the small strain fitting of Figs. 3.28
suggest that, by using Eq. 3.44, the values of c at 0 and 45 can be used to identify the
constitutive parameters and . These are shown in Fig. 3.29 in terms of the heterogeneity
contrast, f , and different values of the filler volume fraction; as expected from the addition
of stiffer and larger inclusions, both and monotonically increase for increasing f and .
As already seen in Fig. 3.26, the increase in f has a prominent effect on that passes from
being almost 0 when f = 10 (soft inclusion) to 0.42 when the platelet is much stiffer than the
matrix ( f = 1000). In the same figure the fitting of with Eq. 3.76 is shown. It is apparent
that this modified form of the H-T equation gives a remarkably good fitting being able to
assess the dependence either on f and on . The corresponding value of the volume fraction,
, for which this fitting was achieved was 1.2%. On the contrary Eq. 3.77 used in [160] and
in [89] does not provide the same accuracy of fitting for (see Fig. 3.29b) due to the lack of
an additional parameter to account for the aspect ratio of the platelet. Hence the following
alternative empirical expression is here considered, (3.78), which provides a much better
fitting of the numerical results that depends upon heterogeneity contrast, volume fraction
and the additional parameter which is optimised to = 4.93 for an aspect ratio of 40,
and = 2.27 for an aspect ratio of 20. The relationship shows the strong influence of
the parameter with respect to the aspect ratio and reinforcement volume. Additionally
it shows the logarithmic relationship of the heterogeneity contrast with the parameter ,
which is not captured by Eq. 3.77 and similar forms based on continuous fibres [86].
= (1 ) log( f );

(3.78)

109

3.5 Results of the FE Analysis for Platelet Reinforced Elastomers

1.5

0.25
vol = 0.4%
vol = 0.8%
vol = 1.2%
H-T

AR=20
0.2

vol = 0.4%
vol = 0.8%
vol = 1.2%
Eq.(3.79)
Eq.(3.78)

AR=20

0.15
1

/(r) 1 (%)

0.1
0.5
0.05

(a)
0 1
10

(b)
0 1
10

10

10

Heterogeneity Contrast, f

10

0.5
vol = 0.4%
vol = 0.8%
vol = 1.2%
H-T

AR=40
0.4

vol = 0.4%
vol = 0.8%
vol = 1.2%
Eq.(3.79)
Eq.(3.78)

AR=40

0.3

/(r) 1 (%)

10

Heterogeneity Contrast, f

1.5

0.2
0.5

0.1

(d)

(c)
0 1
10

10

Heterogeneity Contrast, f

10

0 1
10

10

10

Heterogeneity Contrast, f

Fig. 3.29 Dependence of the constitutive parameters (a) & (c) and (b) & (d) on the heterogeneity contrast, f , the aspect ratio, and the volume fraction against the model introduced
in [160] and the one introduced with Eqs. (3.76)-(3.78).

110

Numerical Modelling

3.6

Concluding remarks

An incompressible fibre/platelet reinforced neo-Hookean material model, subjected to large


plane stress/strain deformations, has been analysed.
The motivation for investigating the RVE models lies in understanding the non-linear
behaviour of elastomers at large strain, when reinforced with aligned fibres and platelets.
In particular, it has been shown that the ability of a constitutive model to describe the
experimental data of a fibre reinforced PDMS elastomer extends only to 30% strain (See
section 2.3), even though a neo-Hookean based transversally isotropic model would be expected to describe the behaviour up to higher strains (<60%), see e.g. [165]. The inability of the model to describe the behaviour in the strain range 30%-60% suggests that the
aligned dis-continuous reinforcements cause non-linearities that are not described well by
a transversally isotropic material model that assumes the reinforcement is continuous. As a
result, the constituents of the simulation are chosen to model those of the fibre reinforced
elastomer.
A study of the behaviour of the RVEs has been carried out and has shown that the response of the fibre and platelet reinforced materials can be adequately described by single
reinforcement RVE models, which work at a greatly reduced computational cost compared
to multiple reinforcement RVE models. Furthermore, in the case of the fibre reinforced
model, there is a close match between the numerically computed strain energy function and
the transversally isotropic neo-Hookean model up to 60%, whilst for the platelet reinforced
elastomer block, the transversally isotropic neo-Hookean model matches the behaviour to
40%; this slight reduction is thought to be due to the difficulty of the model to accurately
capture the rotation of the platelet, due to the lateral constraint. In all cases, care is required
to consider the instabilities that develop as a result of the compressive forces on the reinforcements in certain configurations (e.g. A 0 = 90 specimen in uniaxial compression).
This behaviour is not captured by the model, and inaccuracies occur; particularly at large
strain.
The advantage of a simplified RVE model, compared to the constitutive model, is that it
could be used to model many complex effects such as fibre interaction, matrix/fibre interfacial properties, and even material fracture behaviour.
Never-the-less, the considered transversally isotropic neo-Hookean model has some remarkable properties:
it is dependent on I1 and I4 only, and has the most general transversely isotropic form
that can be uniquely identified in plane stress. A dependency on an additional invari-

3.6 Concluding remarks

111

ant would require (e.g.) biaxial loading, which is not possible for an incompressible
material in plane strain.
the model has demonstrated that one value of each parameter, and , can simultaneously represent the properties of 19 angular configurations of the RVE, (i.e. at 5
intervals between 0 and 90), in all deformation gradients possible for the hyperelastic
incompressible material.
only two simple shear tests at small strain are actually necessary to identify the nonlinear constitutive parameters and : one with the platelets oriented at 0 and another
with 45 orientation.
The consequence of anisotropic reinforcements at short strain is evident in the stiffness
that is observed to depend strongly on the reinforcement orientation; although in the case
of a thin fibre reinforced elastomer transverse stiffening effect is not as significant, due to
the plane stress boundary conditions. In this case (plane stress), the constitutive model
considers the effect, although it is less pronounced. An investigation on the effects of fibre
mis-alignment at low wt% (when fibre interaction is reduced) has involved studying the
properties of a multiple fibre RVE (2D-MFRVE), which indicates that a reduction in fibre
alignment causes a reduction in these stiffening effects at 0 and 90 , reducing the transverse
isotropy. As a consequence of the misalignment, the angle of minimum stiffness is also
found to alter, in accordance with the fibre angle distribution.
A parametric study has been undertaken to investigate the effect of the heterogeneity
contrast and volume fraction of the platelet reinforced specimen. The results show that the
form of the Halpin-Tsai equation derived in the literature to describe [160] overestimates
the effect of the heterogeneity contrast and an alternative form of the Halpin-Tsai equation
is presented to illustrate the logarithmic relationship that the RVE predicts. As a result, the
values of the two parameters, and , can be estimated from the mechanical properties of
the constituents with the modified forms of the Halpin-Tsai equation introduced.
As a result of these analyses, it is suggested that for moderate strain up to 60%, fibre
reinforced thin composites are modelled as transversally isotropic solids. Whilst, platelet
reinforced elastomer blocks are modelled as transversally isotropic solids up to 40% strain.
Larger deformations are likely to require more advanced models that consider non-linear
material behaviours such as chain inextensibility, or a multiscale analysis to correctly match
the stress-strain behaviour of the composite and the interaction of reinforcements during
deformation.

Chapter 4
Magnetic Response
4.1

Introduction

The advantages of reinforcing networks (Discussed previously in section 2.1) are numerous, and ubiquitous in natural systems. This has led to significant research effort in the
orientation of anisotropic reinforcements in materials; partly due to the highly desirable
engineering materials they can create, which expand the design space, but also due to the
current difficulty in making such anisotropic and hierarchical structures without dramatically increasing the cost.
The remarkable properties of such materials are typically the result of biomimicry. For
example, consider the mineralised outer layer of many fish scales that create an aerodynamic
and protective coating [61] or the lamellar structure of seashells that arrest crack growth with
mechanical gradients [61]. Alternatively, see the microfibril angle of cellulose in spruce
trees, which varies to tailor the response of the material to the expected loading conditions
[59], and is individually tailored to reduce mechanical gradients and homogenise the strain
field to reduce stress concentrations [178].
Conventional composite materials attempt to utilise the advantages of anisotropic materials by using long stiff fibres in a softer matrix. The fibre geometry, volume and geometrical
arrangement can then be tailored to control the properties of the overall material, although
typically the fibres are only be oriented in a 2-dimensional plane and quasi-isotropic material properties are most common. Never-the-less, anisotropic laminates have demonstrated
advantageous properties in terms of, for example, aeroelastic tailoring [122] and are used
in helicopter blades to optimise flight performance [140]. Furthermore, relatively new advances are being made in the use of variable angle tow composite panels, that change the
orientation of fibres within a composite layer [125]; the advantages of which include the
potential for increasing the stability of plates by redistributing structural loads.

114

Magnetic Response

The majority of these applications need only to consider the small strain behaviour of a
planar arrangement of continuous fibres, i.e. conventional composites. However, elastomers
are large strain performing materials that would additionally allow the consideration of 3dimensional orientation of reinforcements; this expands the design space considerably for
an engineer, but also increases the associated manufacturing difficulties. Thus a thorough
understanding of the technical challenges is required.
In addition, in order to achieve spatial orientation it is often necessary to functionalize the reinforcements. For example, magnetic orientation requires the filler to develop a
magnetic torque. The result is that, once the elastomer has cured, the magnetic torque is
transferred to the bulk matrix. The material is now magnetically responsive, with mechanical and magnetic properties that are complimentary.

4.1.1

Manufacturing Tailored Reinforced Networks

Manufacturing Techniques
Shear Processing
The manufacture of bespoke reinforced materials in 3-dimensions is challenging and
many different techniques are available. Perhaps most prevalent is the orientation of fibres
by shear processes, often incidentally as an unavoidable consequence of extrusion or injection moulding, as demonstrated in Ref. [68], where short fibres are oriented towards the
flow direction. Nano-fillers can also be oriented by shear forces, GE [120] and CNTs [30]
have both provided enhanced thermal properties when oriented. Greater flexibility of the
orientation can be achieved by the use of a converging channel that orients the fibres in the
flow direction [119] or by additive layer manufacturing, which allows tracks to be draw
with the oriented fibre mixtures [233]. In this way, 3D constructions of oriented fibres can
be made.
More direct analogues to conventional composites can be seen in tape casting, where
a slurry of polymer is swept under a blade and shear-stresses orient the fillers [155]. The
resulting material is advantageous for its easy operation, stack-ability and therefore scalability. Similar layered nanocomposites can also be achieved by a similar layer-by-layer
technique [17]. More viscous melts can be oriented by the assistance of hot-pressing, which
orient them in a 2D plane [18].
Mechanical Deformation
Similar orientation of the reinforcements can be achieved by mechanical deformation of

4.1 Introduction

115

the final composite specimen. Wang et al. [248] demonstrated, in CNT reinforced epoxy,
that repeated stretching caused the CNTs to rotate in the direction of the applied deformation, whilst Jin et al. [114] used an elevated temperature in a thermoplastic composite to aid
the process.
Freeze casting
The orientation does not have to be achieved in-situ of the polymer matrix. Freeze casting can be implemented, which involves freezing a suspension of fillers in a medium to create a controlled structural order. This technique has demonstrated controlled suspension of
ceramic platelets [179]. The orientation of the fillers can be controlled by the ordered formation of ice crystals, and then the polymer matrix is later added to form the final anisotropic
composite.
Remote orientation
In contrast to many methods of filler orientation, particularly shear processing, there are
also methods that require no physical contact with the constituents once a homogeneous
dispersion is achieved. For example, the use of ultrasonic waves has been demonstrated to
create semi-oriented assemblies of fibres, positioned in discrete pathways along the ultrasonic wave lines [219]. Similarly, networks of filler can be created by the use of electrical
currents, which has shown positive results in successfully orienting CNTs in uncured epoxy
[167].
Magnetic orientation
However, the potential for magnetic orientation is increasingly being realised. It is a
non-contact method that gives orientational and spatial control of the reinforcement within
a matrix. The direction of the magnetic field determines the alignment direction of the
fillers, the magnetic field strength controls the degree of alignment and spatial control can
be obtained by controlling the homogeneity of the field. Moreover, this is all achievable in a
3-dimensional space without the use of electrodes or shear operations that could externally
affect the matrix material.
There have been many reports of magnetic orientation of filler particles within matrices, this can involve the formation of filler networks for applications such as, for instance,
magnetorheological dampers [151]. However, the benefits of oriented continuous fibres in
conventional composites has attracted research into the orientation of discontinuous fibres
[52, 114, 154, 232], driven by the exciting potential of an increased design space. By using a
static magnetic field, aided by an intermittent rotating magnetic field, it has been shown that

116

Magnetic Response

short carbon fibres can be oriented in a viscous fluid [133]. The incorporation of orienting
reinforcements into conventional manufacturing processes is attractive, as it would further
enhance the material properties. This has been attempted by Takahashi et al. [231], who
applied a 2.4T magnetic field during a hot-press operation of carbon fibres in polycarbonate and found increased alignment as a result. A demonstration of radial alignment with a
permanent magnet has also been conducted, in which carbon fibres were aligned in a 0.75T
field to create a bending actuator [131].
Recent interest in nano-dimensioned materials [14, 149, 262], has also been accompanied by attempts to magnetically align them. However, this often requires very high magnetic fields. For instance Kimura et al. [129] oriented CNTs in a polyester matrix under a
field of 10T, before polymerizing the dispersion after alignment. Stronger magnetic fields
of 25T have also been used to align CNTs in epoxy, in order to create filler networks that enhance the electrical conductivity [38]. Magnetic fields of up to 40T have also been tested to
investigate the effect of pulsed magnetic fields [232]; as pulsed fields provide an economical
method of producing a high field strength in a large space.
A solution to counter the requirements for large magnetic fields is to use magnetically
responsive fibres, such as Ni/Co coated CNTs, which have been aligned in polystyrene by
a magnetic field of 5T [222]. Lower magnetic fields of 0.3T have also achieved alignment
by coating CNTs with nanoparticles [126], which is a low enough magnetic field strength
to utilise permanent magnets [197]. The addition of superparamagnetic nanoparticles to
alumina platelets allows alignment in a rotating field of around 0.1T [154]. Alignment in
this latter case was aided by the combination of a magnetic field and mechanical vibration;
the vibration helps to overcome steric hindrance, particularly at high filler volume fractions.
The vibration or sonication of samples before alignment also has the effect of improving the
dispersion of the fillers [31, 134, 220], whilst its use during magnetic orientation has been
shown to improve the alignment of fillers [154].
The morphology of the reinforcement also has a large effect upon the alignment, for example it has been shown that curved CNTs do not align as easily as straight-pristine CNTs
[112]. However, by appropriate choice of the morphology of the filler (e.g. size, aspect
ratio, morphology and magnetic properties), alignment of fillers can be achieved with magnetic fields orders of magnitude lower than previously seen. Erb et al. [52] demonstrated
this, by theoretically predicting the geometry of filler that would give an ultra-high magnetic
response; the result was the alignment of alumina platelets coated with super-paramagnetic
nanoparticles in a magnetic field of just 0.08mT.

4.1 Introduction

117

Orientation schemes
Control of the matrix and filler properties, as well as the processing conditions, has been
shown to be important to the quality of the produced material; especially in its ability to
disperse and align the reinforcements. However, careful control of the magnetic field is
equally important. It can be used to locally and globally control the orientation and concentration of fillers in 3-dimensional space. For instance, magnetic alignment typically orients
the primary axis of fibres towards the direction of the magnetic field, however platelets
have 2 principle axes and so the application of a rotating magnetic field is able to bi-axially
align the platelets. The rotational speed of the magnets is important: at low frequency and
low viscosity, the platelets continuously rotate, but at high frequency and high viscosity the
platelet will bi-axially align. It has been shown that by considering the viscous, magnetic
and gravitational forces, a critical frequency for bi-axial alignment can be calculated [54].
A similar concept is to use a uniform field to orient one of the principal axes, and then use
a rotating magnetic field to orient the other [53].
By using a hydrogel with aluminium platelets coated with superparamagnetic nanoparticles, alternately oriented layers can be created [53]. The anisotropic swelling of each layer
then causes either bending or twisting, depending on the orientation of the platelets. Similar
actuating devices that directly use the magnetic responsiveness of the fillers have also been
produced by the radial orientation of fibres, which allow the specimen to bend under the influence of a magnetic field [131]. In fact, a similar arrangement has demonstrated its ability
to walk using the intermittent application of the magnetic field [132].
Spatial orientation is also a useful tool for engineering design for mechanical performance, for example Erb et al. used it to create vertically aligned platelets along a polyurethane
surface to increase the abrasion resistance, whilst platelets within the material were oriented
in-plane to increase the overall strength and toughness [52]. In addition to spatial orientation, the concentration of magnetic fillers can be controlled locally with the application of
a non-homogeneous magnetic field. This creates mechanical gradients, concentrating reinforcing materials to areas of particular interest, in a similar manner to natural systems such
as tree wood [52].

Measuring the degree of alignment


Knowledge of the dispersion and orientation of fibre reinforced composites can be critical
to understanding the underlying mechanical behaviour of a material. Indeed, many models
have sought to include the effects of orientation distribution on the small strain [64, 119]

118

Magnetic Response

and large strain behaviour [72, 187]. Hence, it is important that methods are available to
verify the state of orientation and distribution. For example, Kimura et al. [131] used a
cylindrical magnet to orient short fibres in a radial configuration in a composite strip. The
use of short carbon fibres of micro-scale dimensions meant that optical microscopy could
be used to verify the results.
In addition to the overall spatial orientation of the fibres, the distribution of the orientation can also be a useful quantity which can vary due to the number of different factors,
for instance the anisotropy of oriented chopped carbon fibres has been shown to be directly related to the strength of a magnetic field [134]. The collection of such information
can be very useful in understanding the material behaviour and in making comparisons to
constitutive/numerical models, however the collection of this data is often difficult and time
consuming. To improve this process, contrast techniques have been employed such as: etching the matrix with oxygen ions to roughen its surface in comparison to the fibres [174] and
the use of penetrative dyes that illuminate the glass fibres in an epoxy matrix [118]. These
methods make identification of the fibres from the matrix easier, and so automatic image
analysis can be undertaken with the appropriate software, however this is more challenging
when nano-scale fibres are employed.
Nano-dimensioned materials require the use of SEM to be able to view the reinforcements, however only small portions of the material can realistically be analysed and the
quality of the image depends on the conductivity of the consituent materials. Despite this,
Kim et al. [126] recorded the orientation of around 100 CNTs in epoxy after analsysis with
SEM in order to view the distribution. Alternatively, X-ray diffraction has been successfully used to indirectly assess the anisotropy of filler particles by taking diffraction patterns
of the specimens and recording the relative intensity, the intensity can then be related to
the orientation state of the material microstructure [4, 154]. This has been demonstrated
with CNTs in polycarbonate [231] and a thermoplastic composite [114], as well as platelet
shaped particles in epoxy [154].

Predicting the degree of alignment


The prediction of particle alignment in the presence of a magnetic field can be useful in
the design of materials with controlled architectures, and a number of methods have been
used in the literature [52, 134]. However, the accuracy of any solution depends on careful
consideration of a number of quantities such as: the filler geometry, orientation, aspect ratio,
magnetic susceptibility, matrix viscosity and the curing rate of the matrix. Even considering
all these factors, discrepancies can still occur due to a number of factors which are difficult

4.1 Introduction

119

to account for, such as movement of the matrix, magnetic coupling of fibres, mechanical
fibre interaction, matrix evaportation and matrix shrinking. Hence, experimental validation
is a vital tool in assessing the predictive capabilities of the models, although limited data is
available in practice due to the difficulty in recording the orientation of a sufficient quantity
of filler particles.
It is possible to construct a model by consideration of the thermal and magnetic energies
of a particle. At thermal equilibrium, a randomised orientation of particles is achieved
that can be given by Boltzmann statistics, however when the magnetic energy of the filler
is greater than the thermal energy, it is possible to achieve orientation if the geometry of
the filler produces an anisotropic magnetic response; the probability function of alignment
can then be calculated [134]. Erb et al. [52] also considered the gravitational energy, in
addition to a direct consideration of the geometry of the filler (1D fibre/2D platelet). The
addition of the gravitational energy takes into account that large particles will be dominated
by gravitational effects, but very small particles are dominated by thermal effects; in doing
so, it was shown that an optimum particle size exists at which high magnetic response can
be achieved. An accompanying distribution function, similar to that seen in Ref. [134],
determined the probability that 90% of the fibres, based on their geometry and magnetic
properties, would be aligned to within 15 degrees of complete alignment.
The previous studies considered a homogeneous steady magnetic field, however a rotating magnetic field has also been employed to biaxially align platelet-like reinforcements
[53, 54]. This has shown that at low frequency and low viscosity the platelets continuously
rotate, but as the frequency and viscosity are increased, the viscous torque begins to dominate and alignment of the platelets occurs. By consideration of the viscous, magnetic and
gravitational torques it has been determined that a critical frequency exists for a given viscosity at which alignment of the platelets occurs [54], and which has been used to bi-axially
align magnetised alumina platelets in an epoxy matrix [154].

120

4.1.2

Magnetic Response

Stimuli Responsive Elastomers

When the tailored magnetic reinforcements, discussed previously in their capacity to control
the mechanical properties of a material, are locked in position within the material, it is no
longer possible for them to rotate or translate in the presence of a magnetic field. However,
the force is now transferred to the matrix to cause deformation and locomotion of the material. This stimuli responsive behaviour of a material is ubiquitous in nature, particularly
in the form of biological muscles [238] and plants [168]. The range of actuation varies
hugely, from mechanical changes, to physical changes, to the rapid release of elastic energy
for propulsive purposes.
For example, differential swelling in pinecone bilayers causes them to open as they
dry. During rain, water is then collected by the pine and travels rapidly to the centre, is
transported to the inner scales and closes the pine again to prevent the seeds from spreading
in humid conditions [224]. Differential swelling is similarly adapted in spruce tree branches
by varying the cellulose fibril angles in the wood [60]. When the fibril is parallel to the long
axis of the branch, the wood is stiffest and restricts swelling; this results in a strong branch
that curves upwards and resists collapse.
However, these processes are generally relatively slow acting. In many applications
it is necessary to enact rapid or propulsive actuation, for example sphagnum moss is also
activated by dehydration which causes a rapid propulsion of spores into the air [168]. As
the moss drys, the cell walls of the capsule begin to collapse laterally, forming a cylindrical
shape with an increased internal pressure; at a critical level, the cap breaks free and the
spores are explosively released [168]. Similarly, the bunchberry bogwood flower has an
interesting pollen dispersal mechanism reminiscent of a catapult. Stimulated by the force of
a large insect or strong wind, the flower opens up in under 0.5ms and accelerates the pollen
at up to 2200g before quickly reaching terminal velocity [80, 168]; this gives it the best
chance of cross-pollination.
The venus fly trap is a well known example of an actuating plant that uses a bi-stablility
in the leaf to rapidly close and capture its prey [80]. The process consists of sensory and
actuating mechanisms: firstly the active sensory process is initiated by a loss of pressure
in the motor cells caused by the prey. This then triggers an elastic response in the form
of a switch in the stable configuration, trapping the prey by closing the leaf [266]. This
has inspired an electromagnetically actuated bi-stable composite plate that replicates this
change in stability under the action of a magnetic field [266]. Novel devices such as this
could potentially lead to solutions for engineering design in areas such as morphing wings,
hinge mechanisms and other deployable structures.
Indeed, the creation of new, exciting materials and structures is a challenging task in

4.1 Introduction

121

engineering, and has attracted considerable interest in recent years [228, 234]. The examples
found in nature concerning the actuation of materials to various stimuli have encouraged
many other novel designs that efficiently use multi-functionality to achieve shape change
through elegant design [26].
The actuation of structures in engineering is typically driven by mechanical forces, however a number of different stimuli have been demonstrated to cause actuation. For example,
Li et al. [152] demonstrated the temperature dependent actuation of gel blends into various
shapes by controlling the composition of bi-layers. The position of the temperature responsive material is varied within the structure to create junctions that act as shape-memory
hinges. This memory effect is also shown in a pH responsive elastomer [254], which slightly
changes its elastic properties in different pH environments. Here, the cellulose whisker reinforced polyurethane can be stretched in distilled water and held until dry. It will then
indefinitely hold its shape until immersion in water again.
These stimuli owe the shape of the deformed material to a combination of the undeformed structural architecture and the configuration of the material constituents, however
this behaviour can also be coupled to the directional properties of the stimuli, e.g. the electric field. Fehr et al. [57] showed the alignment of TiO2 particles in a PDMS gel, under
the action of an electric field. The resulting material is responsive to an electric field, and
can be made to bend into an undulating pattern by changing polarities of neighbouring electrodes [57] or elongate due to the electrostriction of a dielectric elastomer [257]. Similarly,
Varga et al. [242] showed the alignment of particles in different fields (i.e. electric and
magnetic), however they subsequently showed that there is a coupling between the field response and the mechanical properties. This depends on both the anisotropy of the material
and the direction of the applied field, and in a PDMS gel can additionally affect the swelling
behaviour. The maximum elastic modulus is found when the field and particles are aligned
[242].
However, an electric field requires physical contacts (i.e. electrodes) that are not ideal
in many applications. Infrared-triggered actuation offers an alternative by providing contactless stimulation. Liang et al. shows that thermally active polyurethane can be infrared
active with the addition of photoactive graphene reinforcements; in this case the graphene
increases the mechanical performance and means infrared light is absorbed by the graphene,
causing the polyurethane matrix to contract. This reversible actuation has also been applied
to bi-layer hinges [135]. By reinforcing one layer with CNT reinforced silicone, and leaving
the other as a passive silicone layer, the hinge can bend reversibly; the result is an inchworm
walker that can travel.
Many other types of stimuli have also been successfully demonstrated, such as thermal

122

Magnetic Response

[152], solvent [80] and visible light [26], and are generally characterised by their comparatively weak mechanical response on the structure, which means they are most suited to the
actuation of soft elastomers, for applications as diverse as soft actuating valves [236], artificial muscles [238], tuneable dampers [217] and micro-swimmers [71, 193]. Therefore, due
to the comparatively low elastic modulus of such materials, the effect of the reinforcements
on the mechanical properties is heightened, and magnitude increases in the mechanical properties are possible [13, 16]. Additionally, as shown by the work of Varga et al. [242] the
mechanical and stimuli properties can be coupled, therefore it is very important to consider
the configuration of the material constituents.
A lot of interest surrounds the potential of carefully controlling the alignment of the
constituents [52, 53, 134], particularly as the advantages are clearly demonstrated in nature
(e.g. the varying orientation of fibrils in tree branches along their lengths [228]). Erb et al.
[53] shows the various swelling behaviours of hydrogels, which create different twisting and
bending configurations dependent on the alignment and configurations of the reinforcing
platelets. The platelets are oriented by low magnetic fields, meaning the resulting hydrogel
is also responsive to a magnetic field.
This magnetic field response has been used to control actuation, and has shown a number
of advantages over other stimuli.
First and foremost is its ability to actuate remotely, i.e. without the need for electrode contacts, batteries or particular environmental conditions, which increases its
usefulness in biomedical applications, e.g. inside the human body.
In addition, the field can be controlled in strength, direction and homogeneity in 3dimensional space, creating a variety of responses.
This response can be reversible, repeatable and almost instantaneous. By turning an
electromagnet on and off, an oscillating response has been demonstrated up to 40Hz
[270], with no phase shift.
The mechanical and magnetic behaviour are coupled, and in fact the elastic modulus
can be increased substantially in the presence of an increasing magnetic field [20].
This potential for magnetic actuators has attracted a number of studies on the properties
of magneto-elastomers. Most prominent is the work of Zrnyi et al. concerning magnetic
polymer gels, in particular the dependence of elongation on the non-uniform magnetic field
intensity [230, 267269]. The application of a load to a gel causes it to elongate, however
the addition of spherical superparamagnetic particles to a gel reinforces the material and

4.1 Introduction

123

reduces this effect. Furthermore, the magnetic response of the material under a magnetic
field is to contract and subsequently further reduce the effects of the load [267], in a process
analogous to the motion of a muscle. This contraction is caused by the magnetic locomotion
of the superparamagnetic particles, which attract to each other in a field and increase the
mechanical properties.
For this reason, the mechanical-magnetic properties of homogeneous magneto-elastomers
have been investigated by a number of researchers (See, e.g. [20]), and have shown an increase in the elastic modulus [243] that is proportional to the field intensity and important
to consider in the derivation of theoretical models concerning actuation [268, 269]. The
specimen resists any imparted deformation, and has, for example, shown increases in the
compressive modulus when in a magnetic field [56]. This enhancement in the properties
is immediate in the presence of a magnetic field, and can show dynamic behaviour when
coupled to the actuating behaviour. For instance, Zrnyi et al. [270] showed the actuating
in-phase response up to a oscillating magnetic field of 40 Hz.
The actuation of anisotropic materials is also of interest, especially as the anisotropy can
be produced in a magnetic field prior to curing the specimen. In this case, particle-particle
interactions cause the spherical particles to form chain-like structures in the material that
provide anisotropic properties [213, 242]. In the case of anisotropic magneto-elastomers,
the coupling between the mechanical and magnetic properties is now further complicated;
firstly by the direction dependent behaviour of the mechanical behaviour, and secondly by
the behaviour in each direction, which is dependent on the direction of the applied field.
Varga et al. [242, 243] showed that maximum elastic moduli are obtained when the load,
field and reinforcements are all aligned parallel, although when the field and reinforcements
are parallel but perpendicular to the load, there is still a significant increase in elastic modulus due to the resulting lateral constraint. For homogeneous materials, the increases in
elastic properties are not as great; although not significantly dependent on the field direction. Other configurations of field, load and reinforcement alignment are possible and many
of these are considered in Ref. [244].
However, in general only perpendicular arrangements of magnetic field, mechanical
loading and reinforcement anisotropy are considered due to the complexity of other systems. Never-the-less, the benefits of such arrangements are known and constitutive models
to predict the behaviour are being developed [28, 29, 213]. Among the many effects are
differential stiffening, which can result in, for example, localised swelling of polyurethane
reinforced with magnetite coated alumina platelets [52]. Bending of polymer films is possible by the creation of bi-layers, with the bending dependent on the mechanical gradients,
reinforcement concentrations, stimuli strength and a number of other factors [80]. Alterna-

124

Magnetic Response

tively, anisotropic reinforcement can create interesting actuating behaviour: Kimura et al.
[131] radially oriented short carbon fibres with a circular permanent magnet. The resulting
material aligns its reinforcing fibres to the lines of a magnetic field, and creates a bending
actuation effect. This arrangement can be used to create an actuating and walking elastomer
[132] (Fig. 4.1a). Similar variation of the reinforcement angle is demonstrated by Kim et
al. [127] to create a crawling movement (Fig. 4.1c). The material is made of modules of
transversally isotropic material, each with a different angle of isotropy that creates a zigzag pattern in a magnetic field. This is a direct improvement on the work of Ramanujan et
al. [202], who also use a modular arrangement of ferrogels to create a finger-like actuating
prototype (Fig. 4.1b); although the modules are magnetically homogeneous and limit the
motion.

Fig. 4.1 Examples of magnetic actuation and locomotion in the literature. (a) Inch-worm
like motion of a composite sample by varying the magnetic field on a ratchet strip [132],
(b) Modular, finger-like actuator responsive to a magnetic field [202], (c) A modular microactuator, in which each module has a different reinforcement direction and responds to a
magnetic field accordingly [127], (d) A soft cellular elastomer with embedded magnets that
collapses in a magnetic field [236].
Utilising this actuation to the locomotion of the material would be of interest to a number
of applications, including drug delivery and biological exploration, and has led to attempting
to find suitable actuation and locomotion mechanisms for this purpose. These are often

4.1 Introduction

125

termed as micro-swimmers and micro-actuators because of the small scale requirements of


the intended applications [193]. They can be optimised to a number of situations, making
use of the dimensions and morphology of the propulsive surfaces [71]. However, larger scale
applications can be envisaged. Larsen et al. [144] have developed a magneto-rheological
foam that can, under the action of a magnetic field, vary in shape and mechanical properties.
A novel material such as this could be of interest as an actuator or a damper. Tipton et al.
[236] showed that a cellular array of elastomer material (similar to a honey-comb material)
with embedded permanent magnets could deform and eventually buckle in a magnetic field,
and could have application as a valve or pump (Fig. 4.1d).

126

4.2
4.2.1

Magnetic Response

Magnetically tailoring the orientation of reinforcements


Introduction

The demand for highly innovative, multi-functional, and lightweight structures continues to
increase (see for example the book by Leng and Leu [147]). This demand can be realised,
partly through the use of self-assembly of reinforcements to produce bespoke material architecture.
The self-assembly of fibres has been attempted by many methods, including: shear
[223], electrical [211], ultrasonic [219] and magnetic alignment [131, 134]. In particular,
magnetic alignment has gained considerable interest due a number of advantages compared
to other techniques [207] such as: (1) Magnetic forces being contactless volume forces that
remotely orient magnetically anisotropic materials along field lines into 3D spatial configurations [52], (2) Permanent magnets and electromagnets being readily available and producing strong enough fields to orient the reinforcements [52], (3) Compared to electric fields,
magnetic fields do not produce currents and are not sensitive to surface charge and pH [207].
Reinforcements that are aligned to magnetic field lines, can in turn, offer control of
the material properties. The performance, however, depends critically upon the degree of
orientation; which is in turn dependent not only on the applied magnetic field strength but
also on the size, shape, volume [52] and magnetic properties of reinforcing filler used, as
well as the curing rate and viscosity of the matrix [7].
Therefore, accurate manufacturing requires an understanding of the material system and
its relationship to the experimental set-up. This in turn allows for the prediction of the fibre
dispersion [41] and helps to further understand the limiting factors of the magnetic field in
producing a bespoke material property.
Typically the focus involves the analysis of homogeneous fields [130, 131, 154] as this
results in a homogeneous dispersion of oriented fibres; due to a torque being applied to
rotate the fibres, but no translation of the fibres taking place. However, it is also possible to
expand the design space by considering the use of non-homogeneous fields. Such controlled
motion and transport of fibres would allow similar structures as those demonstrated in, for
example, the non-uniform dispersion of cellulose fibrils in tree structures [172].
In this section, we study the alignment of magnetic nickel-coated carbon fibres suspended in a viscous PDMS pre-cure solution, whose viscosity is increasing with time, under
the action of a homogeneous field. The distribution of fibres is investigated under an electromagnet by varying the magnetic field strength and analysing the fibre distribution in-situ
during the curing process. Comparisons are made to a model developed in Ref. [41], which
considers the effects of a viscosity increase during the cure process and is expected to bet-

4.2 Magnetically tailoring the orientation of reinforcements

127

ter represent the behaviour of real-life materials, in comparison to other models [52, 134],
when the fibres are oriented during the curing process [225].
Furthermore, neodymium permanent magnets are arranged in different configurations
to investigate some of the potential orientational effects that could be produced with such a
method. The configurations include an approximately homogeneous field, through thickness
reinforcement, variable angle reinforcement and a non-homogeneous field. It is anticipated
that an appreciation of the possibilities of magnetic orientation will drive further innovation
into novel engineering materials through self-assembly, which is becoming an increasingly
important area of research as the need for highly optimised structures continues to develop.

4.2.2

Materials characterization and experimental techniques

Materials Characterization and Preparation


PDMS (Sylgard 184) is used, without modification, as the matrix material. The reinforcement is provided by nickel coated carbon fibres of length 250 m and diameter 4.8 m
(aspect ratio 52, Ni/C 40/60 wt% purchased from Marktek Inc. and used as received). The
dimensions of the fibres were chosen for optical observation of the alignment, in-situ and
post-cure, whilst considering the difficulty to orient longer fibres in viscous solutions. The
PDMS fibre reinforced elastomers are made by direct mixing of the two compounds; this
involves the careful addition of a weighed sample of NiC fibres to a petri dish consisting of
a known amount of PDMS resin. Firstly, the mixture is agitated to partially homogenise the
mixture, then sonicated for 1 hour to further separate agglomerated fibres (agglomerations
of fibres can be seen in the fibre compound before processing, Fig. 4.2b). The mixture is
then left to settle for 24 hours to allow larger particles to settle to the bottom; this ensures a homogeneous material upon curing. This supernatant (1) is then separated and the
curing agent is added at a ratio of 1:10; the solution is then added to petri dishes and allowed to cure under the magnetic field. The remaining PDMS residue and NiC sediment is
then cured, burned off at 500 C and weighed to determine the amount of unused NiC fibre
material; this allows the fibre concentration of the PDMS specimens to be determined.
PDMS has been chosen as the matrix material for a number of reasons beyond its wide
use in both industry and research, it has many advantages which make it an attractive material when compared to NR:
It does not undergo crystallisation at large strain (i.e. Strain-induced Crystallisation).
1 The

term supernatant is here used to describe the homogeneous liquid suspended above the sediment.

128

Magnetic Response
It is more viscous in suspension, although the viscosity can be altered by the addition
of solvents.
No evaporation occurs during the cure process, resulting in less shrinkage.
It is transparent in the visible light spectrum, increasing the amount of information
viewable under a microscope.
The rate of cure, network density and stiffness of the material can all be tailored.

In addition, it can undergo isothermal room temperature curing, which will make identification of a magnetic constitutive model easier.
The reinforcements were chosen based on the requirement for a macro-sized magnetically responsive anisotropic filler, allowing it to be easily viewable under an optical microscope; this means the fibre length can easily be calculated, whilst allowing the fibre position
and orientation to be tracked more easily. In turn this allows the effects on the magnetic orientation speed, fibre dispersion and fibre orientation to be assessed in relation to the applied
magnetic field direction, field intensity, field homogeneity and fibre size.
The fibres themselves are not perfect and SEM images of the specimen indicates that
whilst the majority of fibres are in relatively good condition (see Fig. 4.2a), many features
of geometrical irregularities and damage can be observed (Fig. 4.2b - 4.2f). The simple
nature of the process is intended to reduce the damage to the fibres and fibre coating whilst
ensuring the quality of the final specimens. However, some of the fibres are already damaged
before specimen processing (i.e. as received), resulting in a distribution of fibre lengths that
is shown in Fig. 4.3.
Experimental Techniques
All testing was performed on the same batch of pre-cure reinforced PDMS solution at room
temperature. The resulting specimens were analysed by a Carl Zeiss Jenavert optical microscope. The magnetic field was applied by either an electromagnet (Newport Instruments
Electromagnet Type C, 4900 turns on each coil) or a configuration of N52 neodymium
magnets. The approximate concentration of all tested specimens was calculated as 2.1 wt%,
unless otherwise stated.
The initial testing procedure involved analysing the dispersion of fibres under the effect
of the electromagnet: firstly by observing the change of orientation, subject to a 0.08T
magnetic field, and recording the evolution of the fibre distribution at time intervals of,
t = {0, 60, 160, 220, 340, 3600} minutes. Secondly, the fibre distribution after 24 hours was

4.2 Magnetically tailoring the orientation of reinforcements

129

Fig. 4.2 NiC fibre characterisation under SEM. (a) good fibre (b) agglomeration of fibres
(pre-cure) (c) uneven nickel coating (d) Broken fibre (e) broken interface (g) almost fully
broken interface of small fibre.

Fig. 4.3 (a) The fibre length distribution, fitted to Eq. 2.4 (c=23.0, d=1.59). The averaged
values of 6 specimens are c=25.7 4.9 and d=1.61 0.15 (b) Micrograph image of the
aligned fibres, showing the variation in length and orientation (Selected fibres highlighted
in white).

130

Magnetic Response

measured for separate specimens, having been subject to magnetic field strengths of, |B| =
{0.01, 0.03, 0.05, 0.08} T.
A second series involves investigating the potential for curing specimens between magnets for 24 hours. Three configurations are constructed, (1) A homogeneous field between
four neodymium magnets, (2) An out-of-plane reinforcement with fibres reinforcing the
through thickness direction as well as the in-plane direction, (3) A variable angle, in-plane,
configuration to demonstrate the potential to spatially vary the reinforcement direction.

4.2.3

Magnetic Experimental Setup and Discussion

Electromagnetic Homogeneous Field


Fig. 4.4a shows the electromagnetic setup used to apply a homogeneous magnetic field to the
specimens, whereby the magnetic field strength is varied by controlling the applied current.
The homogeneity of the magnetic field ensures the fibres orient themselves parallel to the
magnetic field due to the magnetic torque, but do not translate. The magnetic
fieldstrength

between the electromagnet plates can then be calculated by N.I = B

Lcore
iron

Lgap
0

, where

N is the number of coil turns, I is the applied current, Liron is the length of the iron in the
circuit, Lgap is the air gap between the plates, iron is the magnetic permeability of iron and
0 is the magnetic permeability of free space.
The result is a transversely isotropic PDMS specimen; the transverse direction directed
between the electromagnet plates. However, the orientation of the fibres is not an instant
process and is highly dependent upon many factors, primarily the magnetic properties of the
fibres, the viscosity of the solution and the magnetic field. The effect of the magnetic field
strength is shown in Fig. 4.4b, and indicates the fibres are not completely oriented, even at
0.08 Tesla, most likely due to the observed damage on the fibres (See Fig. 4.2).
To characterise the fibre angle distribution and to aid in comparisons, Eq. 2.3 presented
in section 2.3 can be applied; as shown in Fig. 4.4b. Whilst this function is useful in characterising the resultant distribution of fibres, it provides no predictive capabilities. The
prediction of particle alignment in the presence of a magnetic field is a useful tool, as it
allows for a bespoke fibre alignment with optimal manufacturing efficiency. However, this
is not a trivial matter, and the accuracy of any solution depends on a number of parameters
such as: the filler geometry, orientation, aspect ratio, magnetic susceptibility and matrix
viscosity. A number of methods are available in the literature that take into account these
parameters [52, 134], however further discrepancies can occur due to a number of factors,
such as distribution in fibre lengths, varying magnetic responsiveness of fibres, movement
of the matrix, magnetic coupling of fibres, mechanical fibre interaction, matrix evaporation,

4.2 Magnetically tailoring the orientation of reinforcements

131

Fig. 4.4 (a) The curing PDMS specimen during the alignment procedure between the coils of
the Electromagnet (Newport Instruments Electromagnet Type C), (b) The fibre distribution
after 24 hours under various magnetic fields (|B0 | = {0.01, 0.03, 0.05, 0.08} T), fitted to the
fibre distribution function, Eq. 2.3 (b=0.44, 0.92, 1.58, 2.95).
matrix shrinking and viscosity increases during the cure process. These add to the complexity of the solution and mean that predictions are generally unable to satisfactorily capture
the evolution of the fibre alignment (e.g. see the fitting of the fibre angle distribution in
Ref.[134]).
However we introduce a model that considers the increasing viscosity of the solution
and allows the evolution of the fibre distribution, , to be calculated. The model considers
the motion of a magnetic fibre, embedded in a viscous fluid under a homogeneous magnetic
field. The fibre has a geometry given by the aspect ratio, AR, with the direction of the main
fibre axis given by the vector, n, subject to a magnetic field with direction, B. The fibre
has a cylindrical shape which can be approximated as a prolate spheroid with characteristic
half-lengths and r ( > r), aspect ratio AR = /r and volume = 4/3 r2 .
which, assumThe application of an external magnetic field induces a magnetisation M
ing a linear dependence of the magnetisation upon the field and an arbitrary angle of the
fibre with the direction of B, can be defined by,
= m 1B + a 1 (n n)B
M
0
0

(4.1)

where 0 is the magnetic permeability of free space and a = m m is the anisotropic


magnetic susceptibility, which is positive for a paramagnetic material. The coefficients m
and m denote the magnetic susceptibilities parallel and perpendicular to the main axis of
the fibre.
In addition, the total energy per unit volume of this fibre in a homogeneous field can be

132

Magnetic Response

given by [54, 132, 133],


1
= a 01 (n B)2 .
2

(4.2)

This fibre is acted upon by a magnetic torque, T , when the main axis of the fibre is
not aligned to the magnetic field direction. The resulting rotational hydrodynamic drag is
defined by d , which can be related to the rotational velocity around the secondary axis of
, by the following relationship,
the fibre,
B)
= d1 (I n n)T = d1 (I n n) (M

(4.3)

where T is the torque.


The scalar hydrodynamic drag quantity, d , can be calculated as quasi-static values of
Stokes flow around a spheroid particle (see for instance [92, 212]), i.e.,
d

8 3
.
3 [ln(2 AR) 1/2]

(4.4)

This shows a clear dependence upon the viscosity of the solution, , which is generally
assumed to be constant. However, if the fibres are aligned during the curing phase of the
composite, then the viscosity, and therefore the drag on the fibres, will be time-dependent.
A common empirical expression for a time-dependent viscosity is,
(t) = 0 et/c

(4.5)

where c is the characteristic time for the initial viscosity, 0 , of the solution to increase e
times, in isothermal conditions [194]. Such an expression has been used previously in the
literature [7, 50] and gives a good approximation of the cure cycle up to the gel-point (2)
[5, 210].
As a result, the time dependent normal vector of the fibre can be described in terms of
the magnetic quantities [258], i.e.,
dn
= o1 et/c (n b) [I n n]b
dt

(4.6)

where b := B/B0 is the magnetic field vector normalised to the reference value B0 , i.e.
the intensity of the homogeneous field. o defines the characteristic time constant, which
2 The

gel-point is the intermediate stage between a liquid and a solid, indicated by a rapid increase in
viscosity.

133

4.2 Magnetically tailoring the orientation of reinforcements


assuming a dilute solution of fibres and no fibre-fibre interaction, can be described by,
o =

0 d0
0
2 AR 0
=
,
2
2
a B0
a B0 ln (2 AR) 1/2

(4.7)

Considering that the fibres and magnetic field lie in the same 2D plane, with a magnetic
field that does not change direction, the governing equation can be described in dimensionless form by,
d
1
= e tr sin(2) ,
dtr
2

(4.8)

which is expressed in terms of , the relative angle between the fibres and the magnetic field.
Here, tr := t/o represents the normalised time, and is the ratio between the orientational
time constant, o , and the characteristic curing time, c , i.e. = o /c .
The solution of Eq. (4.8) in terms of the initial angle 0 is,


tan((tr )) = tan(0 ) e

1+e tr

(4.9)

which is a useful tool to predict the fibre distribution achieved after a time tr . In fact, the
total number of fibres, dN, oriented at time tr within an angle d, can be worked out in
terms of the distribution function (,tr ) as, dN = (,tr )d. Hence one has (,tr ) =
dN/d = (dN/d0 )(d0 /d), which allows the distribution to be obtained from (4.9).
By assuming a random orientation at time tr = 0, then the initial distribution dN/d0 is
constant and equal to 1/. By inverting Eq. (4.9) and taking the derivative with respect to
0 , one obtain the following distribution function,
1 d0
1
(,tr , ) :=
=
d

such that

R /2

/2

e
e

1+e tr

2e tr

sec()2

+e

(4.10)

tan()2

d = 1.

For such a distribution function, if the characteristic orientation time, o , is greater than
the characteristic curing time, c , i.e. < 1, the viscosity increases too quickly for the
fibres to fully align to the magnetic field lines and there is a resultant large spread in fibre
orientation. Alternatively, when c is much greater than o , i.e. 1, the fibres will align
quickly and the effect of the increasing viscosity, , is negligible. This effect, on the fibre
angle distribution, is shown in Fig. 4.5 for two values of .
From Eq. 4.10, it is possible to find the probability, p1 , of having a fibre with an orientation of 1 at time tr ,

134

Magnetic Response

(b)

0.4

Probability Density

Probability Density

(a)

0.3

5
2.5
Time, tr

0 90

45

45
Angle,

0
5

90
2.5
Time, tr

0 90

45

90

45
Angle,

Fig. 4.5 Orientation distribution of the fibre as predicted by the model (4.10) at different
times, tr , for (a) = 5 and (b) = 0.5. A random orientation of the fibres is assumed at
tr = 0.



1e tr
2

p :=
(,tr , ) d = atan tan( )e

(4.11)

which can be rearranged with respect to the reduced time, tr ,





tan( p /2)
1
tr = ln 1 ln

tan( )

(4.12)

This gives the time necessary to align a percentage of the fibres, p , to within 1 . It
depends upon the ratio = 0 /c , between the orientation time and the curing time of the
composite, which implies that a large value of means the composite cures too quickly for
orientation to take place.
By using the implications of this relationship for , it is possible to work out the minimum magnetic field that allows the orientation spread of 1 for p of fibres at cure, i.e.,

1/2
tan( p /2)
0 0
2 AR2
B0 min =
ln
.
c a ln(2 AR) 1/2
tan( )


(4.13)

This shows the evident dependence of the magnetic field intensity upon the curing rate
of the composite, i.e., 0 /c , as well as the magnetic and geometric properties of the fibres.
The contour plot of this function, illustrating this dependence, is shown in Fig. 4.6. The red
checked regions of the plots show where the magnetic field is higher than 1 T. Magnetic
fields up to 1 T can be generated in a normal lab environment through the proper assembly
of neodymium magnets [227], without using large and complex electromagnets. Therefore,

4.2 Magnetically tailoring the orientation of reinforcements

135

any configuration which falls within this checked region would likely require fields that
cannot be generated by standard magnets, but requires more complex equipment such as
superconducting magnets [131].

Fig. 4.6 Contour plots of the theoretical minimum magnetic field intensity B0 min (eq. 4.13)
to have 90% of the fibres oriented at 5 deg. B0 min is plotted against (a) the initial curing
rate 0 = 0 /c and the fibre aspect ratio AR (for a = 2.20 103 ), and (b) the initial
curing rate 0 = 0 /c and the fibre magnetic susceptibility a (for ar = 25). The labels
show the magnetic filed intensity in Tesla. The dashed red area represents the region where
B0 is higher than 1 T. Due to the increasing viscosity of the solution, the results are shown
for the steady state (i.e. cured) condition.
To validate the model and improve its predictive capabilities, experimental data has
been collected which records the fibre orientation and fibre length distributions over a 24
hour period subject to a magnetic field of 0.08T. A low wt% (i.e. 2.1 wt%) is used in order
to minimise the effects of magnetic fibre coupling, and similar interactions.
Whilst the model takes into account the effects of geometry, the experimental data indicates a strong dependence of the alignment on the aspect ratio. Smaller fibres are shown
to orient much less than larger fibres; with those fibres with AR greater than 24.1 almost
completely aligned parallel to the direction of the magnetic field (See Fig. 4.7).
This dependence is contrary to the predictions of the model (See Fig. 4.6a), which predicts that given the approximate fibre geometry and magnetic susceptibility, a decrease in
the AR would increase its ability to orient. This is due to the viscous forces on the fibres
decreasing more than any decrease in the magnetic torque on the fibres. However, Fig. 4.6b
suggests a decrease in the magnetic susceptibility of the fibres, which would correlate to

136

Magnetic Response
1.5

0<AR12.7

(a)

0.8
0.6
0.4
0.2
0
100

50

50

Probability Density

Probability Density

12.7<AR24.1
1

0.5

0
100

100

50

Angle, ( )

50

100

Angle, (o )

2.5

Probability Density

(b)

(c)

AR>24.1
2
1.5
1
0.5
0
100

50

50

100

Angle, (o )

Fig. 4.7 The effect of fibre damage and fibre morphology on the magnetic susceptibility.
Orientation of different fibre aspect ratios after 60 minutes under the homogeneous magnetic
field of an electromagnet at a field strength of 0.08T. Fitted to Eq. 2.3 (a) Orientation of fibres
of aspect ratio: {0 < AR 12.7}, b=0.7 (b) {12.7 < AR 24.1}, b=1.4 (c) {24.1 < AR}.
b=37.7.

the damage observed when analysing the nickel-coating of the fibres in Fig. 4.2d-f. These
images indicate that many of the fibres have broken or damaged coatings. Presumably, as
fibres are broken during processing, they also lose some magnetic coating and have a much
lower magnetic susceptibility, a , as a result. This corresponds to a larger value of the
characteristic orientation time, o .
This can be accounted for in the model by assuming a series of characteristic orientation
times, i.e. o1 , o2 & o3 . Each value of o is fitted to a distribution of fibre aspect ratios, and
in this way the variability of the magnetic susceptibility is hoped to be captured. The fitting
of this data can be seen in Fig. 4.8, where observations of the fibre orientations were made
over a 24 hour period.
The orientation of the fibres in-situ, under the influence of a 0.08T field, indicate that
a significant amount of reorientation occurs in the first 60 minutes (31% aligned to within
5 ) and that in 340 minutes the fibres are 56% aligned to within 5 (See Table 4.1). Only a

137

4.2 Magnetically tailoring the orientation of reinforcements

60 mins
160 mins
220 mins
340 mins
24 hours
Model (Eq 4.3)

(a)

0.10

Probability Density

Probability Density

0.15

0.05
0
0
0
0
0
90

60

30

30

60

90

0.10

60 mins
160 mins
220 mins
340 mins
24 hours
Model (Eq 4.3)
01 , 02 , 03

(b)

0.05
0
0
0
0
0
90

60

30

Angle,

30

60

90

Angle,

Fig. 4.8 Fibre distribution measured at intervals over a period of 24 hours, showing the
fitting of Eq. 4.10. (a) One value of is fitted to the interval data, o1 = 15.4. (b) Three
characteristic orientation times improve the fitting of the data, o1 = 20.2, o2 = 15.7, o3 =
12.6
small amount of further orientation occurs after this time due to the increasing viscosity of
the curing solution.
Table 4.1 The percentage of fibres aligned to the magnetic field direction in experimental
testing. The model fittings are shown in square brackets, when considering three characteristic orientation times (o1 = 20.2, o2 = 15.7, o3 = 12.6).
Time (min)
60
160
220
340
3600

1 5 (%)
Experimental, [Model]
31, [17]
40, [39]
34, [47]
56, [55]
57, [59]

1 10 (%)
Experimental, [Model]
43, [31]
60, [59]
55, [66]
67, [73]
66, [76]

1 15 (%)
Experimental, [Model]
48 [43]
66 [70]
64 [76]
71 [81]
70 [83]

The model is able to capture this behaviour, even considering the varying fibre lengths
and the varying magnetic coatings that they are associated with. In particular, the model
is able to give a reasonable approximation of the fibre angle distribution after 24 hours,
whereas conventional models (e.g. See Refs.[52, 134]) would predict a much narrower
distribution due to the increase in viscosity being neglected.
The fitting of the data could be further improved by the use of more characteristic orientation times, i.e. oi . However, around 600 fibres were analysed for each specimen, and so a
distribution of 200 fibres (of approximately equal AR) were used to fit each o value; further
optimisation of Eq. 4.10 would have reduced the significance of the experimental data for

138

Magnetic Response

the parametric fitting.


Neodymium Magnetic Alignment
The design space of reinforced elastomers, wherein the reinforcement is aligned, is potentially huge. However, to achieve such a design space an understanding of the potential
configurations needs to be realised; this can be achieved with an arrangement of permanent
magnets, and this is here demonstrated.
Neodymium Homogeneous Field
An arrangement of four permanent N52 neodymium magnets, positioned around the
curing specimen (Fig. 4.9a), produce a non-homogeneous field, but by careful placement
of the magnets, an approximately homogeneous field can be produced in the region of the
specimen (see Fig. 4.9b). Confirmation of the homogeneity can be seen in the produced
specimens (Here the concentration shown is 6 wt% NiC, as the specimens are used in the experimental tensile testing in section 2.3) which are transversely isotropic and homogeneous
(Fig. 4.9c); the darker area around the specimen is due to the thickness of the specimen
being larger, rather than any accumulation of fibres.

Fig. 4.9 (a) The Neodymium magnetic set-up to create an approximately homogeneous field
for alignment of the reinforcement, (b) The magnetic field strength (0.07-0.08T) approximation calculated using FEMM (ver.4.2), (c) The cured specimen; the arrow indicates the
in-plane direction of the reinforcement
At these increased concentrations (e.g. 6 wt%), the specimens begin to show small
agglomerations such as those shown in Fig. 4.10. Whilst their presence is limited, they are
likely to reduce the mechanical performance of the material and at higher fibre concentrations it is important to develop strategies to reduce their presence [220]. This is often an
issue in fibre reinforced elastomers, however the presence of the agglomerations is likely

4.2 Magnetically tailoring the orientation of reinforcements

139

furthered by the magnetic field; their ferromagnetic nature means they develop a weak magnetisation of their own under the magnetic field, and therefore attract together.

Fig. 4.10 (a) The post-cure fibre distribution from a specimen aligned between the four
Neodymium N52 magnets, fitted to Eq. 2.3 (b=4.3). (b) Manufactured defects and fibre
interaction during the curing process cause some agglomerations to form.
The result is a material with reinforced fibres oriented along the magnetic field lines, the
degree of transverse isotropy indicated by the orientation distribution (Fig. 4.10a).
Neodymium Out-of-Plane Reinforcement
The potential for out-of-plane curing has been explored. In Fig. 4.11 we can observe
the effect of suspending the specimen above the Neodymium magnetic arrangement. Outof-plane reinforcement would be of benefit to applications that require a strong throughthickness reinforcement, and could be considered as an option in through thickness repair
of composites. The angle of the reinforcement through the thickness can be controlled by
consideration of the arrangement of the magnetic field lines.
The variation of the magnetic field is reasonably small throughout the region of the
curing specimen (In respect to its effect upon the magnetic orientation of the fibres, as seen
in Fig. 4.4b), and as a result very little non-homogeneity of fibre distribution is present in
the specimen seen in Fig. 4.11.
Fig. 4.11c shows the in-plane variation of the fibre angle, which indicates the fibre direction is in-plane at the extremities of the specimen mould. There is significant variation
in the fibre angle due to the specimen mould being larger than the magnet. Towards the
magnet centre, indicated by the white dot, the fibres orient themselves increasingly through
the thickness until they are completely vertical. The through-thickness reinforcement can

140

Magnetic Response

Fig. 4.11 (a) The neodymium magnetic set-up to create an out-of-plane reinforcement, (b)
The magnetic field strength approximation calculated using the FEMM software [170], (c)
The cured specimen; the arrow indicates the in-plane direction of the reinforcement. The
centre dot indicates the fibres are oriented orthogonal to the viewed plane.
be viewed in Fig. 4.12.

Fig. 4.12 The out-of-plane cured specimen has a fibre reinforcement angle that varies along
the length of the specimen. The length along the specimen is shown in each image.
Neodymium Variable Angle Reinforcement
When considering complex geometries, or materials undergoing complex loading, it
can be advantageous to consider variable angles throughout the plane of the reinforcement.
Such techniques, which are ubiquitous with natural materials [215], have also been used to
great effect in conventional composite materials [125], and are demonstrated in Fig. 4.13.
Complex variations in the orientation can potentially be produced, and simply depend upon
the configuration of the magnetic field. In this case the in-plane angle of the reinforcement
can be seen to vary along the length and width of the specimen in Fig. 4.13c.

4.2 Magnetically tailoring the orientation of reinforcements

141

Fig. 4.13 (a) The non-homogeneous neodymium magnetic set-up to create an, in-plane,
variable angle reinforcement, (b) The magnetic field strength approximation calculated using FEMM (ver.4.2), (c) The cured specimen; the arrow indicates the in-plane direction of
the reinforcement.

142

4.3
4.3.1

Magnetic Response

Magnetic Actuation
Introduction

The bespoke orientation of reinforcements in solution, by the action of magnetic torques,


is an interesting strategy for material design. However, when the reinforcement is fixed
within the matrix (i.e. post-cure), the magnetic torque is transferred to the overall body of
the material; the result is controlled deformation and locomotion of the specimen.
The exact actuation response of the specimen is determined by the relationship between
the mechanical and field responsiveness, due to the coupled properties that have been tested
in a number of orthogonal configurations [243, 244]. The benefits of oblique arrangements
are predicted by a number of theoretical and numerical models that describe the behaviour
in a magnetic/electric field [22, 70, 213, 214], however there is a lack of experimental investigation of these effects.
This knowledge would allow more complex actuation, as the actuation can be controlled
by both anisotropic mechanical properties and anisotropic magnetic responsiveness. This
magnetic stimulation is contactless and non-invasive, and therefore ideal for bio-medical
applications such as micro-swimmers and micro-actuators [71, 132, 169, 193].
In this section, inspired by the developing interest in nano- and micro-actuating devices
for bio-medical applications such as drug delivery systems, lab-on-chip devices, as well as
new methods for performing micro-surgery and nano/micro fabrication [169], we look at the
creation of a novel magnetically and mechanically anisotropic PDMS elastomer reinforced
composite. The material actuates in a low magnetic field, and shows tailored properties that
depend upon the reinforcing fibre orientation. In particular, the configuration of the actuated
specimen depends critically on the reinforcement orientation, with the intensity of the shape
change dependent on the field intensity. As such, the effect of the magnetic field on the
transversally isotropic mechanical properties is also investigated. It is demonstrated that the
actuation can be controlled in a novel manner, taking advantage of the fibre reinforcement
and the resulting coupled behaviour between the mechanical and magnetic properties.

4.3.2

Experimental Techniques

PDMS (Sylgard 184) is reinforced by nickel-coated carbon fibres (Ni/C 40/60 chopped to
0.25 mm, diameter 4.8m, are purchased from Marktek Inc. and used as received). The fibre reinforced PDMS elastomers are made by direct mixing of the two constituents, and subsequently aligned in the presence of a magnetic field created by four neodymium magnets; a
detailing of the exact sample preparation can be found in section 4.2. The fibres are magnet-

4.3 Magnetic Actuation

143

Fig. 4.14 Experimental setup of the electromagnet for static actuation of the specimens, with
the magnetic field directed orthogonal to the plane of the fibres. Inset: A sample specimen
of 0 = 90 , actuated in the static magnetic field.
ically responsive, due primarily to the ferromagnetic nickel-functionalisation; the result is
an elastomer that is also responsive to a magnetic field. Rectangular specimens of thickness
0.5 mm are produced with fibre alignments at angles of 0 = {0 , 15 , 30 , 45 , 60 ,
75 , 90 }. The specimens are then cut to dimensions of 30 mm length (gauge length of 27.5
mm in experimental testing) and 7mm width. Burn-off of the solution residue indicated a
concentration of nickel-coated carbon fibres in the PDMS of 6 wt%.
To observe the actuation behaviour, samples are clamped at one end and vertically suspended in between the plates of an electromagnet as shown in Fig. 4.14. The electromagnet
produces a homogeneous field, which is orthogonal to the fibres and has an intensity controlled by varying the absorbed current. At each magnetic field intensity the specimen is
allowed to settle to a stable state. To avoid the effects of gravity and self-weight, in all tests
the samples are vertically oriented.
The presence of magnetically responsive fibres within the elastomer mean that, not only
is the material able to actuate in a magnetic field, but the mechanical performance is also
affected. Therefore, the effect of a uniform magnetic field upon the mechanical properties
was investigated. In order to characterise the effects, the magnetic field is oriented in two
configurations with respect to the material fibre alignment; both of which are orthogonal
to the loading direction and are described as perpendicular and transverse (As described in
Fig. 4.17). The effect of the magnetic field was found to be smaller than the deviation of
the moduli between equivalent specimens (The cyclic testing is shown in Appendix C.1),
therefore individual specimens were conditioned (by cycling to 10% strain 3 times) and
then cyclically tested to 10% strain. The testing involved repeated mechanical loading of

144

Magnetic Response

the specimens in the presence of an increasing magnetic field strength up to 0.09 T. The
mechanical testing of specimens up to 10% strain is justified by consideration that the maximum strain encountered during actuation is no more than 6%.

4.3.3

Magnetic Response in a Static Homogeneous Field

In the presence of a magnetic field, the elastomer material actuates due to the magnetically
responsive reinforcing fibres. The alignment of these fibres causes two main effects:
Firstly, assuming the fibres are perfectly aligned and orthogonal to the magnetic field,
no torque acts on the sample and it remains undeformed. However a critical magnetic
(crit)
field can be reached, By , at which the configuration becomes unstable, leading to
an equal probability of actuation in a positive or negative direction. This instability
can be visualised experimentally as a sudden increase in the actuation angle when the
magnetic field is slowly increased around this critical value.
Secondly, the actuation type is determined by the fibre alignment angle, 0 , varying
from bending only (0 ) to twisting only (90 ); angles in between show a combination
of bending and twisting actuation. This is caused by the anisotropic magnetic susceptibility of the fibres. The fibres attempt to align themselves parallel to the applied
homogeneous magnetic field, and the resulting torque causes bending/twisting of the
elastomer (See Fig. 4.15).
The results of the actuation in a static homogeneous magnetic field are shown in Fig. 4.16.
In the case of the 0 specimen and the 90 specimen, it is clear to see the bending and twisting behaviour, respectively. At angles between; a combination of bending and twisting
occurs, although one configuration is typically dominant in observations. It is seen that,
in general, the actuation angle increases slowly as the magnetic field intensity increases.
(crit)
However at a certain intensity, By , there is a more rapid increase that is associated with
an instability of the configuration. The gentle increase in the actuation angle before the
instability is most likely caused by the misalignment of fibres within the specimens (See
Fig. 4.16), and means that some fibres are not orthogonal to the magnetic field; as has been
confirmed in Fig. 4.10.
Magneto-Mechanical Testing
The magnetic actuation can be described by considering the magnetic and mechanical strain
energies of the specimen in a magnetic field. However, the mechanical deformation of a

4.3 Magnetic Actuation

145

Fig. 4.15 Static Actuation of specimens with reinforcement angles, 0 , at 0 , 30 , 60 &


90 . Specimens of length 27.5mm, width 7mm and thickness 0.5mm are presented to a
homogeneous magnetic field. A transition from pure bending for 0 , to pure twisting for
90 is observed. For the orientation of the axes see Fig. 4.18.

magnetically responsive material in the presence of a magnetic field can have significant
effects upon the mechanical properties (See, for example, Refs.[242244]) and therefore requires assessment. The results of this can be seen in Fig. 4.17. They show that the magnetic
field has a weak effect on the mechanical properties of the specimen when the magnetic
field is oriented perpendicular to the mechanical load. The largest effect is observed for the
specimen with fibres aligned at 0 degrees, however the magnitude of the modulus increase
for all alignment configurations is smaller than the standard deviation observed in the experimental testing under no external field (Fig. 2.23) and much smaller than the effect of
transverse isotropy, which demonstrates a near magnitude change in modulus between the
stiffest and softest configurations.
In spite of this low magnetic susceptibility, the low bending/twisting stiffness of the
material allows magnetic actuation to be observed in the presence of a magnetic field. Typically, a magneto-mechanical transversely isotropic model would be required to describe this
behaviour, however the minimal effect of the magnetic field on the mechanical properties
suggests the magneto-mechanical behaviour can be adequately estimated by the transversely
isotropic model described in section 2.3.

146

Magnetic Response
20

20

Bend Angle, ( )

Bend Angle, ( )

(a)
15

10

(b)

15

10

5
Bend

Bend
0
0

0.05

0.1

0
0

0.15

20

Bend Angle, ( )

(c)
15

10

5
Bend
0
0

0.02

0.04

0.06

0.08

0.05

0.1

0.15

Magnetic Field (T)

0.1

Bend/Twist Angle, / ( )

Magnetic Field (T)

0.12

15

(d)
10

5
Bend
Twist
0
0

0.05

0.1

0.15

Magnetic Field (T)

Magnetic Field (T)


20

(e)

Twist Angle, ( )

Twist Angle, ( )

20
15
10
5

(f)
15

10

Twist
0
0

0.05

0.1

0.15

Twist
0
0

Magnetic Field (T)

0.05

0.1

0.15

Magnetic Field (T)

Twist Angle, ( )

25

(g)
20
15
10
5
Twist
0
0

0.05

0.1

0.15

Magnetic Field (T)

Fig. 4.16 Static bending/twisting actuation of specimens with different fibre alignments; (a)
0 = 0 , (b) 0 = 15 , (c) 0 = 30 , (d) 0 = 45 , (e) 0 = 60 , (f) 0 = 75 , (g) 0 = 90 .
The dominant actuation angle (bending or twisting) is recorded in each case.

147

4.3 Magnetic Actuation


3

(a)

Initial Modulus (MPa)

Initial Modulus (MPa)

2.5

2
0.00 T
0.03 T
0.06 T
0.09 T
Eq. 2.8

1.5

20

40

60

(b)
2.5

80

0.00 T
0.03 T
0.06 T
0.09 T
Eq. 2.8

1.5

20

Fibre Angle, 0 ( )

40

60

80

Fibre Angle, 0 (o )

Fig. 4.17 The magneto-mechanical properties of the magnetically responsive specimens, in


a homogeneous magnetic field of maximum strength 0.09 Tesla. (a) The perpendicular arrangement, whereby 0 is orthogonal to the field. (b) The transverse arrangement, whereby
the field lines are in the plane of the reinforcements. At 0 = 90 the field is parallel to the
field lines.

Model Description
To investigate the actuating (i.e. bending/twisting) behaviour, a large rotations beam model
is introduced for which the total energy is the sum of an elastic, e , and a magnetic part,
m , i.e.,
Z L

Etot = ht

(e + m ) ds

(4.14)

where h and t are the beam width and thickness respectively, and s is the abscissa along the
specimen length (0 s L). e and m are given by,
e =

EI 2 GJ 2
(s) + 2 (s) ,
2 L2
2L

m = (n.B)2 .

(4.15)

Here, the elastic energy is expressed in terms of the bending, , and twisting, , angles
which are shown in Fig. 4.18, and accounts for large rotations/displacements. EI and GJ
are the bending and twisting rigidities, respectively. The magnetic energy, m , depends
upon = a f /(20 ), where a represents the magnetic anisotropic susceptibility, 0 the
vacuum permittivity and f is the volume fraction of the fibres [131]. In addition, it depends
upon the relative angle between the magnetic field B = {0, By , 0} (here assumed to be directed along the y-axis in Fig. 4.18) and the fibre orientation in the deformed configuration
n = {nx , ny , nz }, i.e.,

148

Magnetic Response

Fig. 4.18 Schematic representation of the beam model. (a) The undeformed beam showing
the reinforcement angle in the x-z plane, 0 . (b) The twisting angle, . (c) The bending
angle, .

nx = sin(0 ) sin( ) sin() + cos(0 ) cos( )


ny = cos(0 ) sin( ) sin(0 ) cos( ) sin()

(4.16)

nz = sin(0 ) cos() ,
To study the equilibrium shape of the beam it is important to define the bending and
twisting angles,
(s) =

m
s,
L

(s) =

m
s,
L

(4.17)

The deformed configuration of the beam can then be identified by looking at the minimum of the total energy, which takes the following form,

EeTotal =

B2y L cos(0 )2
(2m sin(2m ))
8m


B2y L sin(20 ) 
+
m cos(m ) sin(2m ) 2m cos(2m ) sin(m )
8m


1 2
sin(2(m + m ))
2 sin(2(m m ))
+ By L sin(0 )
+
32
m m
m + m


EIm2 GJm2
1 2
2 sin(m ) sin(2m )
2
By L sin(0 ) 2 +

+
+
16
m
m
2L3
2L3

(4.18)

149

4.3 Magnetic Actuation

The resulting expression is highly nonlinear, and only numerical solutions can be derived. Never-the-less, it is possible to take a fourth order expansion of the total energy with
respect to the maximum bending and twisting angles; in this way, a closed form approximate
expression of the stable and unstable equilibria configurations of the beam can be obtained.
This approximate expression of the total energy in terms of the maximum bending m and
twisting m angles of the beam is thus given by,

ETotal =

4
1 2 4
By m L cos2 (0 ) B2y m3 Lm sin(0 ) cos(0 )
15
15

1
1
+ B2y m2 Lm2 sin2 (0 ) B2y m2 L cos2 (0 )
5
3

1 2
2
By m Lm3 sin(0 ) cos(0 ) + B2y m Lm sin(0 ) cos(0 )
15
3

1
EIm2 GJm2
1 2 4
By Lm sin2 (0 ) B2y Lm2 sin2 (0 ) +
+
15
3
2L3
2L3

(4.19)

The minimum of this expression describes the stable (undeformed) configuration of the
(crit)
specimen at low magnetic field strengths, however when By By
the configuration becomes unstable (a local maximum of the energy) and two symmetric minima appear. Due
to the equal weight of these minima, the configuration has an equal possibility of a nega(crit)
tive or positive actuation angle. The occurrence of By
can be studied by looking at the
determinant of the Hessian matrix, i.e. the second derivatives of the total energy,
2 Eetot

2
0 ) = m
H(
2 Eetot

m m

2 Eetot
m m

2 Eetot
m2

m =0,m =0
(crit)

which gives the following value of the critical magnetic field By


is zero,
s
(crit)
By ) =

2 L4

(4.20)

at which the determinant

3 EI
.
EI/GJ sin2 (0 ) + cos2 (0 )

(4.21)

In this situation the reinforcing fibres attempt to align to the magnetic field lines, how(crit)
ever the magnitude of By , and the actuation angle at magnetic field strengths above this,
are highly dependent on both the magnetic properties, i.e. , and the mechanical properties,
i.e. E, G. These factors are, in turn, highly dependent upon the alignment of the fibres.

150

4.3.4

Magnetic Response

Results and Discussion


(crit)

can be experimentally determined by slowly increasing the intensity of


The value of By
the magnetic field up to the point at which a sudden jump in the deformation of the beam is
observed. This effect has been experimentally observed and is shown in Figs. 4.16a-g. For
example, in Fig. 4.16f, a specimen with a 75 reinforcement alignment is shown to have a
sudden increase in the actuation angle at 0.07T, which is captured in the camera samples
on the figure. The slow increase in the actuation angle at field strengths below the critical
value is likely due to the mis-alignment of some fibres in the material. It is also observed
that in each case, a dominant mode of actuation angle is present that suppresses the other,
as shown by the relatively low bending angle, compared to the twisting angle, for 0 = 45 .
(crit)
predicted by the model through Eq. (4.21) are shown in Fig. 4.19
The values of By
against the experimental data. The fitting in Fig. 4.19 was achieved with a value of the
anisotropic susceptibility, a = 2.20 103 ; which is about thirty times larger than the
one reported in the literature for neat carbon fibres [131] and is likely due to the nickel
functionalisation of the fibres, which increase the fibre magnetic response and make the
specimen easier to orient in low magnetic fields. The results show the close agreement of
the model to the experimental results, in particular showing the increased stability of the
bending modes, e.g. 0 , and the subsequent increased value of the critical field strength,
(crit)
By .
The minimisation of (4.19) allows the actuation angles to be determined immediately
(crit)
after the occurrence of the instability (i.e. By = 101%By ), and compared to the prediction of the model in light of the results from Fig. 4.19. Fig. 4.20 shows these actuation
(crit)
angles; this was obtained with parametric values used in calculating By , in addition to
the determined value of a , i.e. 2.20103 . Although a certain degree of dispersion in the
fibre angle is present (See Fig. 4.10), the actuation of the specimens in the magnetic field
confirms the behaviour predicted by the model. Indeed, as seen in Fig. 4.16, a 0 specimen
exhibited only bending, whilst only twisting is observed at 90 . At angles between, either
twisting or bending is observed as the dominant actuation mechanism; although it is still
possible to observe both (See Fig. 4.16d).

151

4.3 Magnetic Actuation

0.15

By

(crit)

(T)

0.13
0.11
0.09
0.07
0.05

15

30

45

60

75

90

Fibre Angle, ( )
(crit)

Fig. 4.19 Critical magnetic field intensity By


at which the undeformed configuration
becomes unstable, assuming the model (4.21). (Parameters: E(0 ) and G(0 ) taken from the
results of the mechanical testing (i.e. = 0.252MPa, = 1.85), I = 1.261013 (m4 ), J =
4.771013 (m4 ), L = 27.5103 (m), a = 2.20103 and f = 6 %). The black dots with
error bars represent the results of the experiment carried out by the authors for specimens
with fibres at 0 = {0 , 15 , 30 , 45 , 60 , 75 , 90 }.

152

Bending/Twisting Angle ( )

Magnetic Response

12

8
Bending
Twisting
Bending - Model
Twisting - Model

0
0

15

30

45

60

75

90

Fibre Angle,
Fig. 4.20 The bending and twisting angles achieved for an imposed orthogonal magnetic
(crit)
field at a field strength just greater than at the occurrence of instability (i.e. 1.01By ),
assuming the model (4.19). (Parameters: E(0 ) and G(0 ) taken from the results of the mechanical testing (i.e. = 0.252MPa, = 1.85), I = 1.261013 (m4 ), J = 4.771013 (m4 ),
L = 27.5103 (m), a = 2.20103 and f = 6 %). The orange dots and purple dots, represent the bending and twisting experimental results, respectively, for specimens with fibre
angles at 0 = {0 , 15 , 30 , 45 , 60 , 75 , 90 }. The solid line represents the model (4.19).
The error bars represent the experimental prediction of the actuation angle, when extrapolated from the magnetic field 1% before and after the occurence of the instability (Given by
the dashed line in Fig. 4.16).

4.4 Concluding Remarks

4.4

153

Concluding Remarks

In this chapter, motivated by the potential of spatially oriented dis-continuous fibres, we


align magnetic nickel-coated carbon fibres in a viscous PDMS solution, whose viscosity
is increasing with time, with a magnetic field. The novel NiC-PDMS composites were
fabricated by dispersing the fibres in the host matrix by sonication and mechanical mixing,
then using low magnetic fields to orientate the fibres. Successful dispersion and alignment
of the fibres was observed, although a small number of agglomerations, misaligned particles
and varying fibre lengths were evident. These features of the fibres are a result of the fibre
manufacture processes, as confirmed by SEM images, and the loss of some of the nickel
coating causes the presence of unaligned and partially aligned fibres.
The distribution of the aligned fibres is analysed, both during the cure process and after, using an optical microscope; this allows the position and orientation of the fibres to be
recorded over time. Investigations of the fibre distributions under various magnetic field
strengths, and at time intervals, are undertaken. The strength of the magnetic field and the
exposure time both define the distribution of fibres, whilst the field direction determines the
mechanical and magnetic properties (i.e. an angle of isotropy). These results are compared
to the introduced model, which models the orientation of an initially homogenous distribution of fibres under a homogeneous magnetic field, and indicates its alignment; subject to
a number of environmental factors. It is possible to extend the model to the application of
non-homogeneous fields, although it is then difficult to derive an analytical solution.
Discrepancies to the model are found. For instance, larger fibres orient more quickly,
even though the inertial forces are greater than any increase in magnetic torque. This is
caused by damage to the fibres during their manufacture, as indicated by SEM images,
which show the magnetic coating has been damaged badly on many of the smaller fibres.
This damage, and defects such as fibre agglomerations and uneven nickel coating, contribute to the mis-alignment of many of the fibres. In addition, at higher fibre concentrations,
chain-like structures and agglomerations are more likely to form due to the magnetic coupling of the ferromagnetic fibres.
It is also demonstrated that spatial alignment is possible, which is a potentially useful
strategy to optimise material performance. To achieve this, four permanent neodymium
magnets are arranged in different configurations to investigate the orientational potential
of through thickness reinforcement, variable angle reinforcement and the use of a nonhomogeneous field.
Transversely isotropic specimens of the material are produced at a number of oblique
angles. Due to the magnetic responsiveness of the fibres in the cured matrix, the magnetic
torque is transferred to the specimen. The result is that the mechanical and magnetic prop-

154

Magnetic Response

erties are controlled by the magnetic field properties, specifically the angle of isotropy is
defined by the direction of the magnetic field lines during the curing phase.
Thus, in the presence of a magnetic field, the material has mechanical and actuation
properties that are both dependent on the field properties. However, the effects of a magnetic
field on the magneto-mechanical properties are minimal for low fibre content, allowing the
introduction of a simple constitutive model to describe the magnetic behaviour.
This novel composite was magnetically actuated to determine its potential for use in
micro-actuator systems. This can be achieved in a low magnetic field (< 0.2 T), due to
the nickel-functionalisation of the fibres. The average anisotropic susceptibility of the fibres
identified from the experiments, by comparison to the model, was a = 2.20 103 , which
is about thirty times larger than the one reported in the literature for neat carbon fibres.
The composite actuator has multi-stable behaviour, induced by the magnetic torque acting
on the fibres: the samples remained undeformed up to a critical magnetic field strength,
after which a sudden increase in the deformation occurs; at which point either a positive or
negative actuation is equally likely. This is predicted by the introduced model, which also
predicts the twisting and bending behaviour that are independently observed for different
initial orientations of the fibre.
Discrepancies between the predicted behaviour and the experimental results can be observed in the observations of the instability. The onset is more difficult to define, given that
a slow increase in the actuation angle is observed in the experimental tests leading up to
the instability. This discrepancy can be attributed to the mis-alignment of the fibre angles,
which are partially aligned away from the plane of isotropy, and changing the relative angle between the magnetic field and fibre orientation, Eq. 4.16. Incidentally, the introduced
model could be adapted to consider misalignment of the fibres, or even variation of the
alignment angle along the length of the beam (i.e. 0 (s)), however this would likely require
the use of numerical solutions to determine the shape of the deformed beam.
It is anticipated that an appreciation of potential orientation configurations, and the associated magnetic manufacturing difficulties, will drive further innovation into novel materials. Particularly in the field of self-assembly, which is becoming an increasingly important
area of research as the need for highly optimised materials continues to develop. However,
it has been demonstrated that the the oblique angle between aligned & embedded fibres and
an applied magnetic field can induce a bespoke actuation behaviour. Similar actuation of
advanced material designs could be utilised to develop complex steering mechanisms for
use in a wide variety of micro-actuator systems such as micro-pumps and micro-valves.
Additional configurations can be envisaged, by consideration of the fibre alignment in other
planes, and spatially aligned. Future work requires the development of a device to test in

4.4 Concluding Remarks

155

a low Reynolds number flow, to determine its performance compared to similar, existing,
micro-swimming devices.

Chapter 5
Conclusions
5.1

Original Contributions

In studying the anisotropic effects of nano-inclusions, section 2.2, the mechanical and viscoelastic properties of a graphene oxide reinforced unvulcanised natural rubber are investigated. The report demonstrates that:
Natural rubber nanocomposites, reinforced by graphene oxide, have been successfully
produced by latex-mixing in a simple process that requires no additional heat or
chemical input. The process creates homogeneous specimens containing a regular
size distribution of oxidised nanoparticles of single layer thickness.
A thorough mechanical investigation of the filler network is undertaken. The introduction of the nanoparticles was found to increase the stiffness and strength. The
viscoelastic properties are enhanced by the nano reinforcements and a strong strain
rate dependence is found, which is increased by the addition of the fillers. Cyclic tests
show that an increase in the dissipative energy with filler concentration, caused by
friction between the platelets and the matrix, is greater than the increase in the elastic
energy.
An investigation of the filler network has also been undertaken by retesting the specimen 24 hours after an initial large strain deformation. This test indicated that rotation of the GO is likely to have occurred at large strain, which is permanent to some
degree. This is evidenced by the reduction in strain, for increasing GO concentration,
at which the characteristic upturn in the stress-strain curve occurs.
In response, manufacturing methods of producing anisotropic elastomer composites are explored, section 4.2. The need for optical observation of the fillers in-situ, requires the need

158

Conclusions

for micro-dimensioned fibres. Contributions to the open literature include:


It is demonstrated that novel 3-dimensional spatial orientation of the fibre reinforcements can be achieved with the use of a small number of permanent neodymium
magnets. In-plane, through-thickness and variable angle configurations are all developed.
The validation of constitutive models to predict the orientation of the fibres in a solution, for the purposes of spatial orientation in reinforcing materials, is particularly difficult due to the challenges in recording the position and orientation of a large enough
number of fibres over time. Here a parametric study is experimentally achieved in
terms of the effects of magnetic field strength and time, for a large number of fibres
( 600), and is compared to a constitutive model that considers the increasing viscosity of the PDMS solution due to the cure process.
Transversely isotropic manufactured are subsequently experimentally tested to investigate
the anisotropic effect of micro-inclusions, section 2.3.
The produced specimens are shown to have tensile mechanical properties that depend highly upon the reinforcement angle, 0 . This behaviour can be described by
a hyperelastic transversely isotropic constitutive model, for all reinforcement angles,
assuming a plane stress deformation, up to 30%.
The discrepancies between the experimental data at strains above 30% are investigated
and are primarily attributed to the imperfect interface between the fibres and matrix.
Discrepancies can also be attributed to the defined deformation gradient, which is less
effective as the lateral constraint becomes significant (i.e. at reinforcement angles
approaching 90 ).
The rotation of the fibres during mechanical tests is recorded in-situ, up to 60% strain.
The rotation of the fibres helps to validate the constitute model. The rotation of fibres
at large strain is known to cause instabilities, particularly when in compression, and
such a methodology would be useful to a number of investigations.
A large strain numerical model is developed in the FE environment of ABAQUS (ver. 6.12),
section 3, in order to further investigate the effects of the anisotropic inclusion in a controlled environment. To describe the effects of a fibre reinforced composite, a plane stress
numerical model is developed. The multiple fibre representative element (RVE) replicates
the uniaxial tension experiments of the reinforced PDMS specimens of sections 2.2 & 4.2.

5.1 Original Contributions

159

A platelet-like reinforcement has two reinforcing dimensions, and so the lateral constraint
needs to be accounted for. This is modelled as a 3D multiple platelet RVE, subject to a plane
strain deformation. However, it is advantageous to optimise the computational efficiency of
the model.
It is shown that the 2D multiple fibre RVE can be simplified to a single reinforcement
RVE, when the fibre interaction is negligible (i.e. low wt%).
The effects of the fibre angle distribution on the tangent modulus, strength of the
transverse isotropy and the stress-strain behaviour are investigated. The results quantify the effects of fibre misalignment, and suggest the effect is minimal for moderate
levels of misalignment.
It is shown that the 3D multiple platelet RVE can be simplified to a 2D single platelet
RVE, via a 2D multiple platelet RVE. The simplifications result in a small loss of
accuracy, but a large increase in computational efficiency.
A parametric study is undertaken, and a modified form of the Halpin-Tsai equation
is presented to describe the platelet reinforced RVE. This illustrates the logarithmic
relationship of the parameters that the H-T equation does not predict.
Many features of the single reinforcement RVE numerical models are captured by the
neo-Hookean transversely isotropic constitutive model up to 40-60% strain. Some
discrepancies were found, and are generally ascribed to the compression of the reinforcement, causing instabilities to arise. Never-the-less, the simple model is able to
simultaneously describe 19 angular configurations, in all possible deformation gradients for both the plane stress and plane strain numerical models. This is all done with
only 2 constitutive parameters.
When the magnetically responsive fibres are embedded in the cure elastomer, the magnetic
torque is transferred to the specimen and actuation is possible. This magnetic actuation is
explored in section 4.3.
Previous reports have used a non-uniform magnetic field to actuate spherical particle reinforced elastomers. Here, the use of responsive fibres means that actuation is
possible in a uniform magnetic field, due to the created torque.
Never-the-less, anisotropy with spherical particles is possible due to fibre-fibre interactions that form chain-like structures. This is not desirable, as it can result in
non-homogeneity of the filler distribution and therefore a reduction of the mechanical

160

Conclusions
properties. With fibre reinforcements, this anisotropy is possible without translation
of the fibres, which helps to ensure an even distribution.
The bending and twisting actuation effects depend on the fibre orientation and can
be found, in various forms, in the literature. However, we demonstrate the controlled
combination of bending and twisting, that ranges from bending only, to twisting only,
and depends upon the reinforcement angle.
(crit)

An instability is observed at a critical magnetic field strength, By , due to the equal


probability of a negative or positive actuation angle. This is caused by the orthogonal
arrangement of the reinforcing fibres, and could not be observed for homogeneously
distributed spherical particles.
A large rotations/displacements beam model has been developed that describes the ac(crit)
tuating bend/twist behaviour, and predicts the existence of the instability point, By .
Magneto-mechanical testing has shown the effects of the magnetic field on the sample
are negligible, and as a result that a simple hyperelastic model can be used to describe
the transversely isotropic model at small strain.
In turn this has allowed the large rotations/displacements beam model to describe the
experimental actuating behaviour, and its dependence on the magnetic field strength
and reinforcement angle, 0 .

5.2

Further Perspectives

The work has demonstrated the effectiveness the anisotropic reinforcements in elastomers
at large strain. However, the subject is vast and a number of key challenges have been identified for the journey ahead.
Experimental techniques for large strain transverse isotropy
The simple tensile testing of anisotropic materials is challenging, as shear and bending
forces are created; due to the oblique angle of the transferred forces [45]. Xiao et al. [256]
discusses many of the difficulties in testing transversely isotropic materials. Oblique test
fixtures can often be used to better transfer the tensile strains, however this more challenging at large strain due to the rotation of the reinforcing direction. As a result, relatively few
experimental procedures have tested the various loading conditions of anisotropic materials, and none comprehensively [24]. The experimental validation of anisotropic constitutive

5.2 Further Perspectives

161

models is essential, and due to the technical challenges, numerical methods are often used
instead. The development of a successful rotating test fixture or similar apparatus, to follow the oblique angle at large strain, would be a big step in the experimental testing of
anisotropic materials.
Efficient modelling of complex effects with dis-continuous reinforcements
A simplified single reinforcement RVE has demonstrated its ability to numerically simulate the behaviour of dis-continuous reinforcements at low concentrations. Its validation
potentially allows its use in the exploration of more complex phenomenon, such as crack
propagation and interfacial properties, in a computationally efficient manner.
Extending the fibre-rotations model
The fibre-rotations model, used to predict the distribution of suspended fibres in a homogeneous field, can be extended to consider the effects of a non-homogeneous field. This
would be vital in the utilisation of non-homogeneous fields to produce bespoke arrangements of spatially aligned and distributed fibres. Similar experimental validation would be
required, especially as locomotion of the fibres occurs in addition to rotation.
In addition, the damage to the magnetic fibre coating requires 3 characteristic time constants to be used, i.e. 01,02,03 . This can be extended to directly consider multiple types of
fibres, simultaneously oriented.
Magneto-mechanical testing of transversely isotropic materials
Magneto-Mechanical testing showed that the magnetic properties of the field and specimen had no significant effect upon the transversely isotropic properties in the presence of a
magnetic field. However, at larger magnetic field strengths and magnetic susceptibility the
effects will increase. Currently, only perpendicular arrangements of the magnetic field, load
and reinforcement angle are experimentally considered in the literature. However oblique
arrangements are of great interest to the scientific community, due to the interesting effects
that can be achieved, such as mixed modes of bending and twisting.
Extending the large displacements/rotations beam model
The large displacement/rotations beam model only considers a uniformly aligned arrangement of magnetic fibres perpendicular to the magnetic field. This model could be extended to consider oblique angles to the magnetic field, or the effects of fibre-misalignment.
In addition, if a higher magnetic field strength or larger filler concentration were used, then
the model would need to consider the magneto-rheological effects of the elastomer in a

162

Conclusions

magnetic field.
Design and test of a micro-swimmer prototype
The bending and twisting effects are envisaged to produce a propulsive force, potentially
for use in a micro-swimmers or actuating device. Future work would require the testing of a
prototype in a low Reynolds number flow, and if successful, development and further testing
of the device to determine the potential propulsive speeds of this system.

References
[1] Abraham, F., Alshuth, T., and Jerrams, S. (2005). The Effect of Minimum Stress and
Strain Amplitude on the Fatigue Life of Non Strain Crystallising Elastomers. Materials
& Design, 26:239245.
[2] Ajayan, P., Schadler, L., and Braun, P. (2006). Nanocomposite science and technology.
[3] Alexander, H. (1968). A Constitutive Relation Rubber-like materials. International
Journal of Engineering Science, 6(628).
[4] Amnuaypornsri, S., Toki, S., Hsiao, B., and Sakdapipanich, J. (2012). The effects of
endlinking network and entanglement to stress-strain relation and strain-induced crystallization of un-vulcanized and vulcanized natural rubber. Polymer, 53:33253330.
[5] Ampudia, J., Larrauri, E., Gil, E. M., Rodriguez, M., Leon, L. M. (1998). Thermal
Scanning Rheometric Analysis of Curing Kinetic of an Epoxy Resin . I . An Anhydride
as Curing Agent. Journal of applied polymer science, 71:12391245.
[6] Arruda, E. and Boyce, M. (1993). A Three-Dimensional Constitutive Model for the
Large Stretch Behavior of Rubber Elastic Materials. Journal of the Mechanics and
Physics of Solids, 41(2):389412.
[7] Barton, J. M., Greenfield, D. C. L., and Gu, H. (1992). Some Effects of Structure on the
Cure of Glycidylether Epoxy Resins. Polymer, 33(6):11771186.
[8] Benveniste, Y. (1987). A New Approach to the Application of Mori-Tanakas Theory
in Composite Materials. Mechanics of materials, 6:147157.
[9] Berriot, J., Montes, H., and Lequeux, F. (2002). Evidence for the Shift of the Glass Transition near the Particles in Silica-Filled Elastomers. Macromolecules, 35:97569762.
[10] Berthelot, J.-M. (1999). Composite Materials. Mechanical Behavior and Structural
Analysis.
[11] Beurrot, S., Huneau, B., and Verron, E. (2010). In situ SEM study of fatigue crack
growth mechanism in carbon black-filled natural rubber. Journal of Applied Polymer
Science, 117:12601269.
[12] Beurrot-Borgarino, S., Huneau, B., Verron, E., and Rublon, P. (2013). Strain-induced
crystallization of carbon black-filled natural rubber during fatigue measured by in situ
synchrotron X-ray diffraction. International Journal of Fatigue, 47:19.

164

References

[13] Bhattacharyya, S., Sinturel, C., Bahloul, O., Saboungi, M.-L., Thomas, S., and Salvetat, J.-P. (2008). Improving reinforcement of natural rubber by networking of activated
carbon nanotubes. Carbon, 46(7):10371045.
[14] Bokobza, L. (2007). Multiwall carbon nanotube elastomeric composites: A review.
Polymer, 48(17):49074920.
[15] Bokobza, L. (2012). Multiwall carbon nanotube-filled natural rubber: Electrical and
mechanical properties. Express Polymer Letters, 6(3):213223.
[16] Bokobza, L. and Kolodziej, M. (2006). On the use of carbon nanotubes as reinforcing
fillers for elastomeric materials. Polymer international, 55:10901098.
[17] Bonderer, L., Studart, A., and Gauckler, L. (2008). Bioinspired Design and Assembly
of Platelet Reinforced Polymer Films. Science, 319(February):10691073.
[18] Bonderer, L. J., Feldman, K., and Gauckler, L. J. (2010). Platelet-reinforced polymer matrix composites by combined gel-casting and hot-pressing. Part I: Polypropylene
matrix composites. Composites Science and Technology, 70(13):19581965.
[19] Bortz, D. R., Heras, E. G., and Martin-Gullon, I. (2012). Impressive Fatigue Life
and Fracture Toughness Improvements in Graphene Oxide/Epoxy Composites. Macromolecules, 45(1):238245.
[20] Boutevin, B., Boyer, C., Csetneki, I., David, G., Ferguson, J. S., Filipcsei, G., Gong,
B., Li, S., Li, W., Sanford, A. R., Szilagyi, A., and Zrinyi, M. (2007). Advances in
polymer science: Oligomers - Polymer Composites - Molecular Imprinting. Vol. 207
edition.
[21] Boyce, M. and Arruda, E. (2000). Constitutive models fo rubber elasticity: A Review.
Rubber chemistry and technology, 73:504523.
[22] Brigadnov, I. and Dorfmann, A. (2003). Mathematical modeling of magneto-sensitive
elastomers. International Journal of Solids and Structures, 40(18):46594674.
[23] Brinke, J. T. and Litvinov, V. (2002). Interactions of Stber Silica with Natural Rubber
under the Influence of Coupling Agents, Studied by 1H NMR T2 Relaxation Analysis.
Macromolecules, 35:1002610037.
[24] Brown, L. W. and Smith, L. M. (2011). A Simple Transversely Isotropic Hyperelastic
Constitutive Model Suitable for Finite Element Analysis of Fiber Reinforced Elastomers.
Journal of Engineering Materials and Technology, 133(2):021021.
[25] Budiansky, B. (1965). On the Elastic Moduli of Some Heterogeneous Materials. Journal of the Mechanics and Physics of Solids, 13(02):15.
[26] Burgert, I. and Fratzl, P. (2009). Actuation systems in plants as prototypes for bioinspired devices. Philosophical transactions. Series A, Mathematical, physical, and engineering sciences, 367(1893):154157.
[27] Burns, L., Mouritz, A., Pook, D., and Feih, S. (2012). Strength improvement to composite T-joints under bending through bio-inspired design. Composites Part A: Applied
Science and Manufacturing, 43(11):19711980.

References

165

[28] Bustamante, R. (2008). Transversely isotropic non-linear electro-active elastomers.


Acta Mechanica, 206(3-4):237259.
[29] Bustamante, R. (2009). Transversely isotropic nonlinear magneto-active elastomers.
Acta Mechanica, 210(3-4):183214.
[30] Camponeschi, E., Florkowski, B., Vance, R., Garrett, G., Garmestani, H., and Tannenbaum, R. (2006). Uniform Directional Alignment of Single-Walled Carbon Nanotubes
in Viscous Polymer Flow. Langmuir, 22(23):18581862.
[31] Camponeschi, E., Vance, R., Al-Haik, M., Garmestani, H., and Tannenbaum, R.
(2007). Properties of carbon nanotube-polymer composites aligned in a magnetic field.
Carbon, 45(10):20372046.
[32] Caner, F. C. and Carol, I. (2006). Microplane constitutive model and computational
framework for blood vessel tissue. Journal of biomechanical engineering, 128(3):419
27.
[33] Caner, F. C., Guo, Z., Moran, B., Bazant, Z. P., and Carol, I. (2007). Hyperelastic
anisotropic microplane constitutive model for annulus fibrosus. Journal of biomechanical
engineering, 129(5):63241.
[34] Chaiyasat, P., Suksawad, C., Nuruk, T., and Chaiyasat, A. (2012). Preparation
and characterization of nanocomposites of natural rubber with polystyrene and styrenemethacrylic acid copolymer nanoparticles. Express Polymer Letters, 6(6):511518.
[35] Chen, C. and Cheng, C. (1996). Effective elastic moduli of misoriented short-fiber
composites. International journal of solids and structures, 33(17):25192539.
[36] Chen, X. and Liu, Y. (2004). Square representative volume elements for evaluating
the effective material properties of carbon nanotube-based composites. Computational
Materials Science, 29(1):111.
[37] Chenal, J.-M., Gauthier, C., Chazeau, L., Guy, L., and Bomal, Y. (2007). Parameters
governing strain induced crystallisation in filled natural rubber. Polymer, 48:68936901.
[38] Choi, E. S., Brooks, J. S., Eaton, D. L., Al-Haik, M. S., Hussaini, M. Y., Garmestani,
H., Li, D., and Dahmen, K. (2003). Enhancement of thermal and electrical properties of
carbon nanotube polymer composites by magnetic field processing. Journal of Applied
Physics, 94(9):6034.
[39] Choi, W., Lahiri, I., Seelaboyina, R., and Kang, Y. S. (2010). Synthesis of Graphene
and Its Applications: A Review. Critical Reviews in Solid State and Materials Sciences,
35(1):5271.
[40] Ciambella, J. and Stanier, D. (2014). Orientation Effects in Short Fibre-Reinforced
Elastomers. In Proceedings of ASME 2014 International Mechanical Engineering Congress and Exposition, Montreal. American Society of Mechanical Engineers
(ASME).

166

References

[41] Ciambella, J., Stanier, D., and Rahatekar, S. S. (2016). Magnetic Alignment of Short
Carbon Fibres in Curing Composites by Uniform and Non-uniform Fields. (To Be Submitted).
[42] Ciesielski, A. (1999). An Introduction to Rubber Technology, volume 44. Great
Britain: Rapra Technology Ltd, Shawbury, Shrewsbury, Shropshire.
[43] Cotten, G. R., Boonstra, B. B., and Corporation, C. (1965). Stress Relaxation in Rubbers Containing Reinforced Fillers*. Journal of applied polymer science, 9:33953408.
[44] Debotton, G., Hariton, I., and Socolsky, E. (2006). Neo-Hookean fiber-reinforced composites in finite elasticity. Journal of the Mechanics and Physics of Solids, 54(3):533
559.
[45] Destrade, M., Donald, B. M., Murphy, J. G., and Saccomandi, G. (2013). At least
three invariants are necessary to model the mechanical response of incompressible, transversely isotropic materials. Computational Mechanics, 52(4):959969.
[46] Diani, J., Brieu, M., Vacherand, J.-M., and Rezgui, A. (2004). Directional model
for isotropic and anisotropic hyperelastic rubber-like materials. Mechanics of Materials,
36(4):313321.
[47] Donald, A. (2013). A cracking tale: why did the worlds first jetliner fall out of the
sky?
[48] Drozdov, A. and deC. Christiansen, J. (2007). Viscoelasticity and viscoplasticity of
semicrystalline polymers: Structureproperty relations for high-density polyethylene.
Computational Materials Science, 39(4):729751.
[49] Drozdov, A. and deC Christiansen, J. (2012). Cyclic viscoelastoplasticity of
polypropylene/nanoclay hybrids. Computational Materials Science, 53(1):396408.
[50] Dusi, M. R., May, C. A., and Seferis, J. C. (1982). Predictive Models as Aids to
Thermoset Resin Processing. In 1982, chapter 18, pages 301318. American Chemical
Society, Washington D.C.
[51] Eadie, L. and Ghosh, T. K. (2011). Biomimicry in textiles: past, present and potential.
An overview. Journal of the Royal Society, Interface / the Royal Society, 8(59):76175.
[52] Erb, R. M., Libanori, R., Rothfuchs, N., and Studart, A. R. (2012a). Composites
reinforced in three dimensions by using low magnetic fields. Science (New York, N.Y.),
335(6065):199204.
[53] Erb, R. M., Sander, J. S., Grisch, R., and Studart, A. R. (2013). Self-shaping
composites with programmable bioinspired microstructures. Nature Communications,
4(1712):18.
[54] Erb, R. M., Segmehl, J., Charilaou, M., Lffler, J. F., and Studart, A. R. (2012b). Nonlinear alignment dynamics in suspensions of platelets under rotating magnetic fields. Soft
Matter, 8(29):7604.

References

167

[55] Eshelby, J. D. (1957). The Determination of the Elastic Field of an Ellipsoidal Inclusion, and Related Problems. Proceedings of the Royal Society A: Mathematical, Physical
and Engineering Sciences, 241(1226):376396.
[56] Evans, B., Fiser, B., and Prins, W. (2012). A Highly Tunable Silicone-Based Magnetic
Elastomer with Nanoscale Homogeneity. Journal of Magnetism and Magnetic Materials,
324(4):501507.
[57] Feher, J., Filipcsei, G., Szalma, J., and Zrinyi, M. (2001). Bending deformation of neutral polymer gels induced by electric fields. Colloids and Surfaces A: Physicochemical
and Engineering Aspects, 183-185:505515.
[58] Fornes, T. and Paul, D. (2003). Modeling properties of nylon 6/clay nanocomposites
using composite theories. Polymer, 44(17):49935013.
[59] Fratzl, P. (2003). Cellulose and collagen: from fibres to tissues. Current Opinion in
Colloid & Interface Science, 8(1):3239.
[60] Fratzl, P. and Barth, F. G. (2009). Biomaterial systems for mechanosensing and actuation. Nature, 462(7272):4428.
[61] Fratzl, P., Dunlop, J. W. C., and Weinkamer, R. (2013). Materials Design Inspired by
Nature.
[62] Fratzl, P. and Weinkamer, R. (2007). Natures hierarchical materials. Progress in
Materials Science, 52(8):12631334.
[63] Fu, S. and Lauke, B. (1997). The fibre pull-out energy of misaligned short fibre composites. Journal of materials science, 32:19851993.
[64] Fu, S. and Lauke, B. (1998a). The Elastic Modulus of Misaligned Short-FiberReinforced Polymers. Composites science and technology, 58(97):389400.
[65] Fu, S.-Y., Feng, X.-Q., Lauke, B., and Mai, Y.-W. (2008). Effects of particle size,
particle/matrix interface adhesion and particle loading on mechanical properties of particulatepolymer composites. Composites Part B: Engineering, 39(6):933961.
[66] Fu, S.-y. and Lauke, B. (1996). Effects of fiber Length and Fiber Orientation Distributions on the Tensile Strength of Short-Fiber-Reinforced Polymers. Composites Science
and Technology, 56(2):11791190.
[67] Fu, S.-Y. and Lauke, B. (1998b). An analytical characterization of the anisotropy of
the elastic modulus of misaligned short-fiber-reinforced polymers. Composites Science
and Technology, 58(12):19611972.
[68] Fu, S.-Y., Lauke, B., Mder, E., Yue, C.-Y., and Hu, X. (2000). Tensile properties of
short-glass-fiber- and short-carbon-fiber-reinforced polypropylene composites. Composites Part A: Applied Science and Manufacturing, 31(10):11171125.
[69] Gaboriaud, F., de Gaudemaris, B., Rousseau, T., Derclaye, S., and Dufrne, Y. F.
(2012). Unravelling the nanometre-scale stimuli-responsive properties of natural rubber
latex particles using atomic force microscopy. Soft Matter, 8(9):27242729.

168

References

[70] Galipeau, E., Rudykh, S., DeBottom, G., and Ponte Castaeda, P. (2014). Magnetoactive elastomers with periodic and random microstructures. International Journal of
Solids and Structures, 51:30123024.
[71] Gao, W., Feng, X., Pei, A., Kane, C. R., Tam, R., Hennessy, C., and Wang, J. (2014).
Bioinspired helical microswimmers based on vascular plants. Nano letters, 14(1):305
10.
[72] Gasser, T. C., Ogden, R. W., and Holzapfel, G. a. (2006). Hyperelastic modelling of
arterial layers with distributed collagen fibre orientations. Journal of the Royal Society,
Interface / the Royal Society, 3(6):1535.
[73] Geers, M., Kouznetsova, V., and Brekelmans, W. (2010). Multi-scale computational
homogenization: Trends and challenges. Journal of Computational and Applied Mathematics, 234(7):21752182.
[74] Gent, A. and Thomas, A. (1958). Forms for the Stored (Strain) Energy Function for
Vulcanized Rubber. Journal of Polymer Science, 28(3):14.
[75] Gent, A. N. (1996). A New Constitutive Relation for Rubber. Rubber Chemistry and
Technology, 69(1):5961.
[76] Goettler, L., Lee, K., and Thakkar, H. (2007). Layered Silicate Reinforced Polymer
Nanocomposites: Development and Applications. Polymer Reviews, 47(2):291317.
[77] Goldberg, N., Donner, H., and Ihlemann, J. (2014). Evaluation of hyperelastic models
for unidirectional short fibre reinforced materials using a representative volume element
with refined boundary conditions. Unpublished.
[78] Gong, J. R. (2011). Graphene - Synthesis, Characterization, Properties and Applications. Intech, Croatia.
[79] Gonzlez, L., Valentn, J. L., Fernndez-Torres, A., Rodrguez, A., and MarcosFernndez, A. (2005). Effect of the network topology on the tensile strength of natural rubber vulcanizate at elevated temperature. Journal of Applied Polymer Science,
98(3):12191223.
[80] Gracias, D. H. (2013). Stimuli responsive self-folding using thin polymer films. Current Opinion in Chemical Engineering, 2(1):112119.
[81] Guidelli, E. J., Ramos, A. P., Zaniquelli, M. E. D., and Baffa, O. (2011). Green
synthesis of colloidal silver nanoparticles using natural rubber latex extracted from Hevea
brasiliensis. Spectrochimica acta. Part A, Molecular and biomolecular spectroscopy,
82(1):1405.
[82] Gumbrell, S. M., Mullins, L., and Rivlin, R. S. (1953). Departures of the elastic
behaviour of rubbers in simple extension from the kinetic theory. Transactions of the
Faraday Society, 49:14951505.
[83] Guo, Z., Caner, F., Peng, X., and Moran, B. (2008). On constitutive modelling
of porous neo-Hookean composites. Journal of the Mechanics and Physics of Solids,
56(6):23382357.

References

169

[84] Guo, Z. and Caner, F. C. (2010). Mechanical Behaviour of Transversely Isotropic


Porous Neo-Hookean Solids. International Journal of Applied Mechanics, 02(01):11
39.
[85] Guo, Z., Peng, X., and Moran, B. (2006). A composites-based hyperelastic constitutive
model for soft tissue with application to the human annulus fibrosus. Journal of the
Mechanics and Physics of Solids, 54(9):19521971.
[86] Guo, Z., Peng, X., and Moran, B. (2007a). Large deformation response of a hyperelastic fibre reinforced composite: Theoretical model and numerical validation. Composites
Part A: Applied Science and Manufacturing, 38(8):18421851.
[87] Guo, Z., Peng, X., and Moran, B. (2007b). Mechanical response of neo-Hookean fiber
reinforced incompressible nonlinearly elastic solids. International Journal of Solids and
Structures, 44(6):19491969.
[88] Guo, Z., Shi, X., Chen, Y., Chen, H., Peng, X., and Harrison, P. (2014a). Mechanical modeling of incompressible particle-reinforced neo-Hookean composites based on
numerical homogenization. Mechanics of Materials, 70:117.
[89] Guo, Z., Shi, X., Chen, Y., Chen, H., Peng, X., and Harrison, P. (2014b). Mechanical modeling of incompressible particle-reinforced neo-Hookean composites based on
numerical homogenization. Mechanics of Materials, 70:117.
[90] Halpin (1992). Primer on Composite Materials Analysis. 2nd editio edition.
[91] Halpin, J. and Kardos, J. (1976). The Halpin-Tsai equations: a review. Polymer
engineering and science, 16(5).
[92] Happel, J. and Brenner, H. (1983). Low Reynolds number hydrodynamics. Kluwer
Academic Publishers.
[93] Harrington, M. J., Gupta, H. S., Fratzl, P., and Waite, J. H. (2009). Collagen insulated
from tensile damage by domains that unfold reversibly: in situ X-ray investigation of
mechanical yield and damage repair in the mussel byssus. Journal of structural biology,
167(1):4754.
[94] Harrington, M. J., Masic, A., Holten-andersen, N., Waite, J. H., and Fratzl, P. (2010).
Iron-Clad Fibers : A Metal-Based. Science, 328(April):216220.
[95] Hart-Smith, L. (1966). Elasticity Parameters for Finite Deformations of Rubber-Like
Materials. Zeitschrift fr angewandte Mathematik und Physik, 17:608626.
[96] Hasan, M. S., Ahmed, I., Parsons, A., Walker, G., and Scotchford, C. (2012). Cytocompatibility and Mechanical Properties of Short Phosphate Glass Fibre Reinforced
Polylactic Acid (PLA) Composites: Effect of Coupling Agent Mediated Interface. Journal of functional biomaterials, 3(4):70625.
[97] Hashin, Z. and Shtrikman, S. (1962). On some variational principles in anisotropic and
nonhomogeneous elasticity. Journal of the Mechanics and Physics of Solids, 10(42):335
342.

170

References

[98] Hbaieb, K., Wang, Q., Chia, Y., and Cotterell, B. (2007). Modelling stiffness of polymer/clay nanocomposites. Polymer, 48(3):901909.
[99] Hermans, J. J. (1967). The Elastic Properties of Fiber Reinforced Materials when the
Fibers are Aligned. Proc. Roy. Academy Amsterdam, B70:1.
[100] Hernndez, M., Bernal, M. D. M., Verdejo, R., Ezquerra, T. a., and Lpez-Manchado,
M. a. (2012). Overall performance of natural rubber/graphene nanocomposites. Composites Science and Technology, 73:4046.
[101] Hill, R. (1965). A Self-Consistent Mechanics of Composite Materials. Journal of the
Mechanics and Physics of Solids, 13(March 1962).
[102] Hill, R. (1972). On Constitutive Macro-Variables for Heterogeneous Solids at Finite
Strain. Proceedings of the Royal Society of London. Series A, Mathematical and Physical
Sciences, 326(1565):131147.
[103] Hill, R. and Rice, J. R. (1973). Elastic potentials and the structure of inelastic constitutive laws. SIAM Journal on Applied Mathematics, 25(3):448461.
[104] Holzapfel, G., Gasser, T., and Ogden, R. (2000). A new constitutive framework
for arterial wall mechanics and a comparative study of material models. Computational
Biomechanics, (8).
[105] Horgan, C. O. and Murphy, J. G. (2011). On the Modeling of Extension-Torsion Experimental Data for Transversely Isotropic Biological Soft Tissues. Journal of Elasticity,
108(2):179191.
[106] Horgan, C. O. and Saccomandi, G. (2005). A new constitutive theory for fiberreinforced incompressible nonlinearly elastic solids. Journal of the Mechanics and
Physics of Solids, 53(9):19852015.
[107] Hu, H., Onyebueke, L., and Abatan, A. (2010). Characterizing and modeling mechanical properties of nanocomposites-review and evaluation. Journal of Minerals and
Materials, 9(4):275319.
[108] III, C. T. and Liang, E. (1999). Stiffness predictions for unidirectional short-fiber
composites: Review and evaluation. Composites science and technology, 59:655671.
[109] Iijima, S. (1991). Helical microtubules of graphitic carbon. Nature, 354:5658.
[110] Ishai, O., Weller, T., and Singer, J. (1968). Anisotropy of Mylar A Sheets. ASTM
Journal of Materials, 3(2):337351.
[111] Ishikawa, S., Tokuda, A., and Kotera, H. (2008). Numerical Simulation for Fibre
Reinforced Rubber. Journal of Computational Science and Technology, 2(4):587596.
[112] Jang, B. and Sakka, Y. (2009). Influence of shape and size on the alignment of multiwall carbon nanotubes under magnetic fields. Materials Letters, 63(29):25452547.
[113] Jha, V. and Hon, A. (2008). Modeling of the effect of rigid fillers on the stiffness of
rubbers. Journal of applied polymer science, 107(August):25722577.

References

171

[114] Jin, L., Bower, C., and Zhou, O. (1998). Alignment of carbon nanotubes in a polymer
matrix by mechanical stretching. Applied Physics Letters, 73(9):1197.
[115] Jones, R. (1975). Mechanics of Composite Materials. 2nd edtion edition.
[116] Joshi, U. a., Sharma, S. C., and Harsha, S. (2012). Effect of carbon nanotube orientation on the mechanical properties of nanocomposites. Composites Part B: Engineering,
43(4):20632071.
[117] Julio Garca Ruz, M. and Yarime Surez Gonzlez, L. (2006). Comparison of hyperelastic material models in the analysis of fabrics. International Journal of Clothing
Science and Technology, 18(5):314325.
[118] Kacir, L., Narkis, M., and Ishai, O. (1975). Oriented short glass-fiber composites. I.
Preparation and statistical analysis of aligned fiber mats. Polymer Engineering & Science,
15(7).
[119] Kacir, L., Narkis, M., and Ishai, O. (1978). Aligned short glass fibre/epoxy composites. Composites, 9(2):8992.
[120] Kalaitzidou, K., Fukushima, H., and Drzal, L. T. (2007). Multifunctional polypropylene composites produced by incorporation of exfoliated graphite nanoplatelets. Carbon,
45(7):14461452.
[121] Kaliske, M. and Heinrich, G. (1999). An Extended Tube-Model for Rubber Elasticity:
Statistical-Mechanical Theory and Finite Element Implementation. Rubber chemistry
and technology, 72(4):602632.
[122] Kameyama, M. and Fukunaga, H. (2007). Optimum design of composite plate wings
for aeroelastic characteristics using lamination parameters. Computers & Structures,
85(3-4):213224.
[123] Kessler, M., Sottos, N., and White, S. (2003). Self-healing structural composite
materials. Composites Part A: Applied Science and Manufacturing, 34(8):743753.
[124] Kilian, H. G., Enderle, H. F., and Unseld, K. (1986). The use of the van der Waals
model to elucidate universal aspects of structure-property relationships in simply extended dry and swollen rubbers. Colloid and Polymer Science, 264(10):866876.
[125] Kim, B. C., Potter, K., and Weaver, P. M. (2012a). Continuous tow shearing for
manufacturing variable angle tow composites. Composites Part A: Applied Science and
Manufacturing, 43(8):13471356.
[126] Kim, I. T., Tannnenbaum, A., and Tannenbaum, R. (2012b). Anisotropic conductivity
of magnetic carbon nanotubes embedded in epoxy matrices. Carbon, 49(1):5461.
[127] Kim, J., Chung, S. E., Choi, S.-E., Lee, H., Kim, J., and Kwon, S. (2011). Programming magnetic anisotropy in polymeric microactuators. Nature materials, 10(10):747
52.
[128] Kim, J.-H. and Jeong, H.-Y. (2005). A study on the material properties and fatigue
life of natural rubber with different carbon blacks. International Journal of Fatigue,
27:263272.

172

References

[129] Kimura, T., Ago, H., Tobita, M., Ohshima, S., Kyotani, M., and Yumura, M. (2002).
Polymer Composites of Carbon Nanotubes Aligned by a Magnetic Field. Advanced
Materials, 14(19):13801383.
[130] Kimura, T., Goto, T., Shintani, H., and Ishizaka, K. (2003). Magnetic control of
ferroelectric polarization. Nature, 426(November):5558.
[131] Kimura, T., Umehara, Y., and Kimura, F. (2010). Fabrication of a short carbon
fiber/gel composite that responds to a magnetic field. Carbon, 48(14):40154018.
[132] Kimura, T., Umehara, Y., and Kimura, F. (2012). Magnetic field responsive silicone elastomer loaded with short steel wires having orientation distribution. Soft Matter,
8(23):6206.
[133] Kimura, T., Yoshino, M., Yamane, T., Yamato, M., and Tobita, M. (2004). Uniaxial Alignment of the Smallest Diamagnetic Susceptibility Axis Using Time-Dependent
Magnetic Fields. Langmuir, 20:56695672.
[134] Kitamura, N., Fukumi, K., Takahashi, K., Mogi, I., Awaji, S., and Watanabe, K.
(2011). Orientation of Carbon Nano-fiber in Carbon/Silica Composite Prepared under High Magnetic Field. IOP Conference Series: Materials Science and Engineering,
18(5):052008.
[135] Kohlmeyer, R. R. and Chen, J. (2013). Wavelength-selective, IR light-driven hinges
based on liquid crystalline elastomer composites. Angewandte Chemie (International ed.
in English), 52(35):92347.
[136] Koo, J. H. (2006). Polymer Nanocomposites: Processing, Characterization, and
Applications. McGraw-Hill.
[137] Kouznetsova, V. (2001). An approach to micro-macro modeling of heterogeneous
materials. Computational Mechanics, 27(July 2000).
[138] Kouznetsova, V. (2002). Computational homogenization for the multi-scale analysis
of multi-phase materials.
[139] Kuilla, T., Bhadra, S., Yao, D., Kim, N. H., Bose, S., and Lee, J. H. (2010). Recent advances in graphene based polymer composites. Progress in Polymer Science,
35(11):13501375.
[140] Lachenal, X., Daynes, S., and Weaver, P. (2013). Review of morphing concepts and
materials for wind turbine blade applications. Wind Energy, 16(February 2012):283307.
[141] Laiarinandrasana, L., Jean, A., Jeulin, D., and Forest, S. (2012). Modelling the effects
of various contents of fillers on the relaxation rate of elastomers. Materials & Design,
33:7582.
[142] Lakes, R. (2009). Viscoelastic Materials. Cambridge University Press.
[143] Laoui, T. (2013). Mechanical and Thermal Properties of Styrene Butadiene Rubber Functionalized Carbon Nanotubes Nanocomposites. Fullerenes, Nanotubes and Carbon
Nanostructures, 21(2):89101.

References

173

[144] Larsen, J., Jenkins, C., Woo, K., and Denowh, C. (2009). A Bio-Inspired Lightweight
MRF-Foam Actuator. 50th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, (May).
[145] Laws, N. and McLaughlin, R. (1979). The Effect of Fibre Length on the Overall
Moduli of Composite Materials. Journal of the Mechanics and Physics of Solids, 27:1
13.
[146] Le Cam, J.-B. and Toussaint, E. (2010). The Mechanism of Fatigue Crack Growth
in Rubbers under Severe Loading: the Effect of Stress-Induced Crystallization. Macromolecules, 43(10):47084714.
[147] Leng, J. and t-k. Lau, A., editors (2011). Multifunctional Polymer Nanocomposites.
CRC Press.
[148] Li, B., Olson, E., Perugini, A., and Zhong, W. (2011). Simultaneous enhancements in
damping and static dissipation capability of polyetherimide composites with organosilane
surface modified graphene nanoplatelets. Polymer, 52:56065614.
[149] Li, B. and Zhong, W.-H. (2011). Review on polymer/graphite nanoplatelet nanocomposites. Journal of Materials Science, 46(17):55955614.
[150] Li, H. and Bubeck, C. (2013). Photoreduction processes of graphene oxide and related applications. Macromolecular Research, 21(3):290297.
[151] Li, W. H., Zhang, X. Z., and Du, H. (2013). Magnetorheological elastomers and their
applications. In Advances in Elastomers I: Blends and Interpenetrating Networks, pages
357374.
[152] Li, Y., Hu, Z., and Chen, Y. (1997). Shape Memory Gels Made by the Modulated Gel
Technology. Journal of applied polymer science, 63:11731178.
[153] Libanori, R., Erb, R. M., Reiser, A., Le Ferrand, H., Sess, M. J., Spolenak, R., and
Studart, A. R. (2012a). Stretchable heterogeneous composites with extreme mechanical
gradients. Nature communications, 3:1265.
[154] Libanori, R., Erb, R. M., and Studart, A. R. (2013). Mechanics of platelet-reinforced
composites assembled using mechanical and magnetic stimuli. ACS applied materials &
interfaces, 5(21):10794805.
[155] Libanori, R., Mnch, F. H., Montenegro, D. M., and Studart, A. R. (2012b). Hierarchical reinforcement of polyurethane-based composites with inorganic micro- and
nanoplatelets. Composites Science and Technology, 72(3):435445.
[156] Lin, Y., Zhang, A., Wang, L., Pei, C., and Gu, Q. (2012). Carbon Black Filled
Powdered Natural Rubber. Journal of Macromolecular Science, Part B, 51(7):1267
1281.
[157] Liu, H. and Brinson, L. C. (2008). Reinforcing efficiency of nanoparticles: A
simple comparison for polymer nanocomposites. Composites Science and Technology,
68(6):15021512.

174

References

[158] Lopez-Pamies, O., Goudarzi, T., and Nakamura, T. (2013). The nonlinear elastic
response of suspensions of rigid inclusions in rubber: IAn exact result for dilute suspensions. Journal of the Mechanics and Physics of Solids, 61(1):118.
[159] Lopez-Pamies, O. and Idiart, M. I. (2010). Fiber-reinforced hyperelastic solids: a
realizable homogenization constitutive theory. Journal of Engineering Mathematics,
68(1):5783.
[160] Lopez-Pamies, O., Idiart, M. I., and Li, Z. (2011). On Microstructure Evolution
in Fiber-Reinforced Elastomers and Implications for Their Mechanical Response and
Stability. Journal of Engineering Materials and Technology, 133(1):011007.
[161] Lopez-Pamies, O. and Ponte Castaeda, P. (2009). Microstructure evolution in hyperelastic laminates and implications for overall behavior and macroscopic stability. Mechanics of Materials, 41(4):364374.
[162] Lubeck, D. P. and Bunker, J. P. (1981). The Artifical Heart: Costs, Risks and Benefits,
volume 46. DIANE Publishing.
[163] Madkour, T. (2004). Step-Strain Stress Relaxation of Carbon Black-Loaded Natural
Rubber Vulcanizates. Journal of applied polymer science, 92(5).
[164] Malas, A., Das, C. K., Das, A., and Heinrich, G. (2012). Development of expanded
graphite filled natural rubber vulcanizates in presence and absence of carbon black: Mechanical, thermal and morphological properties. Materials & Design, 39:410417.
[165] Marckmann, G. and Verron, E. (2006). Comparison of Hyperelastic Models for
Rubber-Like Materials. Rubber Chemistry and Technology, 79(5):835858.
[166] Mars, W. and Fatemi, A. (2002). A literature survey on fatigue analysis approaches
for rubber. International Journal of Fatigue, 24(9):949961.
[167] Martin, C., Sandler, J., Windle, A., Schwarz, M.-K., Bauhofer, W., Schulte, K., and
Shaffer, M. (2005). Electric field-induced aligned multi-wall carbon nanotube networks
in epoxy composites. Polymer, 46(3):877886.
[168] Martone, P. T., Boller, M., Burgert, I., Dumais, J., Edwards, J., Mach, K., Rowe,
N., Rueggeberg, M., Seidel, R., and Speck, T. (2010). Mechanics without muscle:
biomechanical inspiration from the plant world. Integrative and comparative biology,
50(5):888907.
[169] Masoud, H., Bingham, B. I., and Alexeev, A. (2012). Designing maneuverable microswimmers actuated by responsive gel. Soft Matter, 8(34):8944.
[170] Meeker, D. C. (2006). Finite Element Method Magnetics, Version 4.0.1.
[171] Merodio, J. and Ogden, R. (2005). Mechanical response of fiber-reinforced incompressible non-linearly elastic solids. International Journal of Non-Linear Mechanics,
40(2-3):213227.
[172] Meyers, M. A., Chen, P.-Y., Lin, A. Y.-M., and Seki, Y. (2008). Biological materials:
Structure and mechanical properties. Progress in Materials Science, 53(1):1206.

References

175

[173] Min, C. and Yu, D. (2010). Simultaneously improved toughness and dielectric
properties of epoxy/graphite nanosheet composites. Polymer Engineering & Science,
50(9):17341742.
[174] Mlekusch, B., Lehner, E., and Geymayer, W. (1999). Fibre orientation in shortfibre-reinforced thermoplastics I . Contrast enhancement for image analysis. Composites
Science and Technology, 59:543545.
[175] Mooney, M. (1940). A Theory of Large Elastic Deformation. Journal of Applied
Physics, 11(9):582.
[176] Moraleda, J., Segurado, J., and LLorca, J. (2009). Finite deformation of incompressible fiber-reinforced elastomers: A computational micromechanics approach. Journal of
the Mechanics and Physics of Solids, 57(9):15961613.
[177] Mori, T. and Tanaka, K. (1973). Average stress in matrix and average elastic energy
of materials with misfitting inclusions. Acta Metallurgica, 21(5):571574.
[178] Mller, U., Gindl, W., and Jeronimidis, G. (2006). Biomechanics of a branch stem
junction in softwood. Trees, 20(5):643648.
[179] Munch, E., Launey, M. E., Alsem, D. H., Saiz, E., Tomsia, A. P., and Ritchie, R. O.
(2008). Tough, Bio-Inspired Hybrid Materials. Science, 322(December):15161520.
[180] Mura, T. (1987). Micromechanics of Defects in Solids. Second, re edition.
[181] Ning, X., Zhu, Q., Lanir, Y., and Margulies, S. S. (2006). A transversely isotropic
viscoelastic constitutive equation for brainstem undergoing finite deformation. Journal
of biomechanical engineering, 128(6):92533.
[182] Novoselov, K. S., Geim, A. K., Morozov, S. V., and Jiang, D. (2004). Electric Field
Effect in Atomically Thin Carbon Films. Science, 306(October):666669.
[183] Ogden, R. (1972). Large Deformation Isotropic Elasticity - on the Correlation of
Theory and Experiment for Incompressible Rubberlike Solids. Proceedings of the Royal
Society A: Mathematical, Physical and Engineering Sciences, 326:565584.
[184] Ogden, R. W. (1997). Non-linear Elastic Deformations. Dover Publications, New
York.
[185] Ogden, R, W. (1978). Extremum principles in non-linear elasticity and their application to compositesI: Theory. International Journal of Solids and Structures,
14(4):265282.
[186] Ozbas, B., Toki, S., Hsiao, B. S., Chu, B., Register, R. A., Aksay, I. A., Prudhomme,
R. K., and Adamson, D. H. (2012). Strain-induced crystallization and mechanical properties of functionalized graphene sheet-filled natural rubber. Journal of Polymer Science
Part B: Polymer Physics, 50(10):718723.
[187] Pandolfi, A. and Vasta, M. (2012). Fiber distributed hyperelastic modeling of biological tissues. Mechanics of Materials, 44:151162.

176

References

[188] Papadopoulos, I. C., Thomas, A. G., and Busfield, J. J. C. (2008). Rate Transitions
in the Fatigue Crack Growth of Elastomers. Journal of applied polymer science, 109(3).
[189] Patil, A. J., Vickery, J. L., Scott, T. B., and Mann, S. (2009). Aqueous Stabilization
and Self-Assembly of Graphene Sheets into Layered Bio-Nanocomposites using DNA.
Advanced Materials, 21(31):31593164.
[190] Peckett, J. W., Trens, P., Gougeon, R. D., Poppl, A., Harris, K., and Hudson, M. J.
(2000). Electrochemically oxidised graphite . Characterisation and some ion exchange
properties. Carbon, 38:345353.
[191] Peng, X. Q., Guo, Z. Y., and Moran, B. (2006). An Anisotropic Hyperelastic Constitutive Model With Fiber-Matrix Shear Interaction for the Human Annulus Fibrosus.
Journal of Applied Mechanics, 73(5):815.
[192] Peng, Z., Feng, C., Luo, Y., Li, Y., and Kong, L. (2010). Self-assembled natural
rubber/multi-walled carbon nanotube composites using latex compounding techniques.
Carbon, 48:44974503.
[193] Peyer, K. E., Zhang, L., and Nelson, B. J. (2013). Bio-inspired magnetic swimming
microrobots for biomedical applications. Nanoscale, 5(4):125972.
[194] Polymers, T., Series, A. C. S. S., and Society, A. C. (1982). Chemorheology of Thermosetting Polymers, volume 227 of ACS Symposium Series. AMERICAN CHEMICAL
SOCIETY, WASHINGTON, D.C.
[195] Ponte-Castaneda, P. (1989). The Overall Constitutive Behaviour of Nonlinearly Elastic Composites. Pathematical physics - Proceedings of the Royal Society A, 422(1862).
[196] Potts, J. R., Shankar, O., Du, L., and Ruoff, R. S. (2012).
ProcessingMorphologyProperty Relationships and Composite Theory Analysis of Reduced
Graphene Oxide/Natural Rubber Nanocomposites. Macromolecules, 45(15):60456055.
[197] Prolongo, S., Meliton, B., Del Rosario, G., and Urea, A. (2013). New alignment
procedure of magnetite-CNT hybrid nanofillers on epoxy bulk resin with permanent magnets. Composites Part B: Engineering, 46:166172.
[198] Qian, D., Dickey, E. C., Andrews, R., and Rantell, T. (2000). Load transfer and
deformation mechanisms in carbon nanotube-polystyrene composites. Applied Physics
Letters, 76(20):2868.
[199] Qiu, G. and Pence, T. (1997). Remarks on the behavior of simple directionally reinforced incompressible nonlinearly elastic solids. Journal of Elasticity, 49:130.
[200] Ramanathan, T., Liu, H., and Brinson, L. C. (2005). Functionalized SWNT/polymer
nanocomposites for dramatic property improvement. Journal of Polymer Science Part B:
Polymer Physics, 43(17):22692279.
[201] Ramanathan, T. and Stankovich, S. (2007). Graphitic nanofillers in PMMA nanocompositesan investigation of particle size and dispersion and their influence on nanocomposite properties. Journal of Polymer Science, 45:20972112.

References

177

[202] Ramanujan, R. V. and Lao, L. L. (2006). The mechanical behavior of smart magnethydrogel composites. Smart Materials and Structures, 15(4):952956.
[203] Ramorino, G., Bignotti, F., Pandini, S., and Ricc, T. (2009). Mechanical reinforcement in natural rubber/organoclay nanocomposites. Composites Science and Technology,
69(7-8):12061211.
[204] Rattanasom, N. and Prasertsri, S. (2009). Relationship among mechanical properties,
heat ageing resistance, cut growth behaviour and morphology in natural rubber: Partial
replacement of clay with various types of carbon black at similar hardness level. Polymer
Testing, 28:270276.
[205] Rault, J., Marchal, J., Judeinstein, P., and Albouy, P. a. (2006). Chain orientation in
natural rubber, Part II: 2H-NMR study. The European physical journal. E, Soft matter,
21(3):24361.
[206] Rezende, C. a., Bragana, F. C., Doi, T. R., Lee, L.-T., Galembeck, F., and Bou,
F. (2010). Natural rubber-clay nanocomposites: Mechanical and structural properties.
Polymer, 51(16):36443652.
[207] Rikken, R. S. M., Nolte, R. J. M., Maan, J. C., van Hest, J. C. M., Wilson, D. a.,
and Christianen, P. C. M. (2014). Manipulation of micro- and nanostructure motion with
magnetic fields. Soft matter, 10(9):1295308.
[208] Rivlin, R. (1948). Large Elastic Deformations of Isotropic Materials. IV. Further
Developments of the General Theory. The royal society Publishing, 241(835):379397.
[209] Rivlin, R. S. and Saunders, D. W. (1951). Large Elastic Deformations of Isotropic
Materials. VII. Experiments on the Deformation of Rubber. Philosophical Transactions
of the Royal Society of London. Series A, Mathematical and Physical Sciences, 243(865).
[210] Roller, M. (1975). Characterization of the Time-Temperature-Viscosity Behavior of
Curing B-Staged Epoxy Resin. Polymer Engineering & Science, 15(6):406414.
[211] Rose, K., Meier, J., Dougherty, G., and Santiago, J. (2007a). Rotational electrophoresis of striped metallic microrods. Physical Review E, 75(1):011503.
[212] Rose, K., Meier, J., Dougherty, G., and Santiago, J. (2007b). Rotational electrophoresis of striped metallic microrods. Physical Review E, 75(1):011503.
[213] Rudykh, S. and Bertoldi, K. (2013). Stability of anisotropic magnetorheological elastomers in finite deformations: A micromechanical approach. Journal of the Mechanics
and Physics of Solids, 61(4):949967.
[214] Rudykh, S., Lewinstein, A., Uner, G., and DeBotton, G. (2013). Analysis of microstructural induced enhancement of electromechanical coupling in soft dielectrics. Applied Physics Letters, 102(15):151905.
[215] Ruiz-Molina, D., Novio, F., and Roscini, C. (2015). Bio- and Bioinspired Nanomaterials.

178

References

[216] Saintier, N., Cailletaud, G., and Piques, R. (2011). Cyclic loading and crystallization
of Natural Rubber: An explanation of fatigue crack propagation reinforcement under a
positive loading ratio. Materials Science and Engineering: A, 528:10781086.
[217] Sassi, S., Cherif, K., Mezghani, L., Thomas, M., and Kotrane, A. (2005). An innovative magnetorheological damper for automotive suspension: from design to experimental
characterization. Smart Materials and Structures, 14(4):811822.
[218] Schmitt, C. N. Z., Winter, A., Bertinetti, L., Masic, A., Strauch, P., and Harrington,
M. J. (2015). Mechanical homeostasis of a DOPA- enriched biological coating from
mussels in response to metal variation. Interface, 12(110).
[219] Scholz, M.-S., Drinkwater, B. W., and Trask, R. S. (2014). Ultrasonic assembly of
anisotropic short fibre reinforced composites. Ultrasonics, 54(4):10159.
[220] Sharma, A., Tripathi, B., and Vijay, Y. (2010). Dramatic Improvement in properties
of magnetically aligned CNT/polymer nanocomposites. Journal of Membrane Science,
361(1-2):8995.
[221] Sheng, N., Boyce, M., Parks, D., Rutledge, G., Abes, J., and Cohen, R. (2004).
Multiscale micromechanical modeling of polymer/clay nanocomposites and the effective
clay particle. Polymer, 45(2):487506.
[222] Shi, D., He, P., Zhao, P., Guo, F. F., Wang, F., Huth, C., Chaud, X., Budko, S. L., and
Lian, J. (2011). Magnetic alignment of Ni/Co-coated carbon nanotubes in polystyrene
composites. Composites Part B: Engineering, 42(6):15321538.
[223] Silva, C. A., Viana, C., Hattum, F. W. J. V., and Cunha, M. (2006). Fiber Orientation in Divergent / Convergent Flows in Expansion and Compression Injection Molding.
Polymer Composites, 27(5):539551.
[224] Song, K., Yeom, E., Seo, S.-J., Kim, K., Kim, H., Lim, J.-H., and Joon Lee, S. (2015).
Journey of water in pine cones. Scientific reports, 5:9963.
[225] Stanier, D., Ciambella, J., and Rahatekar, S. S. (2016). Bending and Twisting Actuation of Nickel coated Carbon Fibres and Elastomer composites using Low Magnetic
Field. Composites Part A (Under Review).
[226] Stanier, D., Patil, A., Sriwong, C., Rahatekar, S., and Ciambella, J. (2014). The
reinforcement effect of exfoliated graphene oxide nanoplatelets on the mechanical and
viscoelastic properties of natural rubber. Composites Science and Technology, 95:5966.
[227] Stanier, D. C. and Rahatekar, S. (2015). Magnetic Response of aligned nickel coated
carbon fibres in a PDMS matrix. In ECCMR IX, number Sylgard 184.
[228] Studart, A. R. and Erb, R. M. (2014). Bioinspired materials that self-shape through
programmed microstructures. Soft matter, 10(9):128494.
[229] Suk, J. W., Piner, R. D., An, J., and Ruoff, R. S. (2010). Mechanical properties of
monolayer graphene oxide. ACS nano, 4(11):655764.

References

179

[230] Szabo, D., Szeghy, G., and Zrinyi, M. (1998). Shape transition of magnetic field
sensitive polymer gels. Macromolecules, 31:65416548.
[231] Takahashi, T., Yonetake, K., Koyama, K., and Kikuchi, T. (2003). Polycarbonate
Crystallization by Vapor-Grown Carbon Fiber with and without Magnetic Field. Macromolecular Rapid Communications, 24(13):763767.
[232] Takeyama, S., Nakamura, S., and Uchida, K. (2006). Dynamical orientation of carbon
nanotubes by pulsed magnetic fields. Journal of Physics: Conference Series, 51:446
449.
[233] Tekinalp, H. L., Kunc, V., Velez-Garcia, G. M., Duty, C. E., Love, L. J., Naskar, A. K.,
Blue, C. a., and Ozcan, S. (2014). Highly oriented carbon fiberpolymer composites via
additive manufacturing. Composites Science and Technology, 105:144150.
[234] Thill, C., Etches, J., Bond, I., Potter, K., and Weaver, P. (2008). Morphing skins. The
Aeronautical Journal, 112(3216):123.
[235] Thomson, J. (1922). The Outline of Science - A Plain Story Simple Told. G. P.
Putnams Sons, New York and London.
[236] Tipton, C. R., Han, E., and Mullin, T. (2012). Magneto-elastic buckling of a soft
cellular solid. Soft Matter, 8(26):68806883.
[237] Toki, S., Sics, I., Ran, S., Liu, L., Hsiao, B. S., Murakami, S., Senoo, K., and Kohjiya,
S. (2002). New Insights into Structural Development in Natural Rubber during Uniaxial
Deformation by In Situ Synchrotron X-ray Diffraction. Macromolecules, 35(17):6578
6584.
[238] Tondu, B., Emirkhanian, R., Math, S., and Ricard, A. (2009). A pH-activated artificial muscle using the McKibben-type braided structure. Sensors and Actuators A:
Physical, 150(1):124130.
[239] Tosaka, M. (2007). Strain-Induced Crystallization of Crosslinked Natural Rubber
As Revealed by X-ray Diffraction Using Synchrotron Radiation. Polymer Journal,
39(12):12071220.
[240] Tosaka, M., Murakami, S., Poompradub, S., Kohjiya, S., Ikeda, Y., Toki, S., Sics, I.,
and Hsiao, B. S. (2004). Orientation and Crystallization of Natural Rubber Network As
Revealed by WAXD Using Synchrotron Radiation. Macromolecules, 37(9):32993309.
[241] Treloar, L. R. G. (1944). The Physics of Rubber Elasticity. Oxford University Press,
3rd edition.
[242] Varga, Z., Filipcsei, G., Szilagyi, A., and Zrinyi, M. (2005a). Electric and Magnetic
Field-Structured Smart Composites. Macromolecular Symposia, 227(1):123134.
[243] Varga, Z., Filipcsei, G., and Zrinyi, M. (2005b). Smart composites with controlled
anisotropy. Polymer, 46:77797787.
[244] Varga, Z., Filipcsei, G., and Zrinyi, M. (2006). Magnetic field sensitive functional
elastomers with tuneable elastic modulus. Polymer, 47(1):227233.

180

References

[245] Vidyashankar, B. R. and Murty, A. V. K. (2001). Analysis of laminates with ply


drops. Composites Science and Technology, 61:749758.
[246] Walpole, L. (1966). On the bounds for the overall elastic moduli for inhomogeneous
systems II. Journal of the Mechanics and Physics of Solids, 14:289301.
[247] Wang, H., Zhou, H., Peng, R., and Mishnaevsky, L. (2011). Nanoreinforced polymer
composites: 3D FEM modeling with effective interface concept. Composites Science
and Technology, 71(7):980988.
[248] Wang, Q., Dai, J., Li, W., Wei, Z., and Jiang, J. (2008). The effects of CNT alignment
on electrical conductivity and mechanical properties of SWNT/epoxy nanocomposites.
Composites Science and Technology, 68(7-8):16441648.
[249] Weiss, J., Maker, B., and Govindjee, S. (1996). Finite Element Implementation of
Incompressible, Transversely Isotropic Hyperelasticity. Computer Methods in Applied
Mechanics and Engineering, 135(1-2):107128.
[250] Weng, G., Huang, G., Qu, L., Nie, Y., and Wu, J. (2010). Large-Scale Orientation in
a Vulcanized Stretched Natural Rubber Network: Proved by In-Situ Synchrotron X-ray
Diffraction Characterization. The Journal of Physical Chemistry B, 114:71797188.
[251] Wetherhold, R. and Jain, L. (1993). The effect of crack orientation on the fracture
properties of composite materials. Materials Science and Engineering: A, 165:9197.
[252] Wiley, J. (2011). Properties and Behavior of Polymers.
[253] Winkelstein, B. A. (2013). Orthopaedic Biomechanics.
[254] Wu, T., Su, Y., and Chen, B. (2014). Mechanically adaptive and shape-memory
behaviour of chitosan-modified cellulose whisker/elastomer composites in different pH
environments. Chemphyschem : a European journal of chemical physics and physical
chemistry, 15(13):2794800.
[255] Xia, Z., Zhang, Y., and Ellyin, F. (2003). A unified periodical boundary conditions for
representative volume elements of composites and applications. International Journal of
Solids and Structures, 40(8):19071921.
[256] Xiao, Y., Kawai, M., and Hatta, H. (2010). An integrated method for off-axis tension
and compression testing of unidirectional composites. Journal of Composite Materials,
45(6):657669.
[257] Yadav, S. K., Kim, I. J., Kim, H. J., Kim, J., Hong, S. M., and Koo, C. M. (2013).
PDMS/MWCNT nanocomposite actuators using silicone functionalized multiwalled carbon nanotubes via nitrene chemistry. Journal of Materials Chemistry C, 1(35):5463.
[258] Yamaguchi, M. and Tanimoto, Y. (2006). Magneto-science. Magnetic Field Effects
on Materials: Fundamentals and Applications.
[259] Yan, N., Xia, H., Zhan, Y., and Fei, G. (2013). New Insights into Fatigue Crack
Growth in Graphene-Filled Natural Rubber Composites by Microfocus Hard-X-Ray
Beamline Radiation. Macromolecular Materials and Engineering, 298(1):3844.

References

181

[260] Yang, J., Tian, M., Jia, Q.-X., Shi, J.-H., Zhang, L.-Q., Lim, S.-H., Yu, Z.-Z., and
Mai, Y.-W. (2007). Improved mechanical and functional properties of elastomer/graphite
nanocomposites prepared by latex compounding. Acta Materialia, 55(18):63726382.
[261] Yeoh, O. H. (1993). Some Forms of the Strain Energy Function for Rubber. Rubber
chemistry and technology, 66(5):754771.
[262] Young, R. J., Kinloch, I. a., Gong, L., and Novoselov, K. S. (2012). The mechanics of
graphene nanocomposites: A review. Composites Science and Technology, 72(12):1459
1476.
[263] Yu, W., Chen, X., Wang, Y., Yan, L., and Bai, N. (2008). Uniaxial ratchetting behavior of vulcanized natural rubber. Polymer Engineering & Science, 48(1):191197.
[264] Zhan, Y., Lavorgna, M., Buonocore, G., and Xia, H. (2012). Enhancing electrical conductivity of rubber composites by constructing interconnected network of selfassembled graphene with latex mixing. Journal of Materials Chemistry, 22(21):10464.
[265] Zhan, Y., Wu, J., Xia, H., Yan, N., Fei, G., and Yuan, G. (2011). Dispersion and
Exfoliation of Graphene in Rubber by an Ultrasonically-Assisted Latex Mixing and In
situ Reduction Process. Macromolecular Materials and Engineering, 296(7):590602.
[266] Zhang, Z., Chen, D., Wu, H., Bao, Y., and Chai, G. (2016). Non-contact magnetic
driving bioinspired Venus flytrap robot based on bistable anti-symmetric CFRP structure.
Composite Structures, 135:1722.
[267] Zrinyi, M. (2000). Intelligent polymer gels controlled by magnetic fields. Colloid
and Polymer Science, 278:98103.
[268] Zrinyi, M., Barsi, L., and Buki, A. (1996). Deformation of ferrogels induced by
nonuniform magnetic fields. The Journal of Chemical Physics, 104(21):8750.
[269] Zrinyi, M., Barsi, L., Szabo, D., and Kilian, H.-G. (1997). Direct observation of
abrupt shape transition in ferrogels induced by nonuniform magnetic field. The Journal
of Chemical Physics, 106(13):5685.
[270] Zrinyi, M., Szab, D., and Kilian, H. (1998). Kinetics of the shape change of magnetic field sensitive polymer gels. Polymer Gels and Networks, 6:441454.

Appendix A
Supplementary Material - Chapter 2
A.1

The Stress-Strain Curve for Reinforced PDMS

Natural Rubber shows an S shaped stress-strain curve (Fig. 2.9a). This behaviour is similarly observed in PDMS, although the S shape is weaker due to the limited extensibility of
the elastomer chains. In Section 2.3, the nominal stress-strain behaviour of PDMS is shown
to 60% strain. Here, we display the full stress-strain behaviour.

Nominal Stress (MPa)

0.8

0.6

00
15
30
45
60
75
90

0.4

0.2

0
1

1.2

1.4

1.6

1.8

2.2

Stretch,
Fig. A.1 The stress-strain behaviour of 7 selected specimens of NiC fibre reinforced PDMS
samples, shown up to mechanical failure.

Appendix B
Supplementary Material - Chapter 3
B.1

Comparison of the RVE to the Constitutive Model Plane Stress

The capabilities of the model are assessed by comparison to the numerical model in Fig. 3.18
(Section 3.4.5). The vs. strain plane, and the stress-strain curves are compared for
the angles 0 = {0 , 45 , 75 }. Here, we supplement the data by including the remaining
angular configurations.
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 5
1

0.5

0
60

20

20

Strain (%)

40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

(a)
40

11 , 12 (MPa)

1.5

(b)
60

2
60

40

20

20

40

60

Strain (%)

Fig. B.1 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The plots
represent (a)-(b) 0 = 5 . The other values were (r) = 2 MPa and f = 1000.

186

Supplementary Material - Chapter 3

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 10
1

0.5

0
60

11 , 12 (MPa)

1.5

20

20

40

2
1
0
1

(a)
40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

(b)
2
60

60

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 15
1

0.5

0
60

20

20

40

3
2
1
0

2
60

60

40

20

20

40

60

Strain (%)

0.5

11 , 12 (MPa)

Strain Energy,

60

(d)

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 20

40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

20

(c)
40

Strain (%)

11 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

0
60

(f)

(e)
40

20

20

Strain (%)

40

60

2
60

40

20

20

40

60

Strain (%)

Fig. B.2 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 10 , (c)-(d) 0 = 15 , and (e)-(f) 0 = 20 . The other values
were (r) = 2 MPa and f = 1000.

187

B.1 Comparison of the RVE to the Constitutive Model - Plane Stress

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 25
1

0.5

0
60

11 , 12 (MPa)

1.5

20

20

40

2
1
0
1

(a)
40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

(b)
2
60

60

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 30
1

0.5

0
60

20

20

40

3
2
1
0

2
60

60

40

20

20

40

60

Strain (%)

0.5

11 , 12 (MPa)

Strain Energy,

60

(d)

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 35

40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

20

(c)
40

Strain (%)

11 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

0
60

(f)

(e)
40

20

20

Strain (%)

40

60

2
60

40

20

20

40

60

Strain (%)

Fig. B.3 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 25 , (c)-(d) 0 = 30 , and (e)-(f) 0 = 35 . The other values
were (r) = 2 MPa and f = 1000.

188

Supplementary Material - Chapter 3

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 40
1

0.5

0
60

11 , 12 (MPa)

1.5

20

20

40

2
1
0
1

(a)
40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

(b)
2
60

60

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 50
1

0.5

0
60

20

20

40

3
2
1
0

2
60

60

40

20

20

40

60

Strain (%)

0.5

11 , 12 (MPa)

Strain Energy,

60

(d)

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 55

40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

20

(c)
40

Strain (%)

11 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

0
60

(f)

(e)
40

20

20

Strain (%)

40

60

2
60

40

20

20

40

60

Strain (%)

Fig. B.4 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 40 , (c)-(d) 0 = 50 , and (e)-(f) 0 = 55 . The other values
were (r) = 2 MPa and f = 1000.

189

B.1 Comparison of the RVE to the Constitutive Model - Plane Stress

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 60
1

0.5

0
60

11 , 12 (MPa)

1.5

20

20

40

2
1
0
1

(a)
40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

(b)
2
60

60

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 65
1

0.5

0
60

20

20

40

3
2
1
0

2
60

60

40

20

20

40

60

Strain (%)

0.5

11 , 12 (MPa)

Strain Energy,

60

(d)

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 70

40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

20

(c)
40

Strain (%)

11 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

0
60

(f)

(e)
40

20

20

Strain (%)

40

60

2
60

40

20

20

40

60

Strain (%)

Fig. B.5 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 60 , (c)-(d) 0 = 65 , and (e)-(f) 0 = 70 . The other values
were (r) = 2 MPa and f = 1000.

190

Supplementary Material - Chapter 3

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 80
1

0.5

0
60

11 , 12 (MPa)

1.5

20

20

40

2
1
0
1

(a)
40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

(b)
2
60

60

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 85
1

0.5

0
60

20

20

40

3
2
1
0

2
60

60

40

20

20

40

60

Strain (%)

0.5

11 , 12 (MPa)

Strain Energy,

60

(d)

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 90

40

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

20

(c)
40

Strain (%)

11 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

0
60

(f)

(e)
40

20

20

Strain (%)

40

60

2
60

40

20

20

40

60

Strain (%)

Fig. B.6 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 80 , (c)-(d) 0 = 85 , and (e)-(f) 0 = 90 . The other values
were (r) = 2 MPa and f = 1000.

191

B.2 Comparison of the RVE to the Constitutive Model - Plane Strain

B.2

Comparison of the RVE to the Constitutive Model Plane Strain

The capabilities of the model are assessed by comparison to the numerical model in Fig. 3.27
(Section 3.5.4). The vs. strain plane, and the stress-strain curves are compared for the
angles 0 = {25 , 45 , 90 }. Here, we supplement the data by including the remaining
angular configurations.
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 0
1

0.5

0
60

(a)
40

20

20

40

11 33 , 12 (MPa)

1.5

3
2
1
0
1

(b)
2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 5
1

0.5

0
60

(c)
40

20

20

Strain (%)

20

40

60

Strain (%)

40

60

11 33 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

(d)
2
60

40

20

20

40

60

Strain (%)

Fig. B.7 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 0 , and (c)-(d) 0 = 5 . The other values were (r) = 2 MPa
and f = 1000.

192

Supplementary Material - Chapter 3

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 10
1

0.5

0
60

(a)
40

20

20

40

11 33 , 12 (MPa)

1.5

3
2
1
0
1

(b)
2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 15
1

0.5

0
60

(c)
40

20

20

40

4
3

1
0
1

(d)

(e)
20

20

Strain (%)

40

60

11 33 , 12 (MPa)

Strain Energy,

0.5

40

60

40

20

20

40

60

Strain (%)

0
60

40

2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 20

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

Strain (%)
11 33 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

(f)
2
60

40

20

20

40

60

Strain (%)

Fig. B.8 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 10 , (c)-(d) 0 = 15 , and (e)-(f) 0 = 20 . The other values
were (r) = 2 MPa and f = 1000.

193

1.5

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 30
1

0.5

0
60

(a)
40

20

20

40

11 33 , 12 (MPa)

B.2 Comparison of the RVE to the Constitutive Model - Plane Strain

4
3
2
1
0
1

(b)
2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 35
1

0.5

0
60

(c)
40

20

20

40

4
3

(e)
20

20

Strain (%)

40

60

11 33 , 12 (MPa)

Strain Energy,

0.5

40

60

1
0
1

(d)
40

20

20

40

60

Strain (%)

0
60

40

2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 40

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

Strain (%)
11 33 , 12 (MPa)

1.5

20

4
3

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

(f)
2
60

40

20

20

40

60

Strain (%)

Fig. B.9 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 30 , (c)-(d) 0 = 35 , and (e)-(f) 0 = 40 . The other values
were (r) = 2 MPa and f = 1000.

194

1.5

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 50
1

0.5

(a)

0
60

40

20

20

40

11 33 , 12 (MPa)

Supplementary Material - Chapter 3

4
3
2
1
0
1

(b)
2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 55
1

0.5

0
60

(c)
40

20

20

40

4
3

(e)
20

20

Strain (%)

40

60

11 33 , 12 (MPa)

Strain Energy,

0.5

40

60

1
0
1

(d)
40

20

20

40

60

Strain (%)

0
60

40

2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 60

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

Strain (%)
11 33 , 12 (MPa)

1.5

20

4
3

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

(f)
2
60

40

20

20

40

60

Strain (%)

Fig. B.10 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 50 , (c)-(d) 0 = 55 , and (e)-(f) 0 = 60 . The other values
were (r) = 2 MPa and f = 1000.

195

1.5

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 65
1

0.5

0
60

(a)
40

20

20

40

11 33 , 12 (MPa)

B.2 Comparison of the RVE to the Constitutive Model - Plane Strain

4
3
2
1
0
1

(b)
2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 70
1

0.5

0
60

(c)
40

20

20

40

4
3

(e)
20

20

Strain (%)

40

60

11 33 , 12 (MPa)

Strain Energy,

0.5

40

60

1
0
1

(d)
40

20

20

40

60

Strain (%)

0
60

40

2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

0 = 75

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain (%)
1.5

Strain (%)
11 33 , 12 (MPa)

1.5

20

4
3

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

(f)
2
60

40

20

20

40

60

Strain (%)

Fig. B.11 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 65 , (c)-(d) 0 = 70 , and (e)-(f) 0 = 75 . The other values
were (r) = 2 MPa and f = 1000.

196

Supplementary Material - Chapter 3

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 80
1

0.5

0
60

(a)
40

20

20

40

11 33 , 12 (MPa)

1.5

3
2
1
0
1

(b)
2
60

60

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

40

Strain (%)
ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

Strain Energy,

0 = 85
1

0.5

0
60

(c)
40

20

20

Strain (%)

20

40

60

Strain (%)

40

60

11 33 , 12 (MPa)

1.5

20

ABAQUS Uniaxial
ABAQUS Shear
Model Uniaxial
Model Shear

2
1
0
1

(d)
2
60

40

20

20

40

60

Strain (%)

Fig. B.12 Comparison between model (3.22) and FE results for different values of the orientation angle in terms of strain energy vs. strain and stress vs. strain curves. The different
plots represent (a)-(b) 0 = 80 , and (c)-(d) 0 = 85 . The other values were (r) = 2 MPa
and f = 1000.

197

B.3 Comparisons of the SPRVE and MPRVE - Plane Strain

B.3

Comparisons of the SPRVE and MPRVE - Plane Strain

2D-SPRVE Uniaxial
2D-SPRVE Shear
3D-MPRVE Uniaxial
3D-MPRVE Shear
Model Uniaxial
Model Shear

0 = 0

11 33 , 12 (MPa)

11 33 , 12 (MPa)

In Section 3.5.3 the results of simple compression/tension tests for the 3D-MPRVE and
the 2D-SPRVE are shown in Fig. 3.21 for reinforcement angles, 0 = {30 , 45 , 75 , 90 }.
Here, we supplement the data by including the remaining angular configurations.

2
60

(a)
40

20

20

40

60

2D-SPRVE Uniaxial
2D-SPRVE Shear
3D-MPRVE Uniaxial
3D-MPRVE Shear
Model Uniaxial
Model Shear

(b)

60

40

20

11 33 , 12 (MPa)

Strain (%)
4

0 = 15

20

40

60

Strain (%)
2D-SPRVE Uniaxial
2D-SPRVE Shear
3D-MPRVE Uniaxial
3D-MPRVE Shear
Model Uniaxial
Model Shear

0 = 60

2
60

(e)
40

20

20

40

60

Strain (%)

Fig. B.13 Comparisons between the 3D-MPRVE and 2D-SPRVE for the cauchy stress difference (11 33 ) against strain for simple tension and compression, and the cauchy shear
stress (12 ) against the shear strain. Displayed for reinforcement orientations, (a) 0 , (b)
15 , (c) 60 .

Appendix C
Supplementary Material - Chapter 4
C.1

The Magneto-Mechanical Experimental Data

0.4
0.3

E/Einitial

Engineering Stress (MPa)

In Section 4.3, the effects of the magnetic field on the magneto-mechanical properties of the
nickel-coated fibre reinforced PDMS are assessed. Initial testing indicated the effect of the
magnetic field to be lower than the variability between equivalent specimens. Therefore,
the same specimen is tested multiple times, so as to more accurately assess the effect of the
magnetic field. The procedure involves conditioning of the specimen, by cycling 3 times to
10% strain. A number of subsequent cycles are then undertaken, with the specimen tested
under different magnetic field strengths. A cyclic test of a sample is shown here, to indicate
the minimal effect on the mechanical properties after the specimen is conditioned.
1

0.5

0
1 2 3 4 5 6 7 8 9 10

Cycle (No.)
0.2
0.1
0
0

10

Strain (%)

Fig. C.1 The cyclic testing of a 0 specimen. The specimen is cycled 10 times, to 10% strain
at a loading rate of 10mm min1 .

You might also like