Sunteți pe pagina 1din 234

Investigating the Recycled Rubber

Granulate-Virgin Rubber Interface

by

Praveen Kumar

A thesis submitted to the University of London


for the degree of Doctor of Philosophy

Department of Materials
Queen Mary, University of London

June 2007

10th June 2007

I hereby declare that the work carried out and presented in this thesis for
the degree of Doctor of Philosophy is original and my own.

.
Praveen Kumar

Abstract

Abstract
Due to environmental concerns and increasingly for economic reasons due to punitive
legislation that is being currently introduced world-wide, there is extensive interest in
materials recycling. The work undertaken and presented in this thesis focuses on the
recycling of rubber materials. Extensive previous research has investigated ways of
reclaiming rubber materials and some research has focused on measuring the behaviour
of materials that incorporate some quantities of recycled material. This project will
concentrate upon the reuse of ground-up rubber granulate mixed within a virgin
material.
Reusing granulates derived from old tyre stock and other sources in high tech
engineering applications is still considered a high risk option as typically it reduces the
strength of the whole composite. This work aims to develop techniques to identify the
rubber granulates, develop techniques to predict the behaviour of the final composite
and to explore in detail the reduction of strength properties resulting from the
incorporation of granulates.
This work shows that addition of granulates increases the intrinsic flaw size of a
finished product. The flaw size plays a vital role in determining the fatigue life and the
strength of the finished product, with larger flaws producing a weaker material. A
technique that can compare the strength of the interface between the granulate and the
virgin material with the basic strength of the virgin stock and the granulates is
developed. This thesis also looks at volume changes that take place in the matrixgranulate composite as a result of cavitation phenomena at the interface when the rubber
is subjected to cyclic loading. The development of these new techniques allows an
informed evaluation of the effect of reincorporating granulates to be established, so that
tailored interfaces that maximise the performance of rubber granulate-virgin rubber
composites can be optimised in the future.

Acknowledgements

Acknowledgements
This thesis would be incomplete without this page as it would not have been possible
without the support of various people that have been a part of my life during the course
of my doctorate at Queen Mary.
Its been a pleasure to work as a member of the Rubber group which is rightly headed
by Dr James Busfield. He has been an excellent supervisor and most importantly has
always been there when he was needed the most. Whatever I do in life, I hope and pray
that I am in the company of people like James who is very understanding, has faith in
his team and has such a great sense of humour. Sincere gratitude is extended towards
Prof. Alan Thomas who has helped steer this work with his vast knowledge and
experience. The bit that amazed me the most about Alan was how quickly he could
make reference to work done in the 50s or 60s. I salute his memory and consider it a
privilege that I had the opportunity to meet, talk and have lunch sessions with such a
well renowned physicist. I would also like to thank Prof. Yoshihide Fukahori for his one
on one discussions, supervision and advice. His involvement in this project was timely
and he brought in the industrial relevance to this work. I have been blessed to have coauthored publications with these three well respected individuals. Sincere thanks is also
extended to Prof. Craig Davies for his valuable input to this work and for his beneficial
advice at the time of my transfer.
My stay at Queen Mary was special because of the friends that I made here. Carmen,
who has always been there for a chat, laugh has been a true friend. Aqif, for looking
after me. Alan, Bill, Caroline, Dave, John, Kirsty, Monisha, Sandra, Steve, Vicky and
Vince who were very helpful. I shared my office with the weird bunch in 235 namely
Hano, Arab, Rob, Philip, Mac, Ioannis and had a great time mixing work with lots and
lots of fun filled activities. Another individual that I must name is Saba without whom I
doubt whether I would have been able to write up this thesis or survive the QM years.
She is a true friend and I wish her the world. The title MS wizard is apt for her.
I will take this opportunity to thank my mum and dad for believing in me. My family
has been very supportive and my brother and bhabhi deserve a special mention for
looking after me and putting up with my tantrums. With the little one due any time now,
they will have someone new to spoil.
3

Contents

Contents

Title page

Abstract

Acknowledgements

List of Contents

List of Figures

List of Tables

22

List of Symbols

24

Chapter 1: General Introduction

30

Chapter 2: Literature Review

34

2.1

34

Rubber
2.1.1

Rubber types & crosslinking

36

2.1.2

Other additives

38

2.2

The problem Why recycle rubber?

39

2.3

Tyre Raw material for recycling

43

2.4

Reuse of whole tyres & retreading

44

2.5

Recycling treatments and technologies

46

2.5.1

Granulates

48

2.5.2

Reclamation of rubber

56

2.5.3

Pyrolysis Carbon black production

57

2.5.4

Devulcanisation

57

2.5.5

Surface modification/activation

58

2.6

Energy recovery

59

2.7

Incorporation of granulates into rubber vulcanisates

61

2.7.1

Fracture behaviour

63

2.7.2

Adhesion at the interface

66

Contents

2.8

2.9

Strength/Crack growth phenomenon

71

2.8.1

Tear strength

73

2.8.2

Interfacial strength

75

Intrinsic flaws (crack initiation)

77

2.9.1

Edge cuts

79

2.9.2

Intrinsic flaw size (IFS) estimation

81

2.9.3

Corner cuts

82

2.10 Volume inflation


2.10.1

Dilatation and hydrostatic stress

85

2.10.2

Cavitation Volume change

94

2.11 Finite element analysis (FEA)

2.12

85

99

2.11.1

FEA models

101

2.11.2

Types of elements

102

2.11.3

Analysis procedure

103

2.11.4

Strain energy density

104

Summary and aim of this study

107

Chapter 3: Experimental Techniques & Materials

109

3.1 Introduction

109

3.2

Materials used in this study

110

3.3

Vulcanisation

112

3.4

Tensile strength and IFS measurement

113

3.5

Tear strength

115

3.6

Peel strength

117

3.7

Volume change measurement

119

3.8

Finite Element Analysis (FEA)

121

3.8.1

Pre-processing

122

3.8.2

Material characterisation

129

3.8.3

Boundary conditions

130

3.8.4

Analysis / Solution

131

3.8.5

Post processing

132

3.9

Granulate characterisation

132

3.9.1

132

ThermoGravimetric Analysis (TGA)

Contents

3.9.2

Fourier-Transform Infrared Spectroscopy (FTIR)

136

3.9.3

Pycnometry

139

3.9.4

Particle size measurement

141

Chapter 4: Intrinsic Flaw Size

145

4.1

Introduction

145

4.2

Experimental procedure for IFS

149

4.3

Mechanical characterisation stress strain behaviour

149

4.4

4.5

4.3.1

Characterising the stress versus strain behaviour of SBR 0

149

4.3.2

Characterising the stress versus strain behaviour of SBR 70

151

4.3.3

Tensile strength versus cut lengths

155

Intrinsic flaw size (IFS)

158

4.4.1

Strain energy release rate for edge and corner cut specimens

158

4.4.2

Flaw size equivalent to corner razor cut

160

Conclusions

167

Chapter 5: Matrix-Granulate Interface

168

5.1 Introduction

168

5.2

Experimental procedure

169

5.2.1

Measurement of tearing energy

169

5.2.2

Measurement of interfacial strength

169

5.3

5.4

Tearing and peeling energy

170

5.3.1

Matrix and granulate interface

176

5.3.2

Granulate and granulate interface

178

Matrix-granulate model
5.4.1

Use of energy balance approach to compute strain energy


release rates

179

5.4.2

Crack propagation path

184

5.4.3

Relation between crack growth rate and strain energy


release rate

5.4.4

185

Interpreting experimental and FEA data to predict crack


growth time

5.5

179

Conclusions

188
194

Contents

Chapter 6: Volume Change

195

6.1 Introduction

195

6.2

Experimental procedure for density determination

195

6.3

Experimental volume change trends

197

6.4

Cavitation model

199

6.5

Debonding in the model

202

6.6

Prediction of crack growth time

211

6.7

Conclusions

214

Chapter 7: Conclusions & Future Work

215

7.1

Conclusions

215

7.2 Future work

217

References

219

Appendix 1

229

Appendix 2

230

Appendix 3

232

Appendix 4

233

List of Figures

List of Figures
Chapter Two
Figure 2-1

Tensile stress-strain curves for polymers in three physical states. X


34

denotes rupture [Gent et al. (1992)].


Figure 2-2

(a) Rubber is composed of immensely long molecules. (b) The


effect of vulcanisation is to form crosslinks between the molecules
so that the whole specimen becomes one single molecular
network. (c) The rubber molecules are attached by the sulphur
crosslinks.

Figure 2-3

35

Chemical structures of commonly used general purpose elastomers


a) NR b) SBR [Gent et al. (1992)].

Figure 2-4

37

Plot of the comparison of the magnitudes of tearing energy


calculated from equation (2-7) and (2-15).

Figure 2-5

42

Data from Shulman (2004) showing whole tyre usage in the EU in


2003.

Figure 2-6

44

Schematic from Shulman (2004) showing the various treatments


and technologies used to recycle postconsumer tyres.
46

Figure 2-7

Figure 2-8

SEM scan by Jacob et al. (2003) of rubber granulates (a)


aggregates (b) single particle.

47

Use of granulates in the EU in 2003 Shulman (2004).

48

List of Figures

Figure 2-9

Illustration of mesh screens taken from report by Evans and Evans


(2006).

50

Figure 2-10

Mesh to microns particle size comparison [De et al. (2005)].

51

Figure 2-11

Sieves that are mounted on the sieve shaker [De et al. (2005)].

51

Figure 2-12

Sieving process used to separate granulates Evans and Evans


(2006).

Figure 2-13

52

Schematic showing the roll design with corrugations on them De


et al. (2005).

Figure 2-14

53

SEM contrasting the typical surface structures of cryogenically


ground rubber granulates (left) and ambiently ground rubber
granulates (right) [De et al. (2005)].

Figure 2-15

54

SEM images of ground tyre rubber produced by cryogenic


grinding (left) and pulverised rubber granulates (right) [De et al.
(2005)].

Figure 2-16

55

Data from Shulman (2004) showing the energy recovery from


post-consumer tyres in the EU.

Figure 2-17

59

The effect on torque with loading of granulates [Han and Han


(2002)]. 1 lbs.in = 0.113 Nm. Tmax: maximum torque; T40:
scorch time.

Figure 2-18

62

SEM contrasting fracture surfaces of similar SBR matrix with


granulates (volume fraction 25%) containing either 75 phr CB
(left) or 12.5 phr CB (right) [Burford and Pittolo (1984)].

63

List of Figures

Figure 2-19

Schematic of a rubber particle embedded in rubber matrix [Gibala


et al. (1998)].

Figure 2-20

64

Optical micrograph showing (a) Interfacial failure with a


cryogenically ground particle (b) fracture within an ambiently
ground particle [Gibala et al. (1998)].

Figure 2-21

65

Schematic of fracture process in composites due to (a) cohesive


failure in the matrix (b) adhesive failure between matrix and
granulate [Hilyard et al. (1983)].

Figure 2-22

67

SEM scan showing fracture surfaces of a specimen of a rubber


compound with 9% (by weight) cryogenically ground granulates
.[Burford and Pittolo (1983)]. The demarcation that separates the
top and bottom fracture surfaces can be seen.

Figure 2-23

68

SEM scan showing fracture surfaces of a specimen of a rubber


compound with 9% (by weight) ambiently ground granulates
.[Burford and Pittolo (1983)]. The demarcation that separates the
top and bottom fracture surfaces can be seen.

Figure 2-24

69

Changes in Ultimate tensile strength and elongation at break with


addition of scrap rubber into parent rubber compound [Burgoyne

70

et al. (1976)].
Figure 2-25

Effect of granulate quantity on tensile properties of the composite


[Han and Han (2002)]. 1 Kg/cm2 = 0.098 MPa.

Figure 2-26

Figure 2-27

71

Effect of average particle size of granulates on tensile properties of


the composite [Han and Han (2002)].

72

Schematic of a trouser test specimen.

73

10

List of Figures

Figure 2-28

Schematic of a peel test specimen.

Figure 2-29

Schematic of a specimen with an edge cut [Rivlin and Thomas

76

(1953)].

78

Figure 2-30

Experimental findings of Greensmith (1963).

80

Figure 2-31

Plot showing the effect of precut length and crosslinking on work


to break per unit volume. The crosslinking was done via
irradiation [Hamed (1983)].

Figure 2-32

82

(a) A schematic showing the corner cut in a tensile test specimen


(b) A schematic showing the calculation of the corner cut length c
from the fracture surface [Busfield et al. (1999)].
83

Figure 2-33

Findings of Busfied (2000) for a corner cut in a tensile test piece.


Comparison with Figure 2-30 shows that the shape of the function
with extension is similar.

Figure 2-34

84

The stretching frame used by Gee et al. (1950). in their study (a)
Elevation (b) Plan.

Figure 2-35

86

Sketch of the stretching rig devised by Mullins and Tobin (1957).


87

Figure 2-36

Change in volume with extension of carbon black filled natural


rubber [Mullins and Tobin (1957)].

Figure 2-37

88

Results of experiments by Christensen and Hoeve (1970) for


polyisoprene where volume change is plotted against extension
ratio. The dotted line represents the calculated volume change
using equation 2-12.

89

11

List of Figures

Figure 2-38

Relative volume change versus extension ratio for experiments on


unfilled SBR carried out by Fedors and Landel (1970). The data is
represented by filled circles and the curve is the predicted volume
change using equation 2-11.

Figure 2-39

91

Results of the study carried out by Fedors and Landel (1970) to


study the dependence of weight of immersed rig + rubber rings set
up on time and extension ratio.

Figure 2-40

92

Schematic showing formation of hollow spaces formed while


deforming a specimen [Reichert et al. (1987)].

95

Figure 2-41 Cavitation and debonding at the surface of a spherical inclusion in


an elastic matrix under tension [Gent and Park (1984)].
Figure 2-42

97

Cavitation occurring in a transparent silicone elastomer with two


rigid fillers in close proximity subjected to tensile strain [Gent and
Park (1984)].

Figure 2-43

98

Schematic from Busfield (2000) showing the four types of


symmetry that are commonly adopted to simplify the FE
modelling of a component.

102

Chapter Three
Figure 3-1

(a) SBR 0 and (b) SBR 70 granulates of 30 mesh size used in this
110

study.
Figure 3-2

Digital photograph showing the distribution of 30 mesh SBR 70


granulates in synthetic isoprene rubber. Granulates were mixed
using the two roll mill.

Figure 3-3

112

Instron 5567 with video extensometer setup that was used to

12

List of Figures

carry out tensile tests.


Figure 3-4

114

Microscopic image of a fracture surface showing the original cut


length.

Figure 3-5

115

(a) Schematic showing a trouser test specimen with a cut along


the centre line to allow the two legs to be gripped (b)
Experimental setup showing a trouser test specimen.

Figure 3-6

116

Schematic diagrams of peel test between sheets having same


composition as the matrix and granulates.

Figure 3-7

117

(a) Schematic of a peel test specimen with force applied across


the interface (b) Experimental setup showing a peel test
specimen.

Figure 3-8

118

Schematic diagrams of peel test between sheets having same


composition as the granulates.

119

Figure 3-9

Stretching rig used for the volume change experiments.

120

Figure 3-10

A schematic of the Mettler-Toledo balance that was used to


measure the volume change of the rings that were suspended by
means of the platinum wire shown above.

Figure 3-11

121

Half symmetry model of a tensile test specimen with an edge cut


(a) undeformed model (b) deformed model showing the region
with fine mesh containing the cut. The uncracked part of the
bottom surface is constrained in the 2 direction.
123

Figure 3-12

Deformed half 3D model of a tensile test specimen with a corner


cut. Nodes at the top are pulled in the 2 direction. The uncut

13

List of Figures

bottom surface is constrained in the 2 direction.


Figure 3-13

124

2D representation of the matrix-granulate model showing a


circular granulate embedded in the matrix.

Figure 3-14

125

Schematic of the matrix-granulate model indicating the two


paths along which nodes were released to simulate crack growth.
Bottom surface was fixed in both the 1 and 2 directions whereas
the top surface was constrained only in the 1 direction and was
moved in the two direction.

Figure 3-15

126

Deformed models of M70-G70(35%) showing the (a) tear


propagation through the granulate (b) tear profile under strain
and its final path.

Figure 3-16

126

Deformed models of M70-G70(35%) showing how rubber might


peel around the interface.

Figure 3-17

127

The axisymmetric FE model used for studying the debonding (a)


Granulate G is embedded in the Matrix M and the debonding
will take place along the dotted line which represents the
interface between the M-G. (b) 2D representation of an
undeformed model with 30 phr granulate incorporated by weight.
The arrow indicates the path along which the flaw size is
increased during the analysis. (c) 2D representation of model
strained to 50%. The arrow shows the direction where the
interface is debonding.

Figure 3-18

Figure 3-19

128

TGA plot showing the weight loss with temperature for SBR 70.
The dotted line is the DTGA curve.

134

TGA plot of granulates labelled compound X.

135

14

List of Figures

Figure 3-20

Diagram showing the main three types of vibrational modes of a


methylene group.

137

Figure 3-21

PA-FTIR scan of SBR 0 granulates.

138

Figure 3-22

PA-FTIR scan of compound X.

139

Figure 3-23

Particle size analysis for SBR 70 30 mesh granulates.

142

Figure 3-24

Particle size analysis for compound X.

143

Chapter Four
Figure 4-1

SEM scans to compare fracture surfaces of SBR 0 with no


granulates (left) and with 100 phr 30 mesh SBR 0 granulates
incorporated (right).

Figure 4-2

145

Tensile test piece with an edge cut a) Schematic b) is a


undeformed half FEA model of the same and c) is half FEA
model deformed to extension ratio of 2.

Figure 4-3

148

(a) Schematic representing a tensile test specimen with a corner


cut (b) 3D half symmetry FEA model deformed to extension
ratio of 2 (c) undeformed half FEA model.

149

Figure 4-4

Plot to determine the coefficients for Mooney SEF of SBR 0.

151

Figure 4-5

Stress versus extension ratio curves for SBR 0 measured in


tension and the Mooney function used to fit the experimental

Figure 4-6

data.

152

Plot to determine the coefficients for Yeoh SEF of SBR 70.

154

15

List of Figures

Figure 4-7

Stress versus extension ratio curves for SBR 70 measured in


tension and the Yeoh function used to fit the experimental data.

Figure 4-8

154

Effect of precut length on stress at break M70-G70(X) (SBR 70


matrix and 30 mesh SBR 70 granulates).

Figure 4-9

156

Effect of precut length on stress at break M0-G0(X) (SBR 0


matrix and 30 mesh SBR 0 granulates).

Figure 4-10

157

Effect of precut length on stress at break of M0-G70(X) (SBR


0 matrix and 30 mesh SBR 70 granulates).

Figure 4-11

158

Effect of precut length on stress at break of M70-G0(X) (SBR


70 matrix and 30 mesh SBR 0 granulates).

Figure 4-12

158

Plot to compare the amount of energy available to propagate a


crack for a M70 tensile test specimen with an edge cut to that
with a corner cut at 100% strain.

Figure 4-13

160

Specimen with a cut along A-A cross-section (a) Fracture


surface with edge cut (b) Fracture surface with corner cut. c
indicates the length of the cut.

Figure 4-14

161

Effect of precut length on stress at break for M70 (SBR 70


without granulates). The point of intersection is the corner cut
equivalent IFS for this material.

163

Figure 4-15

Comparison of IFS values for various materials tested.

166

Figure 4-16

b versus cut lengths curve for M70-G70(X) specimens. b of


M70 without a cut is extrapolated to the curve to obtain the
corner cut equivalent IFS.

166

16

List of Figures

Figure 4-17

b versus cut lengths curve for M0-G0(X) specimens. b of M0G0(47%) without a cut is extrapolated to the curve to obtain

167

the corner cut equivalent IFS.

Chapter Five
Figure 5-1

Plot comparing the tearing force for single cured SBR 70 and
SBR 0 specimens at a strain rate of 0.77 min-1

Figure 5-2

171

Comparison between force required to propagate a tear through


a single cured SBR 70 (representing matrix) and double cured

171

SBR 70 (representing granulate).


Figure 5-3

Plot used to obtain the value of strain energy density, W at any


strain for SBR 70 in order to compute the tearing energy.

Figure 5-4

172

Comparison between force required to propagate a tear through


a single cured SBR 0 (representing matrix) and double cured
SBR 0 (representing granulate).

Figure 5-5

173

Plot used to show the relation between W and extension ratio


for SBR 0 double cured (representing granulate) in order to
compute the tearing energy.

Figure 5-6

174

Normalised plot that compares the peeling force between filled


matrix-filled granulate and unfilled matrix-unfilled granulate
combinations.

Figure 5-7

174

Normalised plot that compares the peeling force between filled


granulate-filled granulate and unfilled granulate-unfilled
granulate combinations.

Figure 5-8

175

FEA models of M0-G0(47%) held at 300% strain showing the

17

List of Figures

crack propagating through the granulate (tearing) [Tavakkoli


(2006)].
Figure 5-9

179

FEA models of M70-G70(35%) held at 300% strain showing


the crack propagating at the matrix-granulate interface
(peeling) [Tavakkoli (2006)].

Figure 5-10

180

Change in total elastic strain energy with crack length for M0G0(47%) at 300% strain [Tavakkoli (2006)].
181

Figure 5-11

Change in total elastic strain energy with crack length for M70G70(35%) at 300% strain [Tavakkoli (2006)].

Figure 5-12

182

A plot comparing the energy required for crack propagation for


two different routes in the matrix-granulate model strained at
300%. In this case the matrix is SBR 0 and it contains 47%
SBR 0 granulate by volume [Tavakkoli (2006)].

Figure 5-13

183

A plot comparing the energy required for crack propagation for


two different routes in the matrix-granulate model strained at
300%. In this case the matrix is SBR 70 and it contains 35%
SBR 70 granulate by volume [Tavakkoli (2006)].

Figure 5-14

184

Cut growth rate, dc/dt as a function of tearing energy, T for


unfilled SBR. () tensile test pieces, (+) pure shear test pieces
and the line is according to eq. ( 5-4) [Lake and Lindley

185

(1964)].
Figure 5-15

Double log plot of crack growth rate versus strain energy


release rate (i.e. tearing or peeling) for M0-G0 and M70-G70
type materials.

Figure 5-16

187

Plot that shows the crack growth rate variation with crack
length as the crack propagates through a SBR 0 double cured
strip (G0).

189

18

List of Figures

Figure 5-17

Time required for crack to propagate from one point to another


across the G0 granulate can be estimated using this plot. The
integral shown being the time taken by the crack to grow from
c1 = 6.6 x 10-6 m to c2 = 6.2 x 10-4 m as 7.02 seconds.

Figure 5-18

190

Time required for crack to propagate from one point to another


across the M0-G0 interface can be estimated using this plot.
The plot inset shows the area under the curve from c1 to c2. The
estimated time for crack to grow from c1=6.4 x 10-6 m to c2 =
4.5 x 10-4 m is 1.88 seconds.

Figure 5-19

191

Time required for crack to propagate from one point to another


across the G70 granulate can be estimated using this plot. The
plot inset shows the area under the curve from c1 to c2. The
estimated time for crack to grow from c1=6.6 x 10-6 m to c2 =
6.2 x 10-4 m is 55.08 seconds.

Figure 5-20

192

Time required for crack to propagate from one point to another


across the M70-G70 interface can be estimated using this plot.
The plot inset shows the area under the curve from c1 to c2.
Time taken for crack to propagate from c1=6.4 x 10-6 m to
c2 = 4.5 x 10-4 m is estimated to be 15.3 seconds.

193

Chapter Six
Figure 6-1

Experimental values of the fractional increase in volume observed


for all the materials with SBR 70 as matrix with various quantities
of granulates incorporated plotted against extension ratio. These
are compared with the dilatational values expected from
hydrostatic stress for M70 calculated using equation 2-18.
197

Figure 6-2

Experimental values of the fractional increase in volume observed


for SBR 0 as matrix with various quantities of granulates

19

List of Figures

incorporated plotted against extension ratio. These are compared


with the dilatational values expected from hydrostatic stress for
M0 calculated using equation 2-18.

Figure 6-3

198

The crack profile is shown as the model is strained with nodes


released along the interface. The void is represented by the shape
outlined by ABCO.

Figure 6-4

199

A schematic showing the upper half of the crack profile in the


axisymmetric model and how the crack volume was calculated.
Volume of the strip under each element edge along the crack front
was considered for this calculation.

Figure 6-5

201

Experimental volume change experienced by M70-G0(41%) due


to cavitation superimposed on the fractional volume change due to
debonding in the FE model for the same material combination.
The values on the right in the figure above indicate the modelled
debond length values (Db).

Figure 6-6

202

Experimental volume change experienced by M70-G70(35%) due


to cavitation superimposed on the fractional volume change due to
debonding in the FE model for the same material combination.
203

Figure 6-7

Experimental volume change experienced by M0-G0(47%) due to


cavitation superimposed on the fractional volume change due to
debonding in the FE model for the same material combination.
204

Figure 6-8

Plot obtained from the intersections in Figures 6-5, 6-6 and 6-7. It
shows how the intrinsic flaw size (IFS) also increase with strain
for materials with a large granulate volume fraction.

Figure 6-9

205

Experimental volume change experienced by M70-G70(14%) due

20

List of Figures

to cavitation superimposed on the fractional volume change due to


debonding in the FE model.
Figure 6-10

205

Experimental volume change experienced by M0-G0(21%) due to


cavitation superimposed on the fractional volume change due to
debonding in the FE model.

Figure 6-11

206

Experimental volume change experienced by M70-G0(17%) due


to cavitation superimposed on the fractional volume change due to
debonding in the FE model.

Figure 6-12

206

Plot obtained from Figures 6-9, 6-10 and 6-11. It shows how the
intrinsic flaw size (IFS) also increase with strain for materials with
a small granulate volume fraction.

Figure 6-13

207

Plot showing that addition of granulates into matrix rubber


increases the average debond length at 300% strain.

Figure 6-14

208

Plot showing how the peel energy varies with growing crack along
the M0-G0 interface.

Figure 6-15

211

Plot showing how the peel energy varies with growing crack along
the M70-G70 interface.

Figure 6-16

212

Plot used to compute the time required for a crack to progress


from one point to another along the interface of the M0-G0 model
held at 300% strain.

Figure 6-17

213

Plot used to compute the time required for a crack to progress


from one to point to another along the interface of the M70-G70
213

model held at 300% strain.

21

List of Tables

List of Tables
Chapter Two
Table 2-1

Disposal route for waste tyres in the UK in 2004 [Chatterley


42

(2004)].

Chapter Three
Table 3-1

Formulation for the materials used in this study.

Table 3-2

Combination of matrix and granulates used for the IFS


estimation.

Table 3-3

Table 3-4

111

113

Comparison of TGA analysis of SBR 70 with the actual


formulation data.

135

Results of density measurements for rubber granulates.

140

Chapter Four
Table 4-1

Stress at break values for the material types used in this study.

155

Table 4-2

IFS values for the different materials tested.

164

Chapter Five
Table 5-1

Energy table to study interface between matrix and granulates


of same composition [M70-G70(X) and M0-G0(X)].

Table 5-2

Table 5-3

176

Energy table to study interface between matrix and granulates


of different composition [M0-G70(X)) and M70-G0(X)].

177

Energy table to study interface between granulates.

178

22

List of Tables

Table 5-4

Table 5-5

Coefficients obtained for crack growth materials in order to


establish the power law relationships.

188

Comparison of time required for crack/flaw to grow.

193

Chapter Six
Table 6-1

Comparison between IFS values [Kumar et al. (2007)] and


average

debond

lengths

(from

Figure

6-13)

for

the

homogeneous matrix-granulate composites.

Table 6-2

209

Comparison between IFS values [Kumar et al. (2007)] and


average

debond

lengths

(from

Figure

inhomogeneous matrix-granulate composites.

6-13)

for

the
210

23

List of Symbols

List of Symbols
English Alphabet
A

Surface area of specimen

Ac

Area of a single surface of the crack

Ao

Unstrained cross-sectional area of the specimen

Bulk modulus

C10, C01

Coefficients for the Mooney SEF to determine material properties

C10, C20,

Coefficients for the Yeoh SEF to determine material properties

C30
c

Crack length cut length precut length

co

Initial crack length

Db

Debond length intrinsic flaw size

Dev70-0

Difference between the calculated volume change due to hydrostatic


effects alone from the experimentally measured volume change values
for M70-G0(41%) at 300% strain

diameter

dc/dt

Crack growth rate

Tensile modulus - Youngs modulus

Free energy of strained rubber

Fliq

Liquid like free energy

Fel

Sum of the free energies of the network chains

Force

FBuoy

Buoyant force

Elongation or retraction force

G70

Granulate is SBR 70

G0

Granulate is SBR 0

Strain energy release rate

Gc

Fracture energy

I1, I2, I3

Strain invariants

24

List of Symbols

Local ratio of deformed to ground-state volumes

Bulk compressibility

K()

Dimensionless function of strain

K()

Dimensionless function of strain similar to K()

lo

Initial unstrained length of specimen

Strained length of specimen

M70

Matrix is SBR 70

M0

Matrix is SBR 0

Change in mass between stretched and unstretched positions

Mass of sample

Peeling energy work of separation per unit area of interface strength


of adhesion strain energy release rate

Pc

Critical pressure

Pr

Pressure

Rate of grip separation in the trouser tear test

Sq

Spring constant

Tearing energy strain energy release rate

T40

Scorch time

Tc

Critical tearing energy

Te

Temperature

Tg

Glass transition temperature

Tmax

Maximum torque

Thickness of specimen

ta

Critical stress for debonding

tc

Critical stress for cavitation

ti

True stress

t1, t2, t3

True principal stresses

Total elastic strain energy

U0

Total elastic strain energy in the test piece without a crack

Ub

Strain energy density at break

Uc

Total elastic strain energy in the test piece with a crack of length c

25

List of Symbols

Volume change

Volume

V0

Initial volume of rubber

Vsolid

Volume of solid

Elastic strain energy density

Wb

Work to break per unit volume

Total width of the specimen

Volume fraction

26

List of Symbols

Greek Alphabet

Material constant

Extension ratio

Extension ratio at break elongation at break

1, 2, 3

Principle extension ratios

Angle subtended by a debonded circular patch on the inclusion that is


located at the pole

Density

fluid

Density of fluid

Density of water

Engineering stress

Reduced stress

1, 2, 3

Principle Cauchy stresses

Tensile strength stress at break

Displacement

Material constant

x&

Cross head separation rate

27

List of Symbols

Abbreviations
CAX4H

Axisymmetric four node elements

CB

Carbon black

CPS4

Plane stress 4 node elements

C3D8H

Solid (continuum) three-dimensional eight-node hybrid brick elements

DPG

Diphenyl guanidine

DOF

Degrees of freedoms

DTGA

Differential thermal gravimetric analysis

ELV

End of life vehicles

EPDM

Ethylene propylene diene monomer

EU

European union

FE

Finite element

FEA

Finite element analysis

FEM

Finite element method

FTIR

Fourier transform infrared spectroscopy

Granulate

HAF N330

High abrasion furnace carbon black

HPPD

N-phenyl-N-(1,3-dimethylbutyl)-p-phenylenediamine

IFS

Intrinsic flaw size

IRB#5

Industry reference carbon black

LALLS

Low angle laser light scattering

Matrix

MBTS

2,2'-Dithiobis(benzothiazole)

MPC

Multi point constraints

NR

Natural rubber cis 1-4 polyisoprene

PA

Photoacoustic

PTFE

Polytetrafluoroethylene

phr

Parts per hundred of rubber by mass

SBR

Styrene butadiene rubber

SBR 70

Styrene butadiene rubber with 70 phr carbon black

SBR 0

Styrene butadiene rubber with 0 phr carbon black

28

List of Symbols

SEF

Strain energy function

SEM

Scanning electron microscopy

TDF

Tyre derived fuel

TGA

Thermo gravimetric analysis

UTS

Ultimate tensile strength

UV

Ultra-violet

VCE

Virtual crack extension

WRAP

Waste and resources action programme

29

Chapter 1

General Introduction

General Introduction

Elastomers or rubbers are polymeric materials that are above the glass transition
temperature, that have low levels of crystallinity and which are usually lightly
crosslinked to form a macromolecular network. From the discovery of the new world
until the second world war, the only source of rubber was natural rubber (NR) obtained
from sap of Hevea Brasiliensis tree. NR found widespread commercial application in
the midnineteenth century in waterproof clothing, balls, erasers and belts for
machinery. Latex, from which NR is produced, is a white milk-like fluid which is
obtained by making cuts in the outer bark of the tree. Synthetic rubbers were developed
during the second world war to ensure continuity of rubber supply.
The commercial potential of all rubber materials was greatly extended in the 1840s
when vulcanisation by sulphur was discovered by Goodyear and patented by Hancock.
This invention was of great significance as it eliminated the unpleasant tackiness of
previous rubber products. Rubber transformed the bicycle from a dangerous mode of
transport into a much safer personal mechanical vehicle. The reason rubber has become
so popular is that once vulcanised the rubber can be stretched easily and almost
completely reversibly to high extensions. It also is able to dissipate energy because of
viscoelastic behaviour and this makes it unusual when compared to other solid materials
such as glass and metals. An example to illustrate this is an ordinary rubber band
usually made of NR that can be stretched up to several times its original length and
upon release it will recover almost completely to its initial unstrained state.
The application of rubber in tyres was commercialised by Michelin in the late 19th
century with the production of pneumatic tyres for the motor car. This resulted in
phenomenal growth of consumption of rubber by the motor industry with the
automotive sector now being the largest consumer of rubber materials in the world. This
is evident from the fact that an automobile vehicle has over 500 different rubber
components that include tyres, belts and suspension systems. The tyre industry alone
encompasses a huge product range from tyres for very wide ranges of vehicles, that can
operate under very diverse conditions to tyre repairing kits and other materials such as
tyre tread buffings and peelings.

30

Chapter 1

General Introduction

The non tyre industry spans from consumer goods like mats, shoe soles, bearings, anti
vibration mountings, conveyor belts, wiper blades, oil seals, stationary items, to
healthcare products such as latex gloves and condoms. All these applications of rubber
were realised after the invention of vulcanisation which has made rubber materials very
versatile. However, the very process that makes the rubber materials so useful also
makes rubber a difficult commodity to recycle and causes difficulties when rubber
products reach the end of their life. With tyres the problem is more severe because tyres
are usually composites that consist of blends of different vulcanised elastomers together
with steel and other fabrics. These products are designed to be very stable and have a
long life.
From a recycling point of view, the items that fall under the category of non tyre
industry are less significant as compared to the ones in the tyre industry. The reason for
this is that in the current scenario, approximately 70% of all the rubber (natural and
synthetic) consumed is used in the production of tyres. Vast quantities of tyres, upon
reaching the end of their lives, are dumped into landfills, stockpiled or exported. The
current disposal methods cannot cope with the ever increasing tyre mountains and this is
not environmentally sustainable. The end of life vehicles directive introduced in the EU
banned tyres from landfills from July 2006 onwards and since this legislation came into
effect a suitable environmentally benign method has been required to dispose off all the
rubber scrap. This directive provides a stronger push for car manufacturers and tyre
companies to upgrade their existing recycling technologies as this helps them to
improve costs and customer perception.
One of the methods of recycling vulcanised rubber is to perform a size reduction or
grinding down of the tyres into finer particles. This size reduction enables their
utilisation in the vulcanised state. Currently these granulates are used as cheap fillers
and processing aids in compounds. Concentration of granulates used for any purpose is
low and these products have low level engineering demands. This is because researchers
have shown that the addition of granulates into virgin rubber compounds results in
reduction in the strength, fatigue and elongation to break properties of the latter. Hence
currently the use of granulates in new tyre compounds does not exceed more than 10%
of the rubber in the tyre. The need for finding better applications for these granulates is

31

Chapter 1

General Introduction

the key to the growth of this recycling so that a higher volume fraction of granulates can
be incorporated.
The aim of this thesis is to develop techniques that can be used to identify the cause for
this reduction of strength properties of the composite material that is formed when
rubber granulates are mixed into virgin rubber compounds. The aim being to establish a
reliable technique that can be used to predict the failure of the composite.
Chapter 2 gives an overview of the legislations that have been introduced in order to
promote rubber recycling. There is then a review of the techniques used by various
researchers in the past who have studied rubber-granulate composite materials and also
discusses their findings. The concept of finite element analysis (FEA) which is used at
various stages in this work, is also explained in this chapter.
Chapter 3 discusses the materials and compounding recipes used to prepare the
compounds investigated in this study. Various experimental and finite element analysis
techniques were used to analyse these specimens and these are also discussed here. In
the initial stages of this work, characterisation techniques were used to determine the
composition of the rubber granulates. The scope of these techniques and the findings are
also discussed here.
The focus of chapter 4 is to determine the cause for the poor strength properties of
virgin rubber-granulate composite materials. Flaws are present in all polymers which
can be either surface flaws or flaws that are inherent to the material. Typically these are
not visible to the naked eye. As rubber is used in so many sectors, flaws are a cause of
concern and identification of the size and the reduction in size of these flaws is a crucial
concern, particularly for the situation when granulates are incorporated into virgin
material. During the course of this study a technique was developed to estimate this
intrinsic flaw size (IFS) of rubber materials and the flaw size values of virgin matrix
with and without granulates could be compared. The effect of incorporation of rubber
granulates on the Intrinsic Flaw Size (IFS) of the virgin rubber could also be examined.
Another aspect of this work was to examine the interface between granulates and virgin
matrix. The interfacial energy would clearly depend on the extent of any bonds that
have been formed across the interface between granulates and matrix. A technique was
developed to quantify this interfacial strength. FEA was used to examine the energy

32

Chapter 1

General Introduction

available for crack propagation in matrix-granulate composites. Combining the


experimental and FEA results it was possible to predict the crack growth rate. This
analysis is discussed in chapter 5.
Chapter 6 discusses the volume changes that take place when matrix-granulate
composites are strained. The technique adopted was based on measuring density
changes using Archimedes principle. In addition, FEA was used to study the separation
along the granulate-virgin rubber interface to estimate if unpeeling would take place.
The conclusions of this work are summarised in chapter 7 where the scope for further
work is also discussed.

33

Chapter 2

Literature Review

2
2.1

Literature Review

Rubber

Polymers are made up of long, flexible molecules that are linked where the atoms along
the polymer backbone are connected to each other by covalent bonds. A polymer
comprises of a series of monomer units connected by a chemical reaction known as
polymerisation. This gives rise to the long chain nature of a polymer that sets it apart
from other materials. As reported by Gent et al. (1992) and shown in Figure 2-1,
different types of polymer can exhibit the following three physical states:

Glassy The polymers are hard and brittle.

Crystalline The polymers typically go through a succession of changes before


they break (elastic yielding plastic flow necking strain hardening
and fracture).

Amorphous The polymers are rubbery.

Figure 2-1 - Tensile stress-strain curves for polymers in three physical states. X denotes
rupture [Gent et al. (1992)].

34

Chapter 2

Literature Review

Rubber or elastomers are amorphous polymers where the chains are interlocked with
each other along their length at various intervals by means of physical or chemical
crosslinks. The chemical crosslinks are formed by a process known as vulcanisation,
which was the first great discovery of rubber technology. Traditionally this process
requires the addition of chemicals which are added to the raw rubber followed by the
addition of pressure and heat [White and De (2001)]. A coherent 3D network is the
result in which all the molecules are linked to each other by bridges of sulphur atoms
and is shown in Figure 2-2. Vulcanisation transforms rubber from a viscous liquid into a
soft, highly elastic solid (vulcanisate) and it is this process that made rubber such a
widely used material. The extent of cross linking must be high enough to prevent failure
by viscous flow but low enough to avoid brittle failure.

Figure 2-2 - (a) Rubber is composed of immensely long molecules. (b) The effect of
vulcanisation is to form crosslinks between the molecules so that the whole specimen
becomes one single molecular network. (c) The rubber molecules are attached by the
sulphur crosslinks.

35

Chapter 2

Literature Review

2.1.1 Rubber types & crosslinking


Gent et al. (1992) have reported that sulphur is the most widely used crosslinking agent
and accelerators are added in order to increase the rate and efficiency of crosslinking.
There are five major categories of accelerators which are guanidines, thiazoles,
dithiocarbamates, xanthates and thiurams. Diphenyl guanidine (DPG) belongs to the
guanidine category and is a secondary accelerator that when added at the time of
vulcanisations brings about a slow onset of vulcanisation (scorch). As DPG gives the
lowest rate of vulcanisation, it is usually used in combination with other types of
accelerators. 2,2'-Dithiobis(benzothiazole) also known as MBTS is a primary
accelerator and belongs to the category of thiazoles which are widely used in
combination with a primarily accelerator. It is reported to give a moderately fast curing
rate for synthetic rubbers such as SBR. Combining accelerators helps achieve scorch
resistance and optimal cure rate. The onset of vulcanisation becomes rapid when
accelerators of the thiurams, dithiocarbamates and xanthates category are added. Thus
while using these accelerators care must be taken to begin processing soon after mixing.
Activators such as zinc oxide and stearic acid are also added to make the accelerated
sulphur curing more efficient. The most common instrument that is used to monitor
crosslinking is the oscillating disk rheometer [Blow (1971)]. It consists of an oscillating
rotor that is housed within a heated chamber. The test specimen is heated in the
chamber. The torque required to oscillate the rotor is monitored as a function of time.
Typically when a new specimen is introduced, a decrease in the torque is observed
because the rubber experiences a drop in viscosity as it is heated. With further heating,
the rubber begins to vulcanise and as it stiffens, the torque rises. This indicates that
crosslinking is taking place.
The properties of the vulcanisates will depend on whether the crosslinks are
monosulphidic (C-S-C) or polysulphidic (C-Sn-C). The proportion of C-S linkages
which are more stable than S-S linkages are higher, if the accelerator/sulphur ratio and
cure time is high. The advantage in such a case is that the vulcanised elastomer will
have better geometric stability and heat resistance throughout the products life. As S-S
bonds in polysulphidic linkages are weaker, these have the ability to break and reform
thus relieving local stresses that could otherwise initiate failure. Thus there is a strange

36

Chapter 2

Literature Review

paradox that compounds that contain the weaker polysulphidic crosslinks have higher
strength properties.
Crosslinked Natural rubber (cis-1,4- polyisoprene) and styrene butadiene rubber (SBR)
are the most commonly used general purpose elastomers and their primary application
is in automotive industry.
Natural rubber (NR) is a homopolymer and is obtained from the latex of the rubber tree,
Hevea Brasiliensis. On the other hand, styrene butadiene rubber (SBR) is a copolymer
of styrene with butadiene and is a large volume synthetic rubber. The chemical
structures of NR (cis 1,4-polyisoprene) and SBR are shown below in Figure 2-3.

Figure 2-3 Chemical structures of commonly used general purpose elastomers a) NR


b) SBR [Gent et al. (1992)].

The major advantage of SBR over a homopolymer such as NR is that by varying the
ratio of the polymers that make up the copolymer, different properties can be realised.
Rubber, whether synthetic or natural is different from other solid materials such as glass
or metal because of the following properties:

The material is highly extensible and soft.

Complete recovery is possible after gross mechanical deformations.

37

Chapter 2

Literature Review

However in order to exhibit these properties the following two conditions must be
satisfied:
1) Rubber must be crosslinked as vulcanised rubber is much more elastic than the
original raw rubber.
2) The temperature must be above the glass transition temperature, Tg (which is
-70C and -55C for the case of NR and SBR respectively).
According to McCrum et al. (1988), Tg, marks the transition region for a material where
its properties change from glassy to rubbery with increasing temperature. This means
that in the case of NR, above -70C, the material behaves as rubber and below it the
material behaves more like glass. This happens because the flexible chains of rubber are
able to move past each other easily due to thermal vibrations above Tg but the molecules
are frozen at temperatures below Tg. The material then behaves like glass that is
shattered easily. The Tg values vary for different elastomers. The specific value is
readily obtained from volume-temperature measurements for a material during cooling.
The Tg shows up as a marked discontinuity to the plot of volume versus temperature.
The properties of rubber also vary depending upon the quantity of fillers and other
compounding ingredients that are mixed into the raw rubber before vulcanisation.
According to the kinetic theory of rubber elasticity in its simplest form, these unique
properties of rubber can be explained by attributing these to the changes in entropy,
which results from a configurational modification in the molecular network during the
deformation of the rubber [Treloar (1975)].
2.1.2 Other additives
Fillers such as carbon black (CB) and silica are added to reinforce the amorphous
rubber. According to Gent et al. (1992) their effect is more pronounced in the case of
SBR which is a non crystallising elastomer. Strain crystallising elastomers such as NR
undergo crystallisation when strained and thus have a self reinforcing mechanism. To be
an effective reinforcing particulate filler, the following conditions must be met:

The specific surface area of the filler must be high which means that the fillers
must be smaller than 1 m as this results in greater interaction with the rubber.
The primary particle size of CB ranges from 8 to 75 nm.

38

Chapter 2

Literature Review

The filler must have an irregular shape so that it will have a void volume within
which the rubber molecules can get occluded which in turn results in greater
interaction and thus more reinforcement.

The extent of interaction between filler and rubber must be high. This can be
improved by modifying the surface chemistry. This is usually done with silica
fillers by treating them with silane coupling agents which helps improve the
bonding between the filler and rubber.

The quantity of filler is conventionally expressed in terms of mass of filler as a


proportion of the mass of the elastomer and is usually termed parts per hundred parts of
rubber (phr). As an example, a compound with 1000g of polymer, 700g of carbon black
filler and 200g of other active ingredients such as antioxidants and curatives is said to
have 70 phr of filler. It is clear that this is equivalent to a mass fraction of filler of
700/1900.
Oils are added towards the end of the mixing cycle to enhance the dispersion of filler
particles in the rubber. As fillers have a tendency to agglomerate oil is added to
maintain a high viscosity of rubber which means that there will be a constant shearing
action that helps break down these agglomerates.
Antioxidants are added at the time of mixing to restrict the alteration/degradation of the
network structure of rubber by oxygen and ozone.
2.2

The problem Why recycle rubber?

A tyre becomes a post consumer tyre after it has served for over thousands of kilometres
and is removed permanently from a vehicle. Post consumer tyres are not classified as
hazardous waste and this is the reason that in the past the bulk of these were sent to
a) landfills b) stored in warehouses or c) exported to other countries in Asia, Africa and
parts of Latin America for overseas retread operations. Some of the alternative proposed
methods to deal with these tyres were for retreading, material recovery or as a fuel
source. The issue with heaps of tyres lying around (landfills) is that they:
1) Act as breeding ground for insects and vermin as water is accumulated in them.
2) Contaminate the ground as a result of the leaching effects of additives that are
toxic, which causes problems for future land use.

39

Chapter 2

Literature Review

3) Do not create a pleasing landscape and they also pose a fire hazard.
4) Do not biodegrade when buried.
Hence landfilling is considered to be a disposal technique where the value added to the
waste material is negative.
A publication by Shulman (2004) puts forward estimates for rubber (natural and
synthetic) produced globally to be approximately 20,000,000 tonnes per year. The
European Union (EU) alone consumes about 4,000,000 tonnes annually and imports
additional 1,000,000 tonnes per year from outside the EU as finished goods including
tyres. As far as post consumer tyres are concerned, in 2003, these amounted to more
than 2,600,000 tonnes of waste in the then 15 member states of the EU. It was
approximated then that the expansion of EU to 25 member states would account for an
additional 250,000 tonnes per year. It is believed that about 75% of the rubber produced
in the world is consumed by the various sectors in the automotive sector. The bulk,
close to 60%, is used in the production of tyres (tyre sector) and the remaining 15% is
used for non-tyre automotive items such as belts, hoses, mountings, seals and
extrusions. The remaining 25% of the natural and synthetic rubber produced is used in
the production of general non-automotive rubber products. Typical applications include
footwear, seals, expansion joints, bladders and gloves.
The world tyre production is estimated to be approximately 1,000,000,000 units per
year since the 1990s. In terms of the number of units sold which is slightly less than
those produced, there are broadly three markets for tyres. The principal one is that of
passenger car tyres which accounts for more than 90% units sold. The second category
is that of truck tyres and the third category comprises of smaller diverse categories such
as aeroplane, bicycle and agriculture. Together these two categories account for the
remaining 10%. Tyres from any one of these categories are labelled as waste when they
are removed from the vehicle without the possibility of it being used again as a regular
vehicle tyre on the road. It is therefore clear that the quantity of tyres generated as waste
is large.
There is added pressure on European tyre industry due to the new EU directives coming
into effect as it is estimated that approximately 1 million tonnes of additional material

40

Chapter 2

Literature Review

will be generated in the EU. These directives are: [Chatterley (2004), Department of
Trade and Industry (2001)]:
1) The Landfill Directive (1999/31/EC) Landfills have been one of the early and
most popular ways of disposal of unwanted rubber products. This directive
specifically bans whole tyres from being dumped into landfills from July 2003
and tyre shreds from July 2006 onwards. As it lists tyres as a major waste
stream, it becomes imperative for this industry to have in place reuse and
recovery systems that are capable of contributing towards handling virtually all
of the waste tyres within EU member states.
2) End of Life Vehicles (ELV) Directive (2000/53/EC) This directive required
that by January 2006, at least 85% of the weight of an average car should be
recovered. It was proposed that the majority of this target could be achieved by
recycling and reusing the vehicles constituent materials. To counteract the
increase in car ownership, a new target of 95% recovery has been set for 2015
and in all possibility this will increase still further in the future. This directive is
aimed to reduce waste materials that go into landfill sites.
3) Waste Incineration Directive (2000/76/EC) European law places restrictions
on the emissions from incineration of waste. It is applicable to most activities or
organisations that are involved in the burning of waste, whether for disposal or
for usage as fuel. It was implemented in the UK in 2002.
Table 2-1 indicates the disposal routes in UK in 2004 and according to Chatterley
(2004) these proportions have not changed significantly since 1999.
Even in 2004, significant quantities of tyres (29%) were disposed via the landfill route.
As the landfill directive has come into effect since July 2006, there is no future
possibility of landfilling. It is therefore important to find alternate route(s) for handling
these tyres that have been historically disposed of in landfills.
The whole drive for rubber reuse/recycling has been driven more from a political
perspective rather than a scientific one. Companies recycle as a part of a marketing
strategy to show an environmentally sound image. The challenge is to look at
costeffective, energy-efficient and environmentally benign methods to reduce, reuse,

41

Chapter 2

Literature Review

recycle or recover rubber. Recycling as a sector can also provide employment to many
people who can contribute towards waste minimisation and resource conservation.

Table 2-1 Disposal route for waste tyres in the UK in 2004 [Chatterley (2004)].

Disposal Route

Percentage Disposed

Approximate Tonnage

Reuse and retreading

16

72,000

Recycling

27

121,500

Energy recovery

16

72,000

Landfill

29

130,500

Other

12

54,000

Total

100

450,000

Figure 2-4 Typical structure of a car tyre (source: http://www.rma.org).

42

Chapter 2

2.3

Literature Review

Tyre Raw material for recycling

A tyre is a round, black, hollow shell that is inflated with compressed air so that it can
provide a flexible cushion between the vehicle and the road surface. It is the only point
of contact between the two. Car and truck tyres have a complex structure that typically
consists of about thirty different components made with different rubber compounds.
This complex structure is illustrated in Figure 2-4 where some of the critical parts of the
tyre are shown. It is clear from the figure that the tread and sidewalls constitute exterior
elements whilst the belts, body ply, beads and inner liner are interior components
[Shulman (2004)].
The part of the tyre that comes in direct contact with the road surface in order to provide
cornering grip and traction in various conditions, is the tread. It is usually made of SBR
with a high concentration of carbon black in case of passenger tyres whereas in case of
truck tyres a highly carbon black filled NR/SBR blend is used depending on the
performance characteristics. So that the tread is able to maintain contact with the road
under all weather conditions it usually consists of various grooves and ridges to allow
water to clear away by flowing outwards.
The reinforcement of the tread and maintaining the shape of the tyre is done by the
belts. These are usually layers of rubber sheets containing brass coated carbon steel
wires.
The beads are structural components that are made from coils of zinc or bronze coated
single filament high strength steel wire coated with rubber. These ensure an air tight fit
between the tyre and the wheel rim so that there is no shift during driving.
Sidewalls protect the side of the tyre from road irregularities by reinforcing the
interface between the tyre and the wheel rim. They cover the area from the bead to the
side edges of the tread. As these sidewalls are exposed to abrasion, ultra-violet (UV)
and ozone degradation, the formulation of a sidewall will definitely contain antioxidants.
The body ply or casing provides the internal structure of the tyre and gives the tyre
strength and flexibility. It is usually made of natural rayon, nylon or polyester cords that
are coated with NR.

43

Chapter 2

Literature Review

The role of keeping the air inside the tyre is that of an integral part of the tyre which is
the inner liner. This is a double layer of synthetic rubber (usually butyl rubber) that
maintains consistent pressure which in turn helps to reduce the rolling resistance and
reduce energy consumption.
Each tyre manufacturer, in an attempt to generate tyres with unique characteristics that
can meet various performance standards and provide maximum comfort in various road
conditions (dry, wet and rutted) at various speeds, has its own special formula for the
tyre compound. However the ingredients are more or less similar and only the
proportions vary slightly depending upon the end use. These ingredients are
compounded and once the structure is assembled, the tyre is vulcanised.

Figure 2-5 Data from Shulman (2004) showing whole tyre usage in the EU in 2003.

2.4

Reuse of whole tyres & retreading

Whole tyres that have reached the end of their lives still possess unique physical
properties which make them suitable for several different applications. According to
Shulman (2004), scrap tyres are suited to installations where lightweight materials are
required to provide support on wet or unstable soil bases. Figure 2-5 shows the whole
tyre usage in EU in 2003 wherein the category other refers to usage as boat fenders.

44

Chapter 2

Literature Review

About 62% of these scrap tyres are used as noise barriers and this will increase further
as the EU directive on noise (EU Directive 2002/49/EC) is implemented. This
application of tyres has been successful predominantly in the Scandinavian countries
where there is an increased governmental pressure to reduce the noise adjacent to
motorways and residential areas. In aquatic environments whole tyres can attract fishes,
algae and corals. These form colonies within them because of their inert structure. In the
oceans the toxicity issues are mitigated by the large volume of the ocean, they are very
popular for building artificial fishing reefs, coastal protection systems and floating
breakwaters in many regions. Many more countries are now adopting this practice. One
of the longest fishing tyre reefs in the world, made of 3 x 106 tyres, is in Florida [De et
al. (2005)].
Retreading is where whole tyres can be reused and hence is considered to make the
best use of the waste tyre resource. This operation involves removal of the old tread of
the tyre mechanically by a process called buffing. A new tread is then applied by either
the pre-cure process where the tread rubber is already vulcanised prior to application or
the mould cure process where unvulcanised rubber is applied to the buffed tyre
following which the tread is vulcanised. According to Chatterley (2004), retreading is
only a temporary reuse option as the retreaded tyres will eventually be worn out and
then be disposed of. Retreading has suffered a marked downturn recently due to factors
such as high labour costs, the lower prices of new tyres especially tyres made outside
the EU [Myhre and Mackillop (2002)]. The business has also come under scrutiny due
to health and safety concerns. Retreading is considered a viable option because it is
about 40-50% cheaper than a new car tyre and the reason for this is that the crude oil
consumed in synthesis of a new tread (~7 gallons) is much less than that for a new tyre
(~22 gallons). It also minimises the number of tyres that would be disposed on an
annual basis which in turn leads to lower levels of new tyre production annually.
However, the number of tyres retreaded continues to drop in UK where the production
of retreads for passenger vehicles has fallen from over 7 million in 1996, to under 2
million in 2001 [Department of Trade and Industry (2001)]. According to a report by
Slater (2006) out of the 480000 tonnes of used tyres that were generated in 2004 in the
UK, approximately 10% were retreaded which is only 49000 tonnes.

45

Chapter 2

2.5

Literature Review

Recycling treatments and technologies

A review by Myhre and Mackillop (2002) points to the fact that the majority of the
technical papers on rubber recycling deal with scrap tyres because the tyre industry
produces the greatest volume of scrap rubber. Scrap tyres consists of blends of
elastomers that are vulcanised with strong chemical cross links that are intended to be
stable and almost irreversible. It is widely understood that cured rubber is difficult to
recycle. Figure 2-6 is a schematic of various treatments and technologies used to recycle
post-consumer tyres.

Figure 2-6 Schematic from Shulman (2004) showing the various treatments and
technologies used to recycle postconsumer tyres.

According to Shulman (2004) these treatments have evolved over the years such that
they attempt to exploit and enhance the tyre compound properties and do not require the
dissolution or melting of the rubber into the virgin compound. The schematic shows
four basic levels and these can be described as follows:
Level 1 Involves use of mechanical means to destroy the structure of the tyre.

46

Chapter 2

Literature Review

Level 2 Uses level 1 output as feedstock and involves treatments that help segregate
the main components of the tyre.
Level 3 This level comprises of treatments and technologies that modify
characteristics of rubber materials that is usually the output of level 2. The outputs of
this level can be used directly in applications or products.
Level 4 Materials that were modified in level 3 are refined further and upgraded in
this stage in order to increase their value.

Figure 2-7 SEM scan by Jacob et al. (2003) of rubber granulates (a) aggregates (b) single
particle.

The next few sections describe some of the methods of reusing rubber that have been
considered as long term solutions to the disposal issues.

47

Chapter 2

Literature Review

Figure 2-8 Use of granulates in the EU in 2003 [Shulman (2004)].

2.5.1 Granulates
As shown in the schematic (Figure 2-6) a grinding/size reduction operation falls under
level 2 treatments where the principal components that constitute a tyre are removed.
The tyre is usually cut into relatively large pieces and then ground into smaller pieces
(Figure 2-7). In this process none of the constituents of the rubber is altered or removed
[Burford and Pittolo (1982)]. Scrap rubber that is mechanically ground to a fine particle
size finds application in many areas as illustrated in Figure 2-8 [Shulman (2004)]:1) Play/sports surfaces and applications

Granulates are used in various soccer pitches, running tracks, artificial turf,
childrens playgrounds, equestrian arenas etc. They help reduce the severity
of injuries during play and improve safety. It is estimated that approximately
2,000,000 square metres of sports fields are built each year in the EU that
utilise approximately 50-80 kg/m2 of granulates. With EU legislation in
place that requires more than 150,000 primary school and municipal
playgrounds to be paved with shock absorbing materials, this is a potential
use of granulates.

48

Chapter 2

Literature Review

Car mudflaps, floor mats, carpet underlays, footwear, crash and noise
barriers are some of other products where rubber granulates are incorporated.

2) Industrial products and applications

Rubberised asphalt are used for road surfaces to reduce skidding, vibration,
noise, reflective glare and fog build up. Granulates are therefore popularly
used for this purpose in large quantities in USA, Canada and Australia.

Rubberised concrete is processed by mixing rubber granulates with concrete.


Addition of granulates to concrete results in a blend that has increased crack
resistance, reduced noise levels, lighter weight and greater resistance to rain
and acid.

Manhole covers are made from rubber granulates as a health and safety
measure for pedestrians. The rubber makes the cover non-conductive and
thus prevents it from becoming electrically charged by stray voltage that
has electrocuted pedestrians in the past.

3) Thermoplastic elastomers (TPE)

These are durable compounds that combine the toughness of rubber and
higher stiffness of thermoplastics (such as Polypropylene, Polyvinylchloride
and Polyethylene) by blending the two together. The new composite material
retains the material qualities of the rubber granulate and processing
behaviour of the thermoplastic. Addition of rubber granulates to plastics
improve certain key properties of plastics such as impact resistance and tear
strengths. TPEs are in great demand in the automotive, industrial and
construction sector.

However, Brown and Watson (2001) have commented that for each of these
applications there is a limit to the quantity of rubber that can be incorporated and hence
it is essential to develop new markets, products and applications. Ultimately several
factors contribute to the performance properties of rubber granulates including particle
size distribution, shape, surface morphology and surface functionality. In a feasibility
analysis study carried out by Sunthonpagasit and Duffey (2004), they reported that the
smaller the size of granulates, greater is the investment and operating cost. It is also

49

Chapter 2

Literature Review

difficult to generalise particle size requirements as different producers in the same


granulates market may require different sizes to produce their unique product.

Figure 2-9 Illustration of mesh screens taken from report by Evans and Evans (2006).

With the purpose of establishing a common nomenclature to discuss post consumer


waste tyres, the waste and resources action programme (WRAP) published a report that
was authored by Evans and Evans (2006). In this report, with reference to rubber
granulates, a screen is defined as a meshed wire fabric that is mounted on a frame for
the purpose of separating coarser from finer granulates. As illustrated in Figure 2-9, the
opening between the wires of a screen is a mesh and is defined as the number of linear
openings per inch (one inch = 0.254 m). Figure 2-10 is a particle size comparison chart.

50

Chapter 2

Literature Review

Figure 2-10 Mesh to microns particle size comparison [De et al. (2005)].

Figure 2-11 Sieves that are mounted on the sieve shaker [De et al. (2005)].

51

Chapter 2

Literature Review

Figure 2-12 Sieving process used to separate granulates [Evans and Evans (2006)].

A sieve is a device that includes the screen for separation of granulates and is shown in
Figure 2-11. These sieves are mounted on a sieve shaker which is then run for a fixed
period of time with rubber granulates placed on the top most screen (Figure 2-12). The
sieve with the screen having the largest opening is placed at the top with a progressive
decrease in the screen size down to the pan. At the end of the shaking operation, the
quantity of granulates on each pan is weighed and measured. The screen on top is
known as the zero screen and one below it is called the designation screen. The mesh
size is established using these two screens.
Popular techniques used for powdering rubber vulcanisates are ambient, cryogenic and
wet grinding.
Ambient grinding is a multistep technology where the material processing takes place
at or above ambient temperature. With this type of grinding, the vulcanised rubber is
passed through the nip gap of a two-roll cracker type mill or a high powered shear mill
at room temperature. The size reduction takes place due to the cutting and shearing
action of the granulator on the rubber. The mills (Figure 2-13) used consist of two large
52

Chapter 2

Literature Review

rollers with serrations cut in one or both of them that run at varying speeds. The particle
size and particle size distribution depends on (i) the period for which the material is run
on the mill and (ii) the number of grinds it undergoes. The higher the number of passes
through the mill, the greater is the extent of size reduction at the expense of increased
cost.

Figure 2-13 Schematic showing the roll design with corrugations on them [De et al.
(2005)].

Burgoyne et al. (1976) used a two stage mechanical grinding process where rubber
granulates with a maximum particle size of 100 mesh (~ 150 m) was achieved. In the
first stage a regular grinder was used and various irregular shaped granulates that were
approximately 5mm in diameter were obtained. These were then passed through a
second grinder with an integrated filter screen. Those granulates that could not pass the
screen were classified as coarse and hence subjected to the process repeatedly until they
were sufficiently small.
Cryogenic grinding or grinding in the glassy state has been described by Kohler and
O'Neill (1996) as a process where the rubber is firstly frozen with liquid nitrogen
following which it is shattered and ground. Rubber becomes brittle as a result of the
cooling and then it can be fractured to the desired shape and size. The heat generated in
this process is much less than with ambient grinding where a lot of shearing and cutting
is involved. This is the reason that the degradation of rubber due to heat build up is

53

Chapter 2

Literature Review

negligible in case of cryogenic grinding compared to ambient grinding techniques


where the granulates are partially oxidised.

Figure 2-14 SEM contrasting the typical surface structures of cryogenically ground
rubber granulates (left) and ambiently ground rubber granulates (right) [De et al. (2005)].

In case of cryogenic grinding the cost is dictated by the amount of liquid nitrogen that is
consumed in the process. However, as the rubber can be broken down easily when it is
embrittled, the machinery undergoes less wear and tear. The maintenance costs for a
cryogenic setup would therefore be less compared to an ambient setup. The efficiency
of cryogenic process is also higher than an ambient setup. It has been reported by
researchers such as Gibala et al. (1996) and De et al. (2005) that cryogenically ground
scrap rubber is smaller and more uniform in size as compared to their ambiently ground
counterparts. Scanning electron microscopy (SEM) examinations of the surface
(Figure 2-14) have shown that the ambient granulates have rough convoluted surfaces in
comparison to the smooth and angular surfaces of the cryoground rubber. Ambient
granulates are also reported to have greater surface area per unit volume than
cryogenically ground granulates. Researchers such as Burford and Pittolo (1982) and
Gibala et al. (1996) carried out experiments where they incorporated rubber granulates
into matrix rubber. They explained that as ambiently ground rubber had a more irregular
shape with a higher porosity than cryogenically ground rubber. The void space in
former was also higher. This enhanced the ability of these ambient granulates to retain
the new matrix rubber in its void volume. The formation of this mechanical key effect

54

Chapter 2

Literature Review

resulted in an increase in the effective granulate volume fraction that leads to better
adhesion between the matrix and granulates.

Figure 2-15 - SEM images of ground tyre rubber produced by cryogenic grinding (left)
and pulverised rubber granulates (right) [De et al. (2005)].

The particle size distribution will vary according to the method used for grinding the
rubber. Cryogenically ground granulates have broader particle size distribution
compared to ambiently ground granulates. Myhre and Mackillop (2002) highlighted that
all these differences could result in different properties of the end product when
granulates are incorporated into virgin rubber.
Shulman (2004) has reported that in order to benefit from the advantages and
characteristics of both these forms of grinding, they can be combined into a single
continuous system where the initial size reduction is done by ambient grinding followed
by cryogenic grinding where the metals and textiles are removed. This also helps reduce
overall costs.
Wet or solution grinding is the third method of grinding and according to Klingensmith
and Baranwal (1998), is a process that is not as commonly used as the previous two. In
this process, granulation is done in the presence of water using closely spaced grinding
wheels. Granulates thus obtained are more uniform and clean with a particle size that
depends upon the time for which the grinding was carried out. However, the
productivity of this process is lower and it is time consuming. Both ambient and wet

55

Chapter 2

Literature Review

grinding are operations carried out with the substrate still in the rubbery state.
Ultimately, irrespective of the grinding method used, their outputs can be used as a
feedstock for further processing.
In his paper Riggle (1995) has summarised the whole process of size reduction of a long
lasting product such as a tyre as cumbersome where equipments like magnets, blowers
and fibre screening systems are employed at various stages in order to remove the metal
and fibre fragments. These are preliminary steps that need to be carried in order to
produce rubber granulates from tyre rubber prior to further operations such as blending
and devulcanisation. Previous studies by Klingensmith and Baranwal (1998) and
Burford and Pittolo (1982) have shown that the extent to which this segregation needs
to be carried will vary according to the method used for grinding the rubber. The size of
these granulates produced can be varied and the individual particle size and particle size
distribution can all be tailored to the requirements of the end user.
2.5.2 Reclamation of rubber
Reclaiming has been described as the process by which the carbon-sulphur bonds
typically present in vulcanised rubber scrap is broken down by the application of energy
so that it can be reused. This breakdown is carried out by either devulcanisation
(thermally) or by depolymerisation (chemically) so that the reclaimed rubber can be
compounded, processed and revulcanised [Myhre and Mackillop (2002].
Reclaiming is the oldest method of reusing rubber. It was very popular in the beginning
of the 20th century as half the rubber consumed then was in the form of reclaim [Dierkes
(1996)]. By the middle 1980s, the worldwide polymer consumption in the form of
reclaim had reduced to less than 1% and this was attributed to the advent in technology
of radial and steel belted tyres. These technologies improved the durability of tyres and
also made tyre processing more difficult. The relative reduction in the cost of the rubber
(natural and synthetic) and the carbon black and an increase in their production capacity
also made virgin materials cost competitive with reclaim [Adhikari et al. (2000)].
On addition to virgin rubber, reclaim reduces the mixing time and power consumption
and hence can be a useful processing aid and can be revulcanised. It also ensures low
mixing temperatures. Its end product has been shown to be an excellent low cost
polymer that can be substituted at high levels in to less demanding products such as
56

Chapter 2

Literature Review

shoe soles and other elements of footwear. However, due to the processes involved,
reclaiming degrades the physical properties of the rubber and is considered to be a
highly complicated, cumbersome and costly process. Payne (1994) has explained that
factors such as heat build up, high hysterisis and poor wear are some of the
disadvantages that limits the application of reclaim for use in high-end products like
tyres.
It is believed that (a) newer legislation and (b) increase in fuel and petroleum derivative
costs for polymers might breathe life into this industry.
2.5.3 Pyrolysis Carbon black production
Pyrolysis is a process where the ground rubber is thermally decomposed to give the
various constituents such as hydrocarbon oil, carbon char, gases, carbon dioxide, carbon
monoxide [Myhre and Mackillop (2002)]. This is carried out in the absence of oxygen
and at temperatures below 400C the major product is carbon char. The relative
proportion of products produced varies as the temperature changes with higher
quantities of gases produced at higher temperatures. It is possible to produce clean oil
free char using this process and as shown in Figure 2-6, it is marketed commercially as
a colorant or filler. Further processing at level 4 is known to upgrade the char to
produce substitutes for carbon black with surface area, structure and distribution in
rubber compounds similar to that of carbon black [Shulman (2004)]. Commercialising
this has met with limited success because it is more expensive and has an unpredictable
composition that makes it unreliable. The oil and gas that is produced during the process
is also marketed (level 2 Figure 2-6 ). However, the process is sensitive to fluctuations
in the oil and gas market [De et al. (2005)].
2.5.4 Devulcanisation
Brown and Watson (2001) have reported that devulcanisation would be the ideal form
of recycling as unlike reclaiming, the technology would aim to cause minimal damage
to the polymer C-C backbone and yet totally or partially cleave the existing sulphur
crosslinks formed during the vulcanisation process. The feedstock for this process are
usually ambiently or cryogenically ground rubber.

57

Chapter 2

Literature Review

Ultrasonic devulcanisation

Review by Myhre and Mackillop (2002) indicates that a lot of work was carried out
in the 1980s in an attempt to use ultrasonic waves that would initiate the scission of
sulphur cross links. Work to improve the extent of devulcanisation using this
technique is going on at a lab scale.

Microbial devulcanisation

Certain microbes are able to selectively break sulphursulphur and carbon-sulphur


bonds in order to restore the flexibility and mobility of rubber molecules so that
there is better interparticle adhesion and better bonding with the matrix when
revulcanised. A review by Warner (1994) has reported that considerable study has
been carried out in the past on degradation of rubber by micro organisms that would
not get physically attached to the rubber surfaces. This technique too is being tested
at a lab scale and has not yet made the difficult transition as a valuable commercial
process.
2.5.5 Surface modification/activation
Granulates are blended into fresh rubber matrix in order to make use of the bulk
properties of these granulates that are rubbery in nature but Brown and Watson (2001)
believe that in order to achieve good performance characteristics, chemical modification
needs to be done at the surface to encourage the best interfacial properties. Surface
modification is therefore the surface activation of granulates in order to achieve better
adhesion between the fresh rubber phase (matrix) and granulates. The idea is to make
the surface of granulates more reactive by adding chemically reactive groups [Fesus and
Eggleton (1991)]. It has been suggested by Kim and Burford (1998) that increasing the
surface energy is vital for enhancing the adhesion. A report by Dierkes (1996) talks
about a surface modifying step that involves addition of a crosslinkable surface layer
consisting of a polymer and a curing system to the granulate as a result of which cross
linking with matrix is achieved. De et al. (2005) and Bauman (1995) have reported an
increase in the extent of interaction by the addition of carbon black and a surface
coating with polar functional groups respectively. One of the important factors that must

58

Chapter 2

Literature Review

be considered for any surface modification process is to retain economically viability


and thus not remove the financial incentive for recycling.
2.6

Energy recovery

According to Ahmed and Klundert (1994) there is a third method of recycling waste
tyres with the first being direct reuse (discussed in 2.4) and second being material reuse
(discussed in 2.5). Figure 2-16 is plotted using data from a report by Shulman (2004)
that clearly shows that the energy recovery from tyres in the EU is steadily increasing.
These are used as a fuel energy source for cement kilns or incinerators to generate
electricity and thus replace conventional oil, coal or gas sources. This tyre derived fuel
(TDF) is widely used as an industrial fuel. It is also one of the principle means of
dealing with waste tyres in countries such as Japan and the US. There are concerns
about using coal as a source of energy due to cleanliness, cost, reliability of supply and
CO2 emissions. TDF seems to offer a good alternative as it produces more energy per
weight in comparison to coal and generates less sulphur dioxide [Myhre and Mackillop
(2002)].

Figure 2-16 Data from Shulman (2004) showing the energy recovery from
post-consumer tyres in the EU.

59

Chapter 2

Literature Review

However, tapping energy from scrap rubber requires the installation of new feed
systems at power stations in order to accept rubber as a fuel source. This is extremely
expensive and acts as a barrier to accepting rubber as a fuel source [Kim and Burford
(1998)]. Other factors such as higher costs associated with meeting emission standards,
public concerns over burning of tyres and unreliability of tyre supply gives traditional
fuel sources an advantage [De et al. (2005)].
To summarise the various rubber recycling treatments and reuse options, the
volume of tyres retreaded is declining and each truck tyre can only be retreaded a
certain number of times after which it will still have to be disposed of. The process of
energy recovery from tyres and pyrolysis is not economical because of the initial costs
involved and there are also environmental concerns. Asphalt when coated with rubber
performs better than normal asphalt in the long run. However, it is not yet very popular
and rubberised asphalt is twice as expensive. All forms of devulcanising rubber are
expensive as they require additional energy and processing. These techniques are still in
their nascent stage and they are not yet suitable for commercial applications like tyre
manufacture. As mentioned previously, tyre manufacturing industry consumes 70% of
all rubber that is produced globally and therefore the scope for using recycled rubber in
non tyre products is limited.
Burford and Pittolo (1982) have pointed out that unlike pyrolysis, reducing vulcanised
rubber into granulates is a non polluting option. Also most of the processes associated
with recycling scrap tyres involve at least a two phase process wherein grinding the
scrap rubber is the first step anyway. The blending of rubber granulates with virgin
rubber has met with limited success. However, the use of rubber granulates for
discovering new applications has a lot of potential because (a) the cost of grinding scrap
rubber is much lower than techniques such as devulcanisation. (b) when scrap rubber is
converted into granulates, a material with modified surface is obtained. This
new/modified surface can then be cross linked with each other or to a virgin rubber or a
different thermoplastic material. Material recycling is therefore emerging as a
technologically, commercially and economically viable option for the future of the
sustainable waste management industry. It is for this reason that in this study the focus
is on the use of rubber granulates that have been ambiently ground and which are
blended with virgin rubber in order to identify difficulties and to optimise the process.

60

Chapter 2

Literature Review

With efficient granulation and reincorporation into virgin material, the properties of
rubber granulates could be put to good use. Thus granulation has the potential to be the
best technique by which a tyre, which is a good resource of hydrocarbons, can find its
way back into the production cycle.
2.7

Incorporation of granulates into rubber vulcanisates

The effect due to incorporation of granulates into rubber compounds, so that they can be
used as a filler in rubber compounds, has been an important area of study for researchers
for the last three decades. Kim and Burford (1998) commented that rubber granulates
are unique as compared to conventional fillers such as carbon black and silica because
of their relatively large size and low modulus. Previous research by Phadke et al. (1984)
has shown that when rubber granulates (that have been vulcanised once) are
incorporated into an uncured rubber sheet which is then subjected to a curing cycle,
these granulates are in effect undergoing a second cure cycle. Hence for the
vulcanisation of the whole matrix-granulates blend to be complete, additional curatives
are required and an optimal curing mix needs to be identified taking into account the
quantity of granulates added into the matrix.
Hilyard et al. (1983) proposed that rubber vulcanisates with rubber granulates
incorporated into them can be treated as rubber-rubber composites where the matrix
(virgin rubber) was the continuous phase and granulates constituted the dispersed phase.
The following variables then affected the mechanical properties of these composites:
1) Granulate composition and their size distribution.
2) The method used for grinding.
3) The matrix rubber type and its composition.
4) The curing conditions for preparation of the composite.
In their previous work, Gibala and Hamed (1994) had observed that there was a
decrease in scorch time and maximum rheometer torques with the addition of ambiently
ground black filled SBR vulcanisates to SBR compounds. Similar observations were
also made by Phadke et al. (1984) when they incorporated rubber granulates that were
ground to 40 mesh (~420 m) cryogenically and which were incorporated into a NR
matrix. Gibala and Hamed (1994) believed that the torque reduction was due to the

61

Chapter 2

Literature Review

transfer of sulphur from matrix rubber into the ground vulcanisate. This in turn
triggered the release of some fraction of bound accelerator from these granulates that
diffused into the matrix resulting in a decrease in scorch time and hence speeding
vulcanisation. Sulphur diffusion from matrix into granulates caused further crosslinking
of granulates and this in effect weakened them as their extensibility was now reduced.

Figure 2-17 - The effect on torque with loading of granulates [Han and Han (2002)].
1 lbs.in = 0.113 Nm. Tmax: maximum torque; T40: scorch time.

Figure 2-17 plots the findings of Han and Han (2002) whose initial rheometry tests
show a decrease in maximum torque (Tmax) and scorch time (T40) with higher
incorporation of granulates. Their results therfore agreed well with previous results
established by Gibala and Hamed (1994). Whilst studying the effect on the Mooney
viscosity, Phadke et al. (1984) reported that the addition of granulates caused an
increase in the viscosity and this was probably due to the addition of crosslinked

62

Chapter 2

Literature Review

granulates. Gibala et al. (1996) found that composites that contained ambiently ground
rubber had a higher viscosity than their cryogenically ground counterparts. They
attributed this to the mechanical key effect formation that was discussed in section
2.5.1.

Figure 2-18 SEM contrasting fracture surfaces of similar SBR matrix with granulates
(volume fraction 25%) containing either 75 phr CB (left) or 12.5 phr CB (right) [Burford
and Pittolo (1984)].

2.7.1 Fracture behaviour


The fracture behaviour of the granulate filled composites has been found to depend
upon the relative modulus of the granulates incorporated into the matrix. Burford and
Pittolo (1984) observed that for the same matrix, when the modulus of the granulates
that were incorporated was varied, there was a difference in the fracture surface. Figure
2-18 shows that the incorporation of a lower modulus granulates produces a smoother
fracture surface when the matrix and the volume fraction of granulates remains the
same.

63

Chapter 2

Literature Review

Figure 2-19 - Schematic of a rubber particle embedded in rubber matrix [Gibala et al.
(1998)].

Using Figure 2-19, Gibala et al. (1998) illustrated the fracture behaviour of a composite
into which a single ground vulcanisate was incorporated. I, is the interface created when
a ground vulcanisate, P, is mixed into cured matrix, M, thus creating a composite M-I-P.
The possible sites then where fracture could initiate were the interface, the matrix or
within the particle itself. It was pointed out that the interfacial fracture would take place
if bonding between M and P was limited and as P was already vulcanised, it was likely
that chains of M and P would not interpenetrate over a sufficient distance. However,
Gibala et al. (1998) concluded at the end of this study that in the case of ambiently
ground rubber, grinding methods caused internal damage to the granulates and hence
there was a deterioration in the properties of the overall composite. For cryogenically
ground rubber, however, interfacial adhesion was poor and this was illustrated using
optical micrographs (Figure 2-20).

64

Chapter 2

Literature Review

Figure 2-20 - Optical micrograph showing (a) Interfacial failure with a cryogenically
ground particle (b) fracture within an ambiently ground particle [Gibala et al. (1998)].

The fracture behaviour of filled SBR and NR compounds containing ground rubber
vulcanisates have been investigated in detail by Han and Han (2002). The examination
of the fracture surfaces showed that the SBR compounds with no granulates had a very
smooth fracture surface and these got visibly rougher as more granulates were
incorporated into the matrix. This roughening of surface indicated tear deviation that in
turn helped the SBR composite to resist and delay the crack growth. However a similar
effect was not seen in case of NR as the addition of granulates into NR did not alter the
fracture surface. The investigators proposed that after incorporation of granulates, the
crack growth rates of SBR and NR compounds were similar even though with no
granulates SBR compounds had a lower crack growth resistance than NR. They
believed that this effect was seen because addition of granulates into NR suppressed
latters strain induced crystallisation which in the first place was usually responsible for
its higher ability to arrest a crack. Their study revealed that effect due to particle size of
granulates was not significant. If tear strength of filler is high and the fracture surface is
rough then it was indicative of the fact that catastrophic failure occurred within the

65

Chapter 2

Literature Review

matrix. [Burford and Pittolo (1984)] On the other hand if tear strength of filler was
lower than that of the matrix, then tear deviation will not occur and hence the fracture
topography would be flat. This implied that the propagating tear would not deviate at
the matrix-filler boundary. In their quest to establish if latex rejects from the health
sector could be reused as fillers in SBR matrix, Mathew et al. (1996) used SEM to study
the morphological features of the fracture surface. The scans revealed the filler particles
as phase separated entities which meant that the NR granulates did not blend well with
the SBR matrix. However, the tear lines of the fracture surface seemed to indicate an
improvement in the mechanical properties of the SBR matrix due to the NR granulates.
2.7.2 Adhesion at the interface
As discussed in section 2.5.1, finer granulates result in a higher surface area to volume
of granulate ratio which in turn results in a better adhesion between the matrix and the
granulates. This would improve the tensile behaviour of the matrixgranulate
composite. Researchers like Burford and Pittolo (1983) and Kim (1999) found that
stretching of matrix-granulate composites resulted in dewetting of granulates that
eventually resulted in voids of varying sizes. As these composites were strained further,
voids around the interface grew until the compound failed completely. They concluded
that if these granulates were of smaller size, then the voids formed at the interface
would have been smaller which meant that the compound could have undergone higher
strains before failure. However, according to them, the end result remained the same
irrespective of the granulate size and morphology.
Hilyard et al. (1983) used granulates that were ambiently ground from whole tyres, the
exact composition of which was not known. They mixed these granulates from tyres
into gum SBR and NR and observed an increase in the initial modulus values of the
entire composite. This indicated that there was some degree of bonding between the
matrix and granulates. In the case of the SBR composite they noted that:

Along with the increase in modulus there was an increase in tensile strength (b)

As higher quantity of granulates were incorporated, the quantity of curatives


required in order to achieve the same hardness, decreased.

66

Chapter 2

Literature Review

This was probably due to some interfacial adhesion between the matrix and granulates.
As the quantity of curing agent was increased, with the quantity of granulates kept
constant, an increase in the initial elastic modulus of the composite was seen and
apparently this was due to the increase in crosslink density of the matrix material.
Burford and Pittolo (1984) also reported an increase in the overall composite modulus
and a corresponding decrease in elongation at break (b) for granulate volume fractions
of 25% when there was an increase in the filler hardness.

Figure 2-21 - Schematic of fracture process in composites due to (a) cohesive failure in the
matrix (b) adhesive failure between matrix and granulate [Hilyard et al. (1983)].

Hilyard et al. (1983) identified two failure mechanisms that are illustrated in Figure
2-21 for SBR matrix that contained scrap rubber granulates. Cohesive failure of the
matrix material adjacent to granulates took place when the quantity of sulphur in the

67

Chapter 2

Literature Review

matrix was low. The failure was marked by the formation of vacuoles that grew with
increasing stress. When sulphur quantity was high, adhesive failure took place as
strength of the matrix exceeded that of the interface. Unlike cohesive failure this form
of failure was more instantaneous and took place at the interface.

Figure 2-22 - SEM scan showing fracture surfaces of a specimen of a rubber compound
with 9% (by weight) cryogenically ground granulates [Burford and Pittolo (1983)]. The
demarcation that separates the top and bottom fracture surfaces can be seen.

Researchers such as Burford and Pittolo (1983), Phadke et al. (1984) and Kim (1998)
have used SEM scans in the past to attribute the loss of physical properties upon
addition of rubber granulates into virgin rubber, to a poor adhesion between the matrix
and the filler. Figure 2-22 depicts the fracture surface of a rubber compound that
contains 9% cryogenically ground granulates by weight with an average particle size of
400 m. Upon comparing this with the fracture surfaces in Figure 2-23 which is that of
a rubber compound that contains 9% ambiently ground granulates by weight with an
average particle size of 200m, it becomes evident that though the ambiently ground

68

Chapter 2

Literature Review

granulates were smaller in comparison with the cryogenically ground granulates, when
9% granulates were incorporated by weight, the extent of roughening was the same.

Figure 2-23 - SEM scan showing fracture surfaces of a specimen of a rubber compound
with 9% (by weight) ambiently ground granulates [Burford and Pittolo (1983)]. The
demarcation that separates the top and bottom fracture surfaces can be seen.

Burford and Pittolo (1983) observed angular craters in some of their SEMs which
seemed to indicate dewetting phenomenon which was indicative of poor adhesion
between the matrix and filler. In case of good adhesion between the matrix and the
granulates, filler rupture may initiate the final rupture of the composite. For such a case
very few signs of dewetting were observed and the strain energy density at break (Ub) of
the composite had a linear relation with that of the granulate. They also pointed out that
the extent of bonding between the granulate and the matrix will determine the portion
on the matrix-granulate composite where maximum stress is concentrated. If the
granulates are harder than the matrix and well bonded, large stress concentrations occur
at the polar interfaces. This means that the stress experienced by the composite at this

69

Chapter 2

Literature Review

particular interface would be greater than the average applied stress. This results in an
interfacial failure at the pole possibly leading to a premature catastrophic failure. The
equatorial interface would have to bear maximum stress if the filler particle is
unbonded. When the granulate is softer than the matrix, the polar interface would
experience a stress that is less than the applied stress. Hence it was concluded that the
strength of the matrix-granulates composite would be significantly affected if the
interfacial adhesion was low and/or the granulate was excessively hard [Burford and
Pittolo (1984)].
Kuznetsova et al. (2004) have carried out a detailed study of the effects of ground tyre
rubber on the elastic modulus, elongation at break, tensile strength of the virgin rubber
granulate blends. They reached the same conclusion as Burford and Pittolo that the
rubber type, granulate size all play a role in the scale of the interfacial interaction.

Figure 2-24 - Changes in Ultimate tensile strength and elongation at break with addition
of scrap rubber into parent rubber compound [Burgoyne et al. (1976)].

70

Chapter 2

Literature Review

Figure 2-25 - Effect of granulate quantity on tensile properties of the composite [Han and
Han (2002)]. 1 Kg/cm2 = 0.098 MPa.

2.8

Strength/Crack growth phenomenon

According to Burgoyne et al. (1976), in the past scrap and flash were accepted as an
unavoidable expense associated with part manufacture. As illustrated in Figure 2-24,
their study confirmed that addition of granulates brought about a reduction in the
ultimate tensile strength (UTS) and elongation at break (b). They observed that even at
low levels of addition, there was a significant drop in the UTS. Their conclusion was
that the granulates could be used as a cost reducing filler which was also the conclusion
that Hilyard et al. (1983) came to based on their own experiments. Similar observations
were made by Han and Han (2002) who also noted a decrease in tensile strength and
modulus with an increase in quantity of granulates incorporated (Figure 2-25) and size
as well (Figure 2-26). However, Swor et al. (1980) had developed in their laboratories a

71

Chapter 2

Literature Review

process for production of granulates whose incorporation into rubber compounds up to a


quantity of 50% by weight did not bring about any performance deficiencies. These
granulates were ultra fine with 90% of the particles (by volume) having an average
diameter of less than 20m.

Figure 2-26 - Effect of average particle size of granulates on tensile properties of the
composite [Han and Han (2002)].

Research work by Gibala et al. (1999) showed a drop in tensile strength wherein
approximately 30 phr of granulates of the size 2030 mesh were incorporated. The
composition of these particles was known as these were ambiently ground from cured
sheets of a known composition. The modulus of granulates was higher than that of the
matrix but their extensibility and tear strength was lower than latter. The investigators
proposed that the drop in tensile strength was a result of multiple cracking that the
granulates underwent during the tensile test which caused it to act as a stress raising
flaw. They explained that as the extensibility of granulates was low, during the tensile
72

Chapter 2

Literature Review

test the particle cracked internally at a relatively lower strain imposed by the matrix.
With further extension of the composite, the deformations imposed by the matrix on the
particle increased as a consequence of which the stress concentration in the matrix in the
vicinity of the particle became sufficient to result in catastrophic failure.
2.8.1 Tear strength

Figure 2-27 - Schematic of a trouser test specimen.

The tearing energy (T) which is also sometimes known as the strain energy release rate
is the energy required to propagate a cut in a specimen by a unit area in size. It was
defined mathematically by Rivlin and Thomas (1953) as:
U
T =

A l

eq. ( 2-1)

where, U is the total elastic strain energy stored in the sample and A is the area of one
fracture surface of the specimen. The partial derivative indicates that the specimen is
held at constant length. This tearing energy is independent of the overall shape of the
test piece and its relationship to the tearing rate is a characteristic of the rubber itself.

73

Chapter 2

Literature Review

The tearing energy can be determined using a trousers tear test piece shown in Figure
2-27. The equation used to compute T is:

T=

2 F
wW
t

eq. ( 2-2)

where, F is the average tear force, t is the thickness of each leg, w is the width of each
leg, is the extension ratio and W is the elastic strain energy density in one of the legs.
W can be determined using an area integral under a stress extension ratio curve of the
same material at the appropriate stress. Thus the equation takes into consideration the
energy stored in the bulk of the material due to its large strain. In the past, Gent (1978)
showed that reinforcing fillers such as carbon black increase the strength of the
elastomer into which they are incorporated. This could happen by either of the
following mechanisms:

An obstruction of crack path to circumnavigate the filler which causes a crack


deviation.

An increase in the energy dissipated by viscoelastic effects.

Han and Han (2002) observed an increase in energy dissipation with an increase in the
quantity of ambiently ground rubber granulates (derived from scrap tyre) incorporated
into SBR matrix. They believed that the rubber granulates obstructed and delayed the
crack growth and as a result the incorporation of rubber granulates resulted in superior
fracture resistance. Earlier studies by Burgoyne et al. (1976) indicated that when the
granulate volume fraction was less than 10% and the matrixgranulate composition is
the same, then there is virtually no change in the tear strength of the matrix-granulate
composite. A slight increase in tear strength with decreasing granulate size was also
observed. According to Gibala et al. (1999) for tensile tests, the stress to break (b) is
determined by the single largest flaw present, which is not the case for tear strength. In
case of a trouser tear test, the stress concentration is only at a small portion at any given
time which is generally the tip of the crack and the manner in which this crack advances
may be:

Steady This results when there are only small fluctuations in the tear force or

Stick slip This results when there are large force fluctuations during the test.

74

Chapter 2

Literature Review

These force fluctuations would occur between the initiation of the crack and its arrest.
Sometimes cracks deviate to a path that is perpendicular to the primary crack path, this
effectively blunts the tear tip and requires the new crack to be reinitiated along the test
piece. This in effect roughens the surface as tearing takes on a stick slip behaviour as
opposed to steady tearing. Thus according to Gibala et al. (1999) the effect on the tear
strength of an elastomer when a discontinuity is introduced would depend on the extent
of crack path deviation that results during tearing.
Pittolo and Burford (1984) explained how the hardness of the granulate and bonding
between the matrix and granulate are the two factors that determine the tear strength of
a granulate filled elastomer. If the interfacial adhesion is high, then the tear strength
might increase with increasing hardness of the granulate as harder granulates might
cause greater tear path deviations. However, if interfacial adhesion is low, then the
matrix-granulate interface becomes more important and the tear strength might
decrease. The examination of tear topography on the fracture surface could help analyse
the tear strengths and thus help determine which factor dominates. So for example, if
the tear topography is rough and the granulate is hard, then tear strength of the overall
composite is expected to be high because the hard filler is expected to cause crack
deviation. However, if the composite tear strength is low, then it can be attributed to the
poor matrix-granulate adhesion. On the other hand, if the tear topography is smooth and
the granulates are very soft, then the overall tear strength of the composite is low and
that is because of the poor tear strength of the granulate.
2.8.2 Interfacial strength

While studying strength properties of rubber-rubber blends, an important factor in


determining the strength of these blends is the interfacial bonding between them [Chang
and Gent (1981)]. Peel tests (Figure 2-28) are commonly used to determine the strength
of adhesion between two adhering layers. The work P of separation per unit area of
interface is given by:P=

2 F
tW
w

eq. ( 2-3)

where F is the average peel force, w denotes the width of the peel strips, t is the
thickness of each strip. W is the elastic strain energy density in the stretched strips.

75

Chapter 2

Literature Review

Burford and Pittolo (1984) used peel tests to determine that matrix-granulate composites
that had the highest interfacial adhesion also had the highest relative UTS and
elongation at break (b).

Figure 2-28 - Schematic of a peel test specimen.

While investigating the effect of granulate modulus on granulate to matrix adhesion,


Burford and Pittolo looked at two types of specimens: Type I was unfilled masterbatch
of SBR whereas type II masterbatch had 50 phr industry reference carbon black
(IRB#5). Strips of granulate compounds of varying degrees of hardness were prepared.
So type I peel specimen consisted of a strip of type I masterbatch (which had no CB)
and a strip of the filler compound sandwiched together. The peeling process showed a
cohesive failure of the type I masterbatch for all type I specimens which implied strong
relative interfacial adhesion and the peel strength values increased with granulate
modulus. Whereas in case of type II specimens which contained a strip of type II
masterbatch and a strip of the filler compound sandwiched together, the peel strength
was a fraction of the tear strength of the type II masterbatch indicative of a weak
interface. Thus they concluded that with a matrix-granulate system, the failure
mechanism would depend on the relative modulus of the matrix and granulate.
76

Chapter 2

2.9

Literature Review

Intrinsic flaws (crack initiation)

All polymers contain flaws or inhomogeneities randomly distributed within them that
act as stress raisers. These may be introduced during polymer processing, perhaps due
to improper mixing, handling or while preparing the test specimen [Gent (1978); Hamed
(1983)]. When a stress is applied on a body, the presence of these flaws results in stress
magnification such that the local stress experienced is much greater than the applied
stress which can result in premature failure [Thomas (1966)]. These flaws are typically
about 25-50 m in size and are not visible to the naked eye until they are about 10 times
the above magnitude. According to Gent et al. (1964) this is the reason that the
impending failure of a rubber component cannot be picked up at an early stage by
observation. If a razor cut is introduced into a material and the size of this cut is less
than the intrinsic flaw size (IFS) of the material, then the strength of the material
remains unaffected. As, in such a case the stress raising effect of the intrinsic flaw is
higher than that of the razor cut.
Gent et al. (1964) examined the fatigue life of dumbbell shaped specimens that had no
intentional flaws introduced. They observed that the die stamping action introduced
some flaws at the cut edge but there were also visible flaws along the narrow central
portion of the dumbbell. They concluded that fatigue failure in the absence of heat build
up resulted from these flaws which act as stress raisers. Variations in cross linking
density or presence of hard particles in the matrix like dust have been found to cause
inhomogeneities that also affect strength and the fatigue life eventually resulting in
failure.[Bhowmick (1986)]
Roland (1993) has reported that intrinsic flaws are also significant in the health care
sector because a lot of latex rubber products are used in the form of surgical gloves,
tubes and condoms. These are required to prevent transmission of very minute particles.
Barrier rubber products are responsible for preventing the spread of viral diseases like
AIDS and hepatitis. Microscopic observations have suggested the presence of flaws
which reduce the barrier performance of rubber films which might result in the passage
of viral sized particles through them. Gent (1978) has also reported that flaw sizes are
independent of the elastomer type or mix formulation even though these factors greatly
alter the strength properties of an elastomer. From his calculations, Hamed (1983)
recorded higher IFS values (~120 m) for lightly cross linked materials as compared to
77

Chapter 2

Literature Review

highly cross linked material (~60 m). His work showed that the work to break
decreased for highly crosslinked networks as the crack tip radius at the limit is
determined by the molecular length between crosslinks. Sharper cracks will act as a
higher stress raiser and fail more easily. All samples irrespective of their cross linking
levels contain flaws of similar dimensions and the ability of a material to arrest these
flaws and thus prevent them from propagating varies with the crosslink density.

Figure 2-29 - Schematic of a specimen with an edge cut [Rivlin and Thomas (1953)].

For elastomers incorporating granulates, a discontinuity will become a critical stress


raiser if it is larger than the IFS of the material [Gibala et al. (1999)]. This would lower
the strength of the composite and initiate fracture. There is a general consensus that the
influence of intrinsic flaws needs to be broadly examined as their minimisation will help
improve the strength and fatigue life of rubbery solids. Continued research in this area
would help understand mechanism of reinforcement of rubber. As described by Gent
and Hwang (1990), there are two ways by which crack initiation can take place. The

78

Chapter 2

Literature Review

first of which is from the naturally occurring flaws and the second one due to dilatant
stress that leads to internal cavitation type fracture. This is discussed in more detail in
section 2.10. In the following sections two crack geometries are discussed and
reviewed.
2.9.1 Edge cuts

Rivlin and Thomas (1953) showed using fracture mechanics for a tensile test specimen
with a cut of length c in the edge of the specimen as shown in Figure 2-29, that the
tearing energy, T is given by:
eq. ( 2-4)

T = 2K()Wc

This expression was derived from dimensional considerations that a small crack of
length c in a specimen of thickness, t that is in simple extension, would reduce the strain
elastic energy (U) by a factor that is proportional to c2.
U0 Uc = U = K()c2tW

eq. ( 2-5)

where the total elastic strain energy in the test piece without a crack is given by U0 and
Uc is total elastic strain energy in the test piece with a crack of length c. W is the strain
energy density in the bulk of the test piece (away from the cut) when the cut grows. K()
is a dimensionless function of strain and is expressed as [Greensmith (1963)]:
K ( ) = (1 + )

1 / 2

eq. ( 2-6)

If w0 and l0 are the width and length of the test specimen in the un-deformed state, then
this relation assumed that c<<w0 and c<<l0. Equation 2-5 when differentiated with
respect to the area of a single surface of the crack Ac (=ct) gives equation 2-4 which is
valid for all highly elastic rubber materials even those with a non-linear stress strain
relationship. From equation 2-6 it is evident that at small strains, K() approaches the
classical elasticity value of which implies that equation 2-4 then becomes the strain
energy release rate for linear fracture mechanics. Greensmith (1963) found the variation
of K() with extension ratio in the bulk of the test piece, empirically (Figure 2-30).

79

Chapter 2

Literature Review

Figure 2-30 Experimental findings of Greensmith (1963).

Now if b is the engineering breaking stress and E is the Youngs modulus, then for a
linearly elastic material [Gent (1978)]:
W=

b2
2E

eq. ( 2-7)

80

Chapter 2

Literature Review

In case of a linearly elastic material, W is the input energy per unit volume of the sample
whilst for a highly hysteretic material, there will be some energy dissipation due to
which the value of W would be less than the input energy.
Using equation 2-7 in equation 2-4 the relation between the engineering breaking stress
and the cut length is obtained as:

b =

ET
K ( )c

eq. ( 2-8)

So here it is seen that applied breaking stress varies inversely with the depth of the flaw
(cut length) and in proportion to c-1/2. Equation 2-4 is valid for all linearly elastic
materials, where all the input energy is stored reversibly and is available to be released
as a cut grows. For rubbery materials equation 2-7 and therefore equation 2-8 are
approximate.
2.9.2 Intrinsic flaw size (IFS) estimation

Thomas (1966), Gent (1978), Choi and Roland (1996) have reported that by
extrapolation of fatigue lives of test pieces with cuts of various sizes in them to fatigue
lives with no cut, the IFS can be estimated for a particular material. It has also been
established in the past that it is possible to estimate the IFS of a rubbery material by
using equation 2-4 directly on a tensile strip specimen tested to break. Hamed (1983)
has studied the effect of cut size on tensile properties of polymers by analysing the
double log plot of work to break per unit volume (Wb) versus cut length. For this
purpose, intentional precuts in the form of razor cuts, perpendicular to the tensile test
direction were introduced on dumbbell shaped rubber specimens prior to testing.
The plot (Figure 2-31) consisted of a straight line with a slope of -1. In order to estimate
the IFS, these lines were extrapolated to values of work to break per unit volume, Wb,
indicated by the arrows in Figure 2-31, when no intentional pre cut was present. A
similar approach has also been reported by Gent (1978) in which strength values, b, for
different edge cut sizes were extrapolated to the b value for a test piece with no cut
introduced at all in order to give the corresponding flaw size (cut length).
Using these techniques most researchers have shown that IFS values are equivalent to
edge crack lengths in the range 25 to 50m. This generic range seemed to indicate that

81

Chapter 2

Literature Review

between these limits the IFS is independent of the particular rubber type and of the
detailed compound formulations, although it has been shown that these greatly affect
the strength.

Figure 2-31 Plot showing the effect of precut length and crosslinking on work to break
per unit volume. The crosslinking was done via irradiation [Hamed (1983)].

2.9.3 Corner cuts

Flaws are random and therefore there is a possibility that when a tensile test specimen is
stretched to break, the largest flaw maybe located within the specimen or near the long
edge of the specimen. Thus, Busfield (2000) and Busfield et al. (1999) explained that
the failure causing flaw is unlikely to be located near the short edge (discussed in 2.5.1).

82

Chapter 2

Literature Review

The geometry of a corner crack is shown in Figure 2-32 where the crack length c is the
perpendicular distance from the crack front to the corner:
c=

a b
a 2 + b 2

eq. ( 2-9)

Drawing parallels with the relationship for an edge cut in a tensile test specimen,
Busfield et al. (1999); Busfield (2000) reported that for a corner cut:
T = 2 K ' ( )Wc

eq. ( 2-10)

where K ' () is a dimensionless function of strain similar to the functional relationship


for K () that is shown in equation 2-4. Figure 2-33 is plot that shows K ' () as a
function of strain and was obtained by Busfield (2000) where the relationship for K ' ()
was derived using finite element analysis (FEA). Comparison of Figure 2-30 and Figure
2-33 shows the similarity in the functional relationship. As the volume of the elastomer
being relaxed with an increase in crack length was higher in case of an edge crack as
compared to the tetragonal shape of a corner crack, the magnitude of K (), as expected,
was greater than K ' ().

Figure 2-32 (a) A schematic showing the corner cut in a tensile test specimen
(b) A schematic showing the calculation of the corner cut length c from the fracture
surface [Busfield et al. (1999)].

83

Chapter 2

Literature Review

Figure 2-33 Findings of Busfied (2000) for a corner cut in a tensile test piece.
Comparison with Figure 2-30 shows that the shape of the function with extension is
similar.

84

Chapter 2

Literature Review

2.10 Volume inflation


2.10.1 Dilatation and hydrostatic stress

Volume change measurement of an elastomer that is subjected to uniaxial stretching is a


sensitive technique for measuring changes in density of the elastomer under strain. It is
accepted widely as an important measurement and yet the data on this subject is limited.
When tensile stresses are applied to a body, the work done is due to two principal
components [Gee (1946)]:
(a) Work done by shear stresses that bring about a change in shape without any
change in volume.
(b) Work done by hydrostatic component of the stress that brings about a volume
change.
For the case described, the hydrostatic component that produces a volume change is
referred to as the dilatational behaviour. For rubber materials, the effective hydrostatic
tension is one-third of the tensile stress which means that the shear stresses dominate
over its hydrostatic counterpart. This implies that it does not contribute significantly
towards the tensile stress.
Gee (1946) derived the following expression for volume change based on
thermodynamic arguments using bulk modulus, B:

V
1
K
=
d or V =

3
3B 1
V

l l
l0

dl

eq. ( 2-11)

Pr ,Te

Where, f is the force required to elongate the material from l0 to l and K is the bulk
compressibility. Gee derived this expression by making the assumption that the material
would remain isotropic in the stretched state. This expression is appropriate for strains
of up to 100%. Thus if one had the stress strain data for a material, the volumeextension response using the above equation could be calculated.
While addressing the significance of the volume change problem, Gee et al. (1950)
confirmed by means of quantitative measurements that finite yet small expansions took
place even at small extensions. They worked with peroxide cured NR vulcanisates
because presence of zinc oxide and other inert fillers were known to result in volume

85

Chapter 2

Literature Review

changes from the formation of vacuoles around the particles. Mullins and Tobin (1957)
compared the volume change curves obtained from natural rubber vulcanisates that
were peroxide cured with carbon black filled vulcanisates that had 5 parts of zinc oxide
per hundred parts of rubber. They found that the increase in volume under strain was
small for vulcanisates reinforced with carbon black.
Gee et al. (1950) used the method of hydrostatic weighing in order to measure volume
changes at relatively small strains and Figure 2-34 is a schematic of the stretching rig
used. Using the known value for the bulk modulus of the rubber, they measured
experimentally and calculated the expected volume change resulting from this
hydrostatic component of the tensile stress and they found that their experimental data
agreed well with the calculated values. They postulated that this volume change
mechanism was applicable to any isotropic body and therefore was applicable to rubber
at low strains.

Figure 2-34 - The stretching frame used by Gee et al. (1950) in their study (a) Elevation (b)
Plan.

86

Chapter 2

Literature Review

In this case when the rubber vulcanisates were stretched to just over 100% strain, the
observed volume changes were of the order of 0.01%. They found that the work done
by applied force for an extension of 50% on one of their specimens was 10,000 times
that due to the volume change mechanism. This finding confirmed previous findings by
Gee (1946) that the volume change mechanism did not contribute significantly towards
total applied stress. This was attributed to the fact that the bulk modulus of rubber is
always much higher than the Youngs modulus typically by a factor of 104. Hence, the
strains that would accompany tensile stress would be higher by a factor of 104 when
compared to the dimensional volume changes due to a hydrostatic tension.
Mullins and Tobin (1957) carried out similar work with NR vulcanisates cured with
peroxide but to higher strains. They used the hydrostatic weighing technique as well and
as they were working with carbon black reinforced vulcanisates they used a more robust
rig (Figure 2-35). They noted a continuous change in weight of their specimens that
were soaked in water with time for the first 24 hours. This was due to very small
quantities of water being absorbed by the rubber and therefore they immersed their
specimens in water for 3 days prior to commencing their experiment.

Figure 2-35 - Sketch of the stretching rig devised by Mullins and Tobin (1957).

87

Chapter 2

Literature Review

Figure 2-36 - Change in volume with extension of carbon black filled natural rubber
[Mullins and Tobin (1957)].

They found that the initial increase in volumetric strain was followed by a very broad
region (Figure 2-36) that contained the maximum. This was followed by a volumetric
reduction with a further increase in strain. Their explanation for this trend was that at
low strains the increase in volumetric strain arises due to the hydrostatic component of

88

Chapter 2

Literature Review

stress. However above a certain strain, strain induced crystallisation that is present at
large strains in NR results in a significant volume reduction.
This effect was seen at extension ratio of 3.5 and once again the relative volume
changes observed were of the order of 10-4. Fedors and Landel (1970) commented on
the finding that the NR crystallisation might not necessarily have started from the point
where the maximum occurred in Figure 2-36. As the volumetric strain observed was of
a very low order of magnitude, small effects of crystallisation could have major
influence on the volume change. It would therefore be important to know the strain
interval for which the effects due to crystallisation would be negligible and thus could
be safely ignored.

Figure 2-37 - Results of experiments by Christensen and Hoeve (1970) for polyisoprene
where volume change is plotted against extension ratio. The dotted line represents the
calculated volume change using equation 2-12.

89

Chapter 2

Literature Review

Christensen and Hoeve (1970) calculated volume change on extension by using the
following expression:
V = KfL(1 + + 2 ) 1

eq. ( 2-12)

where, f is the retractive force and the compressibility, K is given by:


K =

1 V

V p

eq. ( 2-13)

The experiment was carried out using a dilatometer where rings were attached and
extended. They noted that even though at low extensions, the calculated and
experimental data agreed well, at higher extensions, there was a noted difference
(Figure 2-37). The equation was derived using the molecular theory of rubber elasticity
which assumes that the free energy of strained rubber (F) is a sum of both elastic and
liquid like free energies [Flory (1961)]:
F = Fel + Fliq

eq. ( 2-14)

Fliq , is the liquid like free energy that causes compressibility of an ordinary liquid and
is related to change of volume and temperature. It is governed by forces between the
molecules.

Fel , is equal to the sum of the free energies of the network chains and is therefore
related to the network distortion. It does not take into account the interactions between
these chains.
Christensen and Hoeve (1970) concluded that the expression for volume change given
by equation 2-12 was not reliable as the network chains were less restricted than
assumed in theory.
Fedors and Landel (1970) used hydrostatic weighing technique similar to the one used
by Mullins and Tobin (1957) where four ring shaped specimens were stretched under
water using a rig to obtain a direct measure of the volume change taking place in SBR
vulcanisates. SBR was chosen mainly because most of the published data till then that
related volume change to extension was limited to either natural rubber which strain
crystallises or to rubber filled with carbon black which is harder to interpret and make
firm conclusions. Their study focused on effects of water absorption on volume change
90

Chapter 2

Literature Review

and the data is shown in Figure 2-38. While studying the absorption of water by rubber
vulcanisates the specimens were relaxed and reweighed between successive strain
increments. They observed after about 50 minutes that the weight of the immersed
relaxed specimens increased linearly with time as shown in Figure 2-39. After 150
minutes, when the run concluded, the ring specimens were dried by blotting and the
weight of the specimens were measured in air. The difference between the initial and
final weight in air was assumed to be solely due to the water absorbed by the rings.
They found that the volume of water absorbed by the specimens exceeded the relative
volume change expected as a consequence of the hydrostatic component of stress by an
order of magnitude.

Figure 2-38 - Relative volume change versus extension ratio for experiments on unfilled
SBR carried out by Fedors and Landel (1970). The data is represented by filled circles and
the curve is the predicted volume change using equation 2-11.

91

Chapter 2

Literature Review

Figure 2-39 - Results of the study carried out by Fedors and Landel (1970) to study the
dependence of weight of immersed rig + rubber rings set up on time and extension ratio.

Their interpretation of Figure 2-39 was that the increase in weight of relaxed rubber was
a consequence of water entering the rubber and filling up the voids/cavities already
present in them. This however would not increase the volume of rubber thus making the
volume constant and independent of time. Assuming that this is the reason for an
increase in the observed weight of their specimens after blotting dry the surface, they
argued that the deformation of these voids did not contribute significantly to the
volume-extension response.
Several expressions have been proposed in the past that relate volume change to
extension [Fedors and Landel (1970)]. Based on equation 2-12 several one and two
parameter expressions were proposed by researchers that could predict the volumestrain behaviour of elastomers. For instance if the stress-strain response was a linear
elastic response then this would give a one parameter expression for volume-strain
response as:

92

Chapter 2

Literature Review

V E
= ln
V0
3B

eq. ( 2-15)

where E is the Youngs modulus. A simple two parameter equation was derived from
the Mooney equation which for tension is given as:

2C
1
2C10 + 01
2

eq. ( 2-16)

This combined with equation 2-12 gives the volume-extension response of:
V 2C10
=
V0
3B

1 2 2 3C01
3 C
+ 1 + 01

2C10 2 C10
2

eq. ( 2-17)

However, these expressions all assume isotropic behaviour of elastomers upon


extension. Khasanovich (1959) showed that elastomers are not isotropic at large strains
and proposed an additional term that accounted for the strain induced anisotropy. In this
case equation 2-11 becomes:

E 1
V
=
1
V
3B

eq. ( 2-18)

From a phenomenological viewpoint Ogden (1976) described the volume change in


rubber subjected to uniaxial tensile test. He addressed the isochoric and dilatational
aspects of the elastic response of rubber separately and showed mathematically that if
1, 2 and 3 were the principle Cauchy stresses and J was the local ratio of deformed to
ground-state volumes, then the hydrostatic part of stress that was responsible for
causing dilatation was given as:
1
( 1 + 2 + 3 ) = W
3
J

eq. ( 2-19)

Now, In case of simple tension, 2 = 3 =0 and so equation 2-19 reduces to:


1
W
1 =
3
J

eq. ( 2-20)

This agrees with the theory proposed earlier by Gee (1946).


Penn (1970) had also studied the volume changes due to extension of peroxide cured
NR vulcanisates and he found that these volume changes were not proportional to the

93

Chapter 2

Literature Review

isotropic part of the stress. His conclusion was that the strain energy (W) could not be
decoupled and thus expressed it as a sum of both the isochoric part due to shear stresses
and the dilatational part. Ogden upheld this by putting forward the argument that as the
volume changes that occurred due to uniaxial stretching of rubber were of the order of
10-4, a decoupling would imply a volume increase that was linearly proportional to an
increase in the tensile stress. This contradicts all the experimental evidence.
Treloar (1978) maintained that the deviation in experimentally observed volume
changes due to dilatation from the calculated values could only be explained by carrying
out further experiments. This departure led Gee (1987) to develop a systematic
framework to compare various theoretical and experimental strain energy functions. He
concluded that no commonly encountered SEF gave a satisfactory prediction of the
observed volume change.
In case of the materials used in this work, equation 2-18, that was proposed by
Khasanovich (1959), was used to predict the volume change under simple uniaxial
tension due to dilatation effects alone. This equation took into account the anisotropic
behaviour of elastomers at large strains and would be relatively easy to use.
2.10.2 Cavitation Volume change

Fillers such as carbon black when incorporated into rubber matrix are known as
reinforcing fillers as they improve the mechanical properties. One of the most important
factors in determining the extent of reinforcement is the interfacial strength between the
filler particles and the elastomer. Cavities that are oriented in the direction of strain are
formed upon stretching if the rubber matrix pulls away from the filler particles. This
leads to vacuole/cavity formation and this cavitation is directly responsible for an
increase in the sample volume. This obviously negates some of the beneficial
reinforcing effect of the filler.
Investigations in the past have focussed upon cavitation taking place in NR containing
fillers such as lead shot [Schippel (1920)], glass beads [Gent and Park (1984)] and other
filler materials [Jones and Yiengst (1940)] where effects of particle size, filler volume
fraction and degree of vulcanisation were investigated. The relative volume change
under strain was found to be directly proportional to the filler particle size.

94

Chapter 2

Literature Review

Figure 2-40 - Schematic showing formation of hollow spaces formed while deforming a
specimen [Reichert et al. (1987)].

Sekhar and Van Der Hoff (1971) used a dilatometer to measure the volume changes
under strain of cross linked NR filled with glass beads and poly(tetrafluoroethylene)
(Teflon) powder. They found that straining of filled specimens led to opening of
cavities which was followed by diffusion of these gases into the cavities. The greater the
solubility of the gas in the matrix, the higher was the dilatation. Thus determining the
volume of cavities depended to a great extent on the availability of gases that were
dissolved in the elastomer matrix. Teflon was chosen as a particulate filler as it was
known that rubber bonded very poorly with it. They noted that even unfilled compounds
that were mixed mechanically contained intrinsic flaws as described in section 2.9
within them that expanded when the elastomer was stretched, causing dilatation. With

95

Chapter 2

Literature Review

compounds where glass beads were incorporated, cavitation always initiated at the
fillermatrix interface indicative of weak bonding between the two which results in the
formation of vacuum pockets at the poles of the filler under strain. Their experiments
were limited to strains below 100% to avoid complications resulting from strain induced
crystallisation.
Similar findings were reported by Reichert et al. (1987) using a dilatometer when they
incorporated fillers such as glass beads and silicic acid into rubber. Upon strain of these
filled composites, they too observed that the matrix did not adhere to the filler particles
resulting in the formation of hollow spaces. As shown in Figure 2-40, the hollow spaces
would form on both sides in the direction of the deformation. They believed that if this
whole exercise was carried out with these specimens submerged under a liquid, then the
volume dilatation that took place due to this effect could be measured and this could
reflect the interaction between the matrix and filler. Researchers such as Sekhar and
Van Der Hoff (1971), Reichert et al. (1987) have discussed the filler-polymer
interaction and how it changes as the gas that is dissolved in the matrix diffuses into the
resulting hollow spaces.
Gent and Park (1984), in an attempt to identify the precise criteria for failure of bonded
systems, discussed the two distinct modes of failure (Figure 2-41) dominant in filled
elastomers by inserting rigid glass beads into transparent rubber. Previous research by
Gent and Lindley (1958) had shown that elastomers undergo cavitation when they are
subjected to a negative hydrostatic pressure that is approximately given by:
tc = 5E/6

eq. ( 2-21)

where tc is the critical stress for cavitation. This expression was derived by assuming
that all elastomers contain intrinsic flaws which can undergo limitless expansion. When
the elastomer at the surface of this flaw/void reaches its breaking elongation, in order to
relieve the tensile stress at the surface, the void will increase in size until the stress is
reduced enough to stop tearing. For systems with high interfacial strength, cavitation is
the cause of failure.
Systems with a low interfacial strength fail due to the separation of the elastomer from
the surface of the filler which is known as debonding, dewetting or blanching. Oberth
and Bruenner (1965) observed that blanching rarely occurred on both sides of the filler

96

Chapter 2

Literature Review

particles. Composite failure results because of the growth of cavities in the regions of
high stress in the elastomer matrix close to the fillers which then grow around the
interface. Gent and Park (1984) concluded that the critical stress for cavitation, tc, was
dependent on the diameter of the glass bead and the Youngs modulus of the elastomer,
E, and put forth the following expression for the case where the bead is infinitely large:
tc = 5E / 12

eq. ( 2-22)

which is exactly half the value for cavitation of a hollow sphere given by equation 2.21.

Figure 2-41 - Cavitation and debonding at the surface of a spherical inclusion in an elastic
matrix under tension [Gent and Park (1984)].

With debonding, the critical stress, ta, was found to be directly proportional to the
strength of adhesion, P, between the elastomer and the glass bead. However, for an
inclusion of a given size, in contrast to cavitation, it was found to decrease with increase
in E.

97

Chapter 2

Literature Review

This was contradictory to the Griffith treatment of the mechanics of debonding


according to which the debonding stress should increase with E [Gent (1980)]:
t a = (8EP / 3d sin 2 )1 / 2

eq. ( 2-23)

where 2 is the angle subtended by a hypothetical initially debonded circular patch on


the inclusion that is located at the pole and d is the diameter of the inclusion. The
energy required to peel an elastomer from the rigid inclusion, P, is found to directly
depend upon the viscoelastic properties of the elastomer. A harder elastomer would
have a greater value of E and thus would have more intermolecular bonds which would
result in it being less dissipative than a softer elastomer. Hence, the harder elastomer
would show a lower value of P.
The effect of two beads aligned in the direction of applied stress in close proximity was
also studied (Figure 2-42). The specimen failed catastrophically because of the growth
of a second cavity between the two spheres at right angles to the applied stress. They
concluded that it was the second cavitation process that was detrimental to the specimen
as primary cavities always formed at the pole of the inclusion and grew in the direction
of the applied tension but did not result in the final rupture.

Figure 2-42 - Cavitation occurring in a transparent silicone elastomer with two rigid fillers
in close proximity subjected to tensile strain [Gent and Park (1984)].

Ball (1982) studied the radial deformation of a solid sphere composed of an


incompressible, isotropic elastic material and his studies revealed that when the applied
stress on the body reaches a critical value, the body will not deform homogeneously in
98

Chapter 2

Literature Review

pure dilatation. Instead it will take the path that is more favourable energetically which
is to deform inhomogeneously by developing an internal cavity. Stringfellow and
Abeyaratne (1989) used FEA to examine the cavitation that occurs in filled rubber-like
materials. They prepared models of specimens that contained an embedded spherical
particle and stretched them uniaxially to study the hydrostatic stresses that developed
near the poles of the particle. They concluded that the theory proposed by Ball (1982)
was accurate enough to get an estimate of the critical stress level at which cavitation
would occur in rubber-like materials.
In summary, it has been observed in the past that rubbery solids contain cavities of
various sizes which grow into cracks at a hydrostatic stress of approximately 5E/6. This
means that if elastomers could be prepared devoid of any micro-voids/cavities/flaws,
they would be much more resistant to cavitation and thus can be subjected to greater
strains. When fillers such as glass beads are incorporated into rubber, the critical stress
for cavitation is again found to be directly related to the Youngs modulus of the
elastomer. The approach outlined in this thesis is to use the theories that have been
developed for filler-rubber systems and thus draw a parallel to study the cavitiation
phenomenon in granulate filled rubber systems.
2.11 Finite element analysis (FEA)

Finite element analysis (FEA) is an analytical tool that has been used to examine
complex stress analysis problems and it has widespread use in the design of rubber
components. It is a computer simulation technique where the component or the object is
represented by a model with similar geometry. This geometry is then subdivided into a
finite number of small elements that are connected to each other at points known as
nodes. This network of elements and nodes make up a grid that is known as a mesh.
This mesh can be refined by increasing the number of elements and nodes used which in
turn increases the accuracy of the FEA technique. FEA produces approximate solutions
and the effectiveness of the whole approach relies on an appropriate model of structural
geometry being adopted, a high enough mesh density and an appropriate loading
representation.
The equations that define the behaviour of the entire model are obtained from the
equations of the individual elements. These are then put together using matrix algebra

99

Chapter 2

Literature Review

and solved. This is illustrated using the following example of a force displacement
analysis:
[Sq] {} = {F}

eq. ( 2-24)

where, [Sq] is a square matrix, which is also known as the stiffness matrix of the entire
model, {} is the vector of (unknown) nodal displacements and {F} is the vector of the
applied nodal forces.
In case of a one dimensional spring, the force F required to produce a displacement in
a spring with spring constant (stiffness) Sq is given by the above relationship. In order to
compute the displacement for a certain force, the relation can be inverted. In order to
solve structural problems using FEA, this same approach is applied. However, to define
the stiffness matrix fully before inverting the equation, certain boundary conditions
have to be applied to the model. So for example in the case of a stress analysis problem
to inhibit unrestricted rigid body motion, the body must be constrained, or for a problem
involving thermal analysis, the initial temperature must be defined at various nodes in
the model.
Finite Element Method (FEM) helps the analyst solve a large range of problems that
were previously considered complex and obtain the solutions using clear and
identifiable steps. It generally involves 3 phases that can be listed as:
1) Preprocessinginvolving the preparation of the model and defining the
mesh.
2) Analysisinvolving the computation of the solution.
3) Postprocessinginvolving the viewing and analysis of models results.
Nowadays the pre-processors and post-processors have become so advanced that it
gives an appearance of unquestionable accuracy to an inexperienced user. Highly
complex models can now be created with little input from the engineer using the preprocessor. The post-processors also provide very complex design tools.
Ultimately in order for FEA to be used as an effective tool to aid analysis, it is
imperative that the problem must be represented in a reliable manner followed by the
implementation of the correct analysis procedure [Fagan (1992)].

100

Chapter 2

Literature Review

2.11.1 FEA models

Three dimensional (3D) FEA models of real engineering structures can be very time
consuming to prepare and solve. The geometry can be simplified using symmetry
approximations. Two dimensional (2D) models should be used for analysis if the
geometry and loading conditions can be depicted in a single plane.
In this thesis both 2D and 3D models are used. 2D systems can be modelled using either
the plane strain or the plane stress conditions. Plane stress condition can be used to
simplify and analyse the model if the body has an insignificant dimension in one
direction and is loaded in the plane of the body (such as the tensile test piece). The
plane stress condition has been used in chapter 4 in order to simplify the analysis of a

specimen with an edge cut. If the body is long, has a geometry and loading conditions
that do not vary in its longitudinal direction, then the system can be modelled as a 2D
system using plane strain. Axial symmetry is effectively plane strain in a cylindrical
coordinate frame. It assumes uniform geometry and constant variable distribution
around the axis of symmetry. Only radial and axial deformations are possible. An
example of this is illustrated in Figure 2-43.
However, if the plane stress, plane strain and axial symmetry conditions are not
satisfied, then a full 3D model is sometimes necessary. In chapter 4 a simplified 3D
model of a corner cut has been discussed where symmetry assumptions have been made
which helps reduce the analysis time. For the 3D model discussed in chapter 4, only the
repeated part of the structure has been modelled. This can be done provided that the
symmetry of the problem is correctly applied and can accurately represent the loading
and constraints across the appropriate parts of the model [Busfield (2000)]. Common
symmetry conditions used include:
Planar symmetry which can best be explained with the example of a flat plate with a

hole loaded uniformly. In such a case only a quarter of the model has to be used to
represent the whole body. The constraint conditions must be applied correctly at the
symmetry edges for the model (Figure 2-43).
Cyclic symmetry is similar to planar symmetry but the symmetry is described in a

cylindrical coordinate system rather than a rectangular one. Once again the constraint
conditions must be applied correctly (Figure 2-43).

101

Chapter 2

Literature Review

Repetitive symmetry is shown in Figure 2-43. When a certain part of a structure is

found to repeat itself at regular intervals, in a linear sequence, repetitive symmetry can
be used.

Figure 2-43 Schematic from Busfield (2000) showing the four types of symmetry that are
commonly adopted to simplify the FE modelling of a component.

2.11.2 Types of elements

The pre-processor used throughout was I-DEAS v.11 supplied by SDRC in order to
prepare and mesh the model. The analysis and visual examination of the model was
done using the post processor, Abaqus v 6.6 software supplied by Hibbitt, Karlsson and
102

Chapter 2

Literature Review

Sorensen, Inc. Four node rectangular elements were chosen for the mesh. Hybrid
elements were chosen to prevent the over constraint resulting from the use of a virtually
incompressible material model and so that when the force is applied to the model, the
model would deform in the appropriate manner and not remain locked in its original
shape. For plane stress elements, this incompressibility is not a problem and so normal
elements are used.
2.11.3 Analysis procedure

Busfield (2000) has detailed the following five stages that are typically involved in
developing a finite element model:
1. Selection of a suitable analytical model that can be used to accurately represent
component behaviour. The important factors that must be considered while
doing this exercise are expected behaviour when the component is loaded,
typical applied loads and material properties.
2. Meshing is the next step wherein the model is divided into a network of
elements and nodes. Materials, like rubber that can undergo large strains,
occasionally become too highly distorted and the elements assume unacceptable
shapes. This results in an error and as a result the analysis ends before a solution
is achieved.
3. Boundary conditions are defined along with the material properties of all the
elements. The pre-processor then generates an input deck which basically has all
the information required to run the analysis.
4. This input deck is submitted to a computer that has a dedicated equation solver.
Taking into account the boundary conditions and material properties, the
equation solver works out a numerical solution for the mesh. This can often
involve iterative steps.
5. The analysis is followed by the post processing stage where parameters such as
the nodal coordinates, the stress contours and strain energy contours are
examined in detail.
An input deck is generated by the pre-processor and basically describes the problem to
be analysed. It comprises of:

103

Chapter 2

Literature Review

1. Header Statement that defines the problem, title section and format of the
output files.
2. Geometry Defines the spatial arrangement of the material, using elements and
coordinates of nodes.
3. Material definition Ascribe the model as 2D or 3D and the material type. In
case of elastomers a strain energy function (SEF) is used to define the
mechanical behaviour and so a hyperelastic material definition is used.
4. Boundary conditions Symmetry constraints are inserted by specifying the
boundary conditions. Sometimes nodes need to be tied together and in such
cases multi-point constraints (MPCs) are used.
5. Step definition Analysis type and actual loading conditions are specified here.
2.11.4 Strain energy density

At large strains the mechanical behaviour of elastomers is non linear. This is


compounded by the fact that geometric non linearity makes modelling of rubber
materials at large strains a challenging task. A strain energy function (SEF) is used to
describe the amount of energy that is stored elastically per unit volume of the material at
a particular strain. In order to deal with large deformations, one approach is to select an
appropriate mathematical form for the SEF that can fit the measured experimental stress
strain data of the material under consideration. In order to obtain useful and accurate
solutions using numerical techniques such as FEA it is essential that the correct SEF is
used to describe the elastic behaviour of the elastomer.
This approach was initiated by Mooney (1940) and was then further developed by
Rivlin (1956). Rivlins large strain elasticity theory made the fundamental assumption
that rubber is isotropic. This implied that the elastic strain energy per unit volume, W,
would be symmetrical with respect to the three principal extension ratios, 1,2 and 3.
Three strain invariants I1, I2 and I3 that were introduced by Rivlin that could be used to
express the SEF were defined as:
I 1 = 1 + 2 + 3
2

eq. ( 2-25)

I 2 = 1 2 + 2 3 + 3 1
2

104

Chapter 2

Literature Review

I 3 = 1 2 3
2

Mooney (1940) used mathematical reasoning and derived the following SEF:

1
1
1
2
2
2
W = C10 1 + 2 + 3 3) +C01 2 + 2 + 2 3
2
3
1

eq. ( 2-26)

Expressed in terms of the strain invariant and assuming incompressibility this can be
rewritten as the Mooney SEF:
W = C10 ( I 1 3) + C01 ( I 2 3)

eq. ( 2-27)

Rivlin (1956) showed that if t1, t2 and t3 are the true principal stresses relating to
deformed dimensions in the respective directions, then for a pure homogeneous strain:
t1 t 2

1 2
2

t2 t3

2 3
2

t1 t 3

1 3
2

W
2 W
= 2
+ 3
I 2
I 1

eq. ( 2-28)

W
2 W

= 2
+ 1
I 2
I 1

eq. ( 2-29)

W
2 W

= 2
+ 2
I 2
I 1

eq. ( 2-30)

The relation between true stress, ti, and engineering stress, i, is given by:

i =

ti

eq. ( 2-31)

In case of uniaxial deformation, the reduced stress, *, is given by:

* =

W 1 W

= 2
2
( )
I 1 I 2

eq. ( 2-32)

Rivlin (1956) worked on large elastic deformations and proposed a general SEF using
mathematical reasoning where the strain energy density, W, was expressed as a function
of the three strain invariants as:
W =

C (I

ijk = 0

ijk

3) (I 2 3) (I 3 1)
i

eq. ( 2-33)

105

Chapter 2

Literature Review

In case of an incompressible material, I3=1 then equation 2-33 may be reduced to the
following power series:

W = C ij (I 1 3) (I 2 3)
i

eq. ( 2-34)

ij = 0

In an unstrained state, the strain energy is zero and hence C00 is set equal to zero to
reflect the same. The first two terms of this power series (i=1, j=0 and i=0, j=1) yields
equation 2-28 which is the Mooney SEF. In case of a Mooney material, the partial
derivatives W/I1 and W/I2 equal C10 and C01 respectively.
The Mooney SEF ignores the upturn present for all elastomers at large strains due to the
finite extensibility of the network. This SEF is considered adequate to model unfilled
rubbers because the upturn occurs at large strains.
Gregory (1979) investigated the effects of carbon black fillers on stress strain properties
of rubber vulcanisates and found that a simple relation existed between the reduced
stress and (I1-3) only if W/I2 was much smaller than W/I1 and if W/I1 was
independent of I2.
James and Green (1975) showed for filled elastomers that W/I2 was much smaller
than W/I1. The data published by Fukahori and Seki (1992) also showed that W/I1
is much greater than W/I2 (numerically close to zero) for filled rubber. Researchers
such as Gough et al. (1998) and Yeoh and Fleming (1997) have shown that in case of
carbon black filled rubber both the above conditions are met to a satisfactory degree and
hence agreed with Gregorys findings. However, Obata et al. (1970) found that for
unfilled rubber both W/I1 and W/I2 varied with I1 and I2 respectively.
Based on these previous findings, Yeoh (1990) assumed for carbon black filled
materials that W/I2 is approximately equal to zero and derived the following cubic
equation:
W = C10 ( I 1 3) + C 20 ( I 1 3) 2 + C 30 ( I 1 3) 3

eq. ( 2-35)

This SEF contains only I1 terms making it easier to calculate the coefficients from
simple tests. The Yeoh SEF has been used successfully in the past by several
researchers including Busfield et al. (2005) to predict the stress strain behaviour of

106

Chapter 2

Literature Review

filled elastomers at large strains. Hence, the SEF proposed by Yeoh has been used in
this thesis to model the behaviour of filled rubber.
The coefficient values are obtained experimentally and then entered in to the input deck
of the FEA software package.
2.12 Summary and aim of this study

It is widely accepted that a continued research on internal fractures of rubber solids


would shed better light on understanding their strength. This review has identified the
importance of rubber recycling and looked at the various rubber recycling methods that
have been tried and tested. One such method is recycling tyres by grinding the tyre
rubber into fine granulates which can then be reincorporated into virgin rubber
compounds. Even though this route of recycling offers a lot of potential and can tackle
large volumes of waste rubber, it has not been of much commercial significance yet and
the reason for this is the apparent reduction in performance of rubber compounds when
scrap rubber is added into it. With current EU legislation, existing tyre mountains need
to be handled and tyre and rubber companies are looking for ways to tackle their ever
growing volumes of rubber waste. Hence this is a crucial time to examine the reason for
this decline in performance and identify solutions for improving the virgin rubber-scrap
rubber mix.
Initial attempts were made to identify and characterise the rubber granulates. Various
techniques were used to obtain information such as density, particle size, composition
of granulates. These techniques were used on granulates obtained from commercial
sources of unknown origin and composition as well as on our own materials prepared
carefully with a known composition. The study was carried out in three main phases:
Phase I A technique was developed to examine the change in flaw size when rubber

granulates are incorporated into virgin rubber. Flaws that are inherent in the matrix
rubber act as stress raisers and it was necessary to check if addition of rubber granulates
increased the flaw size thus causing it to fail prematurely.
Phase II Poor bonding between the matrix and granulate can lead to a weak interface

which could then be a site for failure initiation. A suitable technique was developed to
study this interfacial strength.

107

Chapter 2

Literature Review

Phase III Rubber materials undergo dilatation when stretched due to the hydrostatic

stresses encountered. But when granulates are incorporated, is there an additional


volume change that takes place due to cavitation? Once a cavity opens up as a flaw, at
what strain does it result in a catastrophic failure? An approach to examine these issues
was identified.
Throughout the work FEA analysis was adopted to corroborate the experimental
findings.

108

Chapter 3

3.
3.1

Experimental Techniques & Materials

Experimental Techniques & Materials

Introduction

This thesis describes a study of the reduction in strength of rubber materials resulting
from the incorporation of ground rubber granulates. A systematic approach has been
developed to investigate what happens to the size of the intrinsic flaws in the rubber
matrix when granulates are incorporated. A fracture mechanics approach has been used
to investigate failure initiation due to flaws located primarily at the matrix-granulate
interface. FEA techniques were used to ascertain debonding at the interface and to
investigate the mechanical behaviour of elastomer blends.
The various different compounding recipes for the materials used in this work are
explained in sections 3.2 and 3.3. Initially, tensile tests were carried out to characterise
the mechanical properties of the material using the strain energy functions discussed in
chapter 2 and to determine the tensile strengths of various matrix-granulate
combinations. This is followed by a detailed description of tests undertaken to
determine the change in the intrinsic flaw size (IFS) of the rubber materials resulting
from the incorporation of granulates. In the past, one of the methods by which IFS has
been estimated is by the extrapolation of fatigue lives of test pieces with cuts of various
sizes in them to fatigue lives with no cut. This method is time consuming as the fatigue
life data of a wide range of specimens has to be measured. The method adopted here in
this work and which has been used previously by Gent (1978) and Hamed (1983) adopts
a strength measurement approach and is described in section 3.4.
After studying flaw/crack initiation and flaw size estimation, a technique was developed
to examine the interface between the matrix and granulates and between the granulates.
The technique developed here is described in sections 3.5 and 3.6, such that the
interfacial strength could be quantified. By merging the experimental data with finite
element analysis (FEA) findings the tear path as the flaw grows through the matrixgranulate composite is investigated.
Section 3.7 describes the experiments that were undertaken to check the extent of
cavitation taking place in the composite when strained using careful volume change
experiments. FEA (described in section 3.8) was again used to study the debonding
109

Chapter 3

Experimental Techniques & Materials

along the matrix-granulate interface and to predict the extent of cavitation as the
composite material is strained.
As the focus of the whole study was on matrix-granulate composite materials, it was
obvious that it would be difficult to make a conclusive study on a rubber-granulate
blend if the composition of the granulate that is being mixed into the rubber matrix is
unknown. To be able to make predictions of the crack growth behaviour, materials
identification and characterisation techniques were used to determine the composition of
the rubber granulates. This is described in section 3.9.

Figure 3-1 Granulates of 30 mesh size used in this study (a) SBR 0 (b) SBR 70 and
(c) magnified single ambiently ground SBR 70 granulate.

3.2

Materials used in this study

Styrene Butadiene Rubber (SBR) was the base polymer used throughout this study. This
is a synthetic copolymer consisting of styrene and butadiene. The chemical structure of
SBR is described in chapter 2 (section 2.1). It does not crystallise when strained and has
a glass transition temperature, Tg close to -55C.

110

Chapter 3

Experimental Techniques & Materials

Table 3-1 - Formulation for the materials used in this study.

SBR,
70 phr
Carbon black

SBR,
No
Carbon black

100

100

70

Stearic Acid

Zinc Oxide

Ingredients &
Curing conditions

Type

SBR 1500
Carbon Black

N330 (HAF)

Antioxidant

HPPD*

Accelerator

DPG

1.3

1.3

Accelerator

MBTS

Sulphur

1.5

1.5

Moulding Temp, C

160C

160C

Time, min

15

15

HPPD - N-phenyl-N-(1,3-dimethylbutyl)-p-phenylenediamine

Two sets of granulates, shown in Figure 3-1, were ambiently ground by J. Allcocks and
Sons Ltd. from carefully moulded SBR sheets which contained either 70 parts per
hundred (phr) of carbon black (SBR 70) or no carbon black (SBR 0). The grinding
operations reduced the vulcanised sheets into granulates. The size of the resulting

111

Chapter 3

Experimental Techniques & Materials

granulates was estimated by sieving through a series of standard screens using a sievetray shaker. A 30 mesh screen would have 30 holes or opening per linear inch of the
screen. Rubber granulates used in this study passed through a 20 mesh screen (~850
m) but not a 30 mesh screen (~650 m). This broadly determines the size range of the
granulate used in this study. The detailed formulation of these materials is shown in
Table 3-1. Varying amounts of granulates were incorporated into new virgin material
(matrix) in order to study their effect.
3.3

Vulcanisation

Figure 3-2 Digital photograph showing the distribution of 30 mesh SBR 70 granulates in
synthetic isoprene rubber. Granulates were mixed using the two roll mill.

Rubber material preparation requires two stages:


1. The first is the preparation of the unvulcanised masterbatch using an internal
mixer where the unfilled SBR, carbon black if required and other ingredients
such as activators and antioxidants were added in the first two minutes. This mix
was then dumped after 4 minutes at about 180C.
2. The second involves incorporating the vulcanising ingredients which were
sulphur and accelerators into the masterbatch using a two roll mill at a lower

112

Chapter 3

Experimental Techniques & Materials

temperature. A combination of primary and secondary accelerators was added


into the rubber in order to impart better storage stability and ageing
characteristics. When required granulates were added during this stage as well.
It was very important that there was a thorough distribution of carbon black in the SBR.
Hence the first stage was carried out at a high temperature as this lowers the rubbers
viscosity and facilitates good mixing. The second stage was carried out at a lower
temperature so that no curing occurred prior to moulding. For the case of the filled SBR,
the carbon black used was high abrasion furnace (HAF N330). A combination of
primary accelerator (MBTS - 2-benzothiazolyl disulfide) and secondary accelerator
(DPG Diphenyl guanidine) was used to adjust the cure characteristics. MBTS gives a
moderately fast cure for sulphur cured synthetic elastomers and the secondary
accelerator improves the vulcanization rate and crosslink density of the primary
accelerator.
Table 3-2 - Combination of matrix and granulates used for the IFS estimation.

Nomenclature

Matrix (M)

Granulates (G)

Volume fraction (X)

M0-G0(X)

SBR 0

SBR 0

21% or 47%

M0-G70(X)

SBR 0

SBR 70

17% or 41%

M70-G70(X)

SBR 70

SBR 70

14% or 35%

M70-G0(X)

SBR 70

SBR 0

17% or 41%

M0

SBR 0

M70

SBR 70

Conditions for the cross linking were maintained throughout the study. After the second
stage mixing in the two roll mill the rubbers were compression moulded into 2 mm
113

Chapter 3

Experimental Techniques & Materials

thick sheets at 160C for 15 minutes. The mould plates had surfaces that were polished
and had mirror like finish so that no surface flaws were introduced. The moulded rubber
sheets were carefully removed from the moulds to avoid pre-straining.

Figure 3-3 Instron 5567 with video extensometer setup that was used to carry out tensile
tests.

Some of the test pieces (approximately 25-30) were tested with corner cuts of length c
that were introduced using a clean and sharp razor. Figure 3-3 shows the Instron 5567
equipped with a non-contacting video extensometer that was used to carry out the
tensile tests. As shown in Figure 3-3, the video extensometer recorded the distance
between optical dots adhered to the rubber surface. The initial cut length, c, was
measured by examining the fracture surfaces (Figure 3-4) after break using an optical
microscope. The cut length c was the perpendicular distance between the corner and the
razor cut edge. There was always a clear demarcation between the rough fracture

114

Chapter 3

Experimental Techniques & Materials

surface and the smooth razor cut surface. IFS values were estimated using log b versus
log c plots.

100 m

c = 332.5 m

Figure 3-4 - Microscopic image of a fracture surface showing the original cut length.

3.4

Tear strength

In order to estimate the tear strength of the materials with same composition as both the
matrix and the granulates, trouser tear testing was done. To investigate the tear strength
of the granulates, a sheet with the same composition as both types was prepared and
cured at 160C twice. Each cure cycle being 15 minutes so that, like the granulates it
would be subjected to a double curing cycle. Another sheet representing the matrix was
mixed and cured once at 160C for 15 minutes. Rectangular strips with dimensions of
100 x 20 x 2 mm were cut from these sheets and along the centre line of these strips a
cut of length 30 mm was introduced to make a trouser test piece as shown in Figure 3-5.
The two legs were pulled apart at a constant strain rate of 0.77 min-1. and the tearing

115

Chapter 3

Experimental Techniques & Materials

energy, T was calculated from the average force F required to propagate the tear using
the relation derived by Rivlin and Thomas (1953):

T=

2 F
wW
t

eq. (2-2)

where w is the width of each leg, ~10 mm, t is the thickness of each leg, ~2mm and W is
the elastic strain energy density in the legs.

Figure 3-5 - (a) Schematic showing a trouser test specimen with a cut along the centre line
to allow the two legs to be gripped (b) Experimental setup showing a trouser test
specimen.

3.5

Peel strength

In order to quantify the strength of the interface between the matrix and granulate, peel
tests were done. The sample preparation involved 2 stages and is illustrated in Figure
3-6. A sheet with the same composition as granulates was mixed and compression
moulded at 160C for 15 minutes (stage 1 in Figure 3-6). After a period of at least three
days, another sheet with the same composition as the matrix was mixed using the two
116

Chapter 3

Experimental Techniques & Materials

roll mill but not cured. These two sheets were then compressed together but separated at
one end by a thin strip of Polytetrafluoroethylene (PTFE) coated tape. The tape was
used to separate a third of each of the two sheets to allow the two free ends to be
gripped during testing. This combination was compression moulded with the same
curing conditions (160C, 15 min). A flat interface was thus created between the two
sheets as shown in Figure 3-6 (stage 2). The sheet representing granulates was double
cured (A in Figure 3-6) and the sheet (B in Figure 3-6) representing the matrix was
cured only once. In order to quantify the interfacial strength, five separate strips were
cut out and as shown in Figure 3-7, the gripped free ends were pulled apart at a constant
rate to separate the strips across the interface.

Figure 3-6 - Schematic diagrams of peel test between sheets having same composition as
the matrix and granulates.

Quantification of interfacial strength between matrix and granulates in this manner has
been done in the past by Burford and Pittolo (1984). The peel energy, P, was calculated
using:

117

Chapter 3

P=

Experimental Techniques & Materials

2 F
tW
w

eq. ( 2-3)

where F is the average peel force, w denotes the width of the peel strips ~30mm, t is the
thickness of each sheet ~2 mm and W is the elastic strain energy density in the stretched
strips.

Figure 3-7 - (a) Schematic of a peel test specimen with force applied across the interface
(b) Experimental setup showing a peel test specimen.

In order to obtain a measure of the interface between two granulates, a similar


procedure (as described above) was adopted to prepare the specimens and is shown in
Figure 3-8. The two sheets (A and B) were separately cured so that they would have
undergone a single cure cycle and then after 3 days those sheets were cured together
again for 160C for 15 minutes with the thin PTFE layer in between. This resulted in a
flat interface between two double cured sheets. Once again in order to obtain the

118

Chapter 3

Experimental Techniques & Materials

interfacial strength between granulates, five strips were cut and each of these were
gripped and pulled apart a constant rate to separate the strips across the interface.

Figure 3-8 - Schematic diagrams of peel test between sheets having same composition as
the granulates.

3.6

Volume change measurement

The technique used to measure volume change under strain was hydrostatic weighing.
The apparatus used (Figure 3-9) was similar to that used by Mullins and Tobin (1957)
that had been further developed from the one developed originally by Gee et al. (1950).
Due to the relative small changes in volume of the order of 10-4, this method gives a
direct measure of the volume change much more accurately than any technique that
relies on accurately measuring the external dimensions alone. Similar to earlier workers,
[Gee et al. (1950) and Mullins and Tobin (1957)] four elastomer ring specimens were
placed on the rig. These rings were cut from moulded sheets, about 2mm thick. The
rings had an internal diameter of about 14 mm and an outer diameter of about 20 mm.

119

Chapter 3

Experimental Techniques & Materials

The whole experiment was carried out with the rig immersed in distilled water to which
a little detergent was added in order to facilitate the complete wetting out of the rig and
sample surfaces and to reduce the risk of air bubbles being trapped. The rig was
suspended by a platinum wire which was a part of the MettlerToledo single-arm
analytical balance used for this study. In order to maintain the accuracy and consistency
of the readings, this balance was mounted on a vibration free table and the water
temperature was maintained at 232C. The temperature was monitored at the time of
each reading using a thermometer that was kept immersed in the water.

Figure 3-9 Stretching rig used for the volume change experiments.

The first mass readings were recorded with the rings in the unstretched position. This
was followed by subsequent readings that were made after the rings had been stretched
to the desired extension. A third reading was then taken with the rings relaxed. This
cycle was repeated 3 times at each extension. Thus an average of three readings was
noted in the case of the stretched and unstretched states at each extension. The change in
volume at a particular extension was then calculated from the difference in these mean
weights between the unstretched and stretched states. Similar to Gee et al. (1950), the
extension ratios ( = l/l0) were calculated on the internal diameter of the rings. In each

120

Chapter 3

Experimental Techniques & Materials

case, 15 minutes were allowed for the system to attain equilibrium before each reading
was noted.
Before commencing the experiment, these specimens were conditioned by immersion in
distilled water for 3 days so that during the experiment, there would be no change in
weight resulting from even very modest amounts of water absorption. Once the
experiment was started with the rig + rings immersed in water, the fixtures were kept
immersed until all measurements were completed.

Figure 3-10 A schematic of the Mettler-Toledo balance that was used to measure the
volume change of the rings that were suspended by means of the platinum wire shown
above.

3.7

Finite Element Analysis (FEA)

The corner cuts that were introduced in the specimens tested in this study are different
to edge cuts. The different cut geometry creates a different strain energy release rate
versus crack length relationship. The actual strain energy release rate for this geometry
was investigated here using FEA.
Busfield (2000) has discussed the selection of an appropriate element type and the
correct representation of test piece geometry to make a reasonable analysis using FEA.
Three different techniques namely: energy balance, J-integral and crack tip closure were
used by Busfield et al. (1996) to calculate the strain energy release rate. They showed

121

Chapter 3

Experimental Techniques & Materials

that even in case of complex 3D models that the calculated strain energy release rate
was independent of the specific technique used. It was therefore established that for a
specific problem the technique chosen should be the one that is the easiest to
implement.
For the purpose of this work, the energy balance approach was used. This involves
measuring the drop in the total strain energy stored in the specimen (dU) between two
models held at a fixed extension where the crack tip is extended by a small area (dA).
The tearing energy is then the ratio of dU over dA. Thus each model with a pre-cut of a
known size was strained to the same extent and T values were obtained. In order to
represent a model with a pre-cut of known length, nodes in the model are released and
so this procedure is also known as the node release or virtual crack extension technique
(VCE).
Finite element analysis was also used to study the peeling across the matrix-granulate
interface. The strain energy release rate with increasing crack length along the interface
was compared with tearing across the granulate in order to deduce the likely path for
crack propagation for different combinations of matrix and granulate.
Finally FEA was used to predict failure at the matrixgranulate interface under strain.
Experimental test data was used in combination with FEA to investigate cavitation at
different strains.
3.7.1 Pre-processing

The pre-processing was done using SDRCs IDEAS software v.11.


Edge cut model

Figure 3-11 is a 2D plane stress model of a tensile test piece with an edge cut that was
prepared and meshed using IDEAS v.11. The model was used to investigate the change
in strain energy release rate, T, with the increase in crack length, c. The elements used in
this model consisted of plane stress four node elements (CPS4). The element sizes
varied along the length and width of the model with a higher mesh density in the
vicinity of the crack initiation site as that was the region with the largest strain
gradients. Symmetry conditions existed in the plane of the crack and so as explained in

122

Chapter 3

Experimental Techniques & Materials

chapter 2, in order to simplify the analysis only the upper half of the test piece was
modelled.

Figure 3-11 Half symmetry model of a tensile test specimen with an edge cut.
(a) undeformed model (b) deformed model showing the region with fine mesh containing
the cut. The uncracked part of the bottom surface is constrained in the 2 direction.

Corner cut model

To study the T versus c relationship of a tensile test piece with a corner cut, a 3D model
was used to represent the geometry. The elements used were solid (continuum) three-

123

Chapter 3

Experimental Techniques & Materials

dimensional eight-node hybrid brick elements (C3D8H). Symmetry constraints were


again used to represent the geometry with a half symmetry model (Figure 3-12).

2
3
Left Surface
Front Surface

Crack
propagation

Bottom
Surface

Figure 3-12 Deformed half 3D model of a tensile test specimen with a corner cut. Nodes
at the top are pulled in the 2 direction. The uncut bottom surface is constrained in the 2
direction.

Matrix-Granulate model

For simplicity, it is assumed that the granulate is spherical in shape and evenly
dispersed throughout the matrix. Figure 3-13 shows a 2D plane stress model of a rubber
matrix with a circular granulate embedded in it. The volume fraction of granulate was
varied in accordance with the matrix-granulate composite materials that were used for
peeling and tearing experiments (listed in Table 3-2). The elements used in this model
also consisted of plane stress four node elements (CPS4). Element biasing was done
which meant that a sufficiently fine grid (fine elements) was allocated in the vicinity of
the crack propagation regions as these underwent the highest strain gradients in the

124

Chapter 3

Experimental Techniques & Materials

model. These high strain gradients usually occurred at discontinuities such as interfaces
or at sharp corners.
Matrix

Granulate
1

Figure 3-13 2D representation of the matrix-granulate model showing a circular


granulate embedded in the matrix.

As mentioned earlier, typical granulates used in the experimental part of this study were
in the range of 600 m to 850 m. The granulates falling in this range are assigned a 30
mesh size label. Tests conducted at J. Allcocks & Sons Ltd. and the particle size
analysis at QMUL confirmed that 30 mesh was approximately 650 m. Therefore, for
all the FEA models, the circular granulate diameter was fixed at 650 microns.

125

Chapter 3

Experimental Techniques & Materials

Figure 3-14 Schematic of the matrix-granulate model indicating the two paths along
which nodes were released to simulate crack growth. Bottom surface was fixed in both the
1 and 2 directions whereas the top surface was constrained only in the 1 direction and was
moved in the two direction.

(a)

(b)

Figure 3-15 Deformed models of M70-G70(35%) showing the (a) tear propagation
through the granulate (b) tear profile under strain and its final path.

126

Chapter 3

Experimental Techniques & Materials

(a)

(b)

Figure 3-16 Deformed models of M70-G70(35%) showing how rubber might peel around
the interface.

A schematic of this model (Figure 3-14) shows two different paths which the crack
might follow. Crack propagation along path 1 simulated tearing through the granulate
and path 2 was representative of peeling across the matrix-granulate interface. Crack
propagation was achieved by releasing nodes along the crack profile as shown in Figure
3-15 and Figure 3-16. For each path, the strain energy release rate was calculated as a
function of crack length and strain using the energy balance approach.
Cavitation model

The model used to study the debonding at the matrix-granulate interface was an
axisymmetric model whose elements were solid (continuum) axisymmetric four node
elements (CAX4H). Different axisymmetric granulate models were prepared for each
different volume fractions of material used in the volume change experiments. Figure
3-17b shows a typical matrix-granulate half symmetry axisymmetric model. Figure
3-17c shows a typical model subjected to a tensile strain of 50%.

127

Chapter 3

Experimental Techniques & Materials

Figure 3-17 The axisymmetric FE model used for studying the debonding (a) Granulate
G is embedded in the Matrix M and the debonding will take place along the dotted line
which represents the interface between the M-G. (b) 2D representation of an undeformed
model with 30 phr granulate incorporated by weight. The arrow indicates the path along
which the flaw size is increased during the analysis. (c) 2D representation of model
strained to 50%. The arrow shows the direction where the interface is debonding.

128

Chapter 3

Experimental Techniques & Materials

The debonding along the matrix-granulate interface was also achieved using a node
release technique. The model was axisymmetric, with the flaw increasing at the pole of
the granulate in the model as shown in Figure 3-17a. Here the flaw opens up as the
vacuole formed at the pole. For each node, n, that was released; the flaw size was
recorded on the undeformed geometry, the actual debond length (Db) being twice the arc
length shown in Figure 3-17a as a result of symmetry.
For each flaw size, the model was stretched under tension to a range of strains. The
volume of the cavity at various extension ratios was calculated using numerical
techniques.
3.7.2 Material characterisation

Section 2.11.4 describes how a strain energy function (SEF) is used to characterise the
elastic behaviour of the rubber material. A mathematical form is used to fit the
measured experimental stress strain data of the material under consideration, which is
then used in the FEA software to predict the behaviour of the elastomer at large strains.
It is, therefore, crucial to select an appropriate SEF to analyse and evaluate the
elastomer behaviour at large deformations. ABAQUS contains different types of
hyperelastic materials each of which can be described by different functions. For the
purpose of this work two different SEFs were used. The Mooney SEF described in
chapter 2 was used to characterise the mechanical behaviour of the unfilled rubber and
is repeated here for convenience:
W = C10 ( I 1 3) + C01 ( I 2 3)

eq. ( 2-27)

C10 and C01 are material coefficients that were used to characterise the material.
The Yeoh SEF as explained in chapter 2 was used to describe the behaviour of filled
rubber and is also repeated here for convenience:

W = C10 ( I 1 3) + C 20 ( I 1 3) 2 + C 30 ( I 1 3) 3

eq. ( 2-35)

This equation when differentiated with respect to I1 can be rewritten for the case of
uniaxial tension as:

129

Chapter 3

Experimental Techniques & Materials

2
= 2C10 + 4C 20 (I 1 3) + 6C 30 I 1 3)
2

eq. ( 3-1)

Here C10, C20 and C30 are the constants measured from tests on material behaviour at
specific temperatures and strain rates.
The coefficients for the Yeoh and Mooney SEFs were determined by plotting the stressextension ratio data for the materials in consideration. The uniaxial tensile tests carried
out are described in section 3.4.
3.7.3 Boundary conditions

Boundary conditions are applied to the model before submitting it to the server for
analysis. These conditions are crucial and must be applied carefully in order to ensure
accuracy and to prevent the analysis from exiting with an error. A fixed constraint is
the most common boundary condition applied for FE analysis and involves attaching a
surface (or node) of the model firmly to the ground or another component thus
rendering it immovable in any direction. This is done by setting the displacements of all
the degrees of freedoms (DOFs) associated with that particular surface (or node) to zero
in the input deck.
For instance if the bottom surface of a component is resting on the ground then the
displacement of the DOF normal to this surface is set to zero. Hence the component is
free to move parallel to the surface. This is equivalent to setting the nodes on the bottom
surface on rollers. For the models that were used in this work, the boundary conditions
applied were as follows:
Edge cut model

As seen in Figure 3-11, the bottom surface symmetry plane was constrained in the
vertical direction (the 2 direction) whilst each node on the top surface was displaced
vertically by an identical amount. Nodes were then released along the crack profile to
represent an actual edge cut.
Corner cut model

The corner cut half symmetry model is shown in Figure 3-12. Only half the total height
was modelled due to symmetry with the bottom surface again being the symmetry
plane. The nodes along the front surface were all constrained in the 3 direction whereas

130

Chapter 3

Experimental Techniques & Materials

the uncut bottom face was constrained in the 2 direction only. A node release approach
was used to vary the corner cut length for a range of strains. Using the energy balance
approach, for every c, the strain energy release rate, T, was readily calculated.
Matrix-granulate interface model

For the matrix-granulate interface model, virtual crack extension (VCE) technique was
used to release nodes along the two different paths, as shown in Figures 3-13, 3-14 and
3-15 in order to determine the path that would have the highest strain energy release
rate. The bottom surface nodes were again all constrained in all directions and so it was
fixed to the ground whereas the top surface was displaced vertically to various different
strains but constrained in the 1 direction.
Cavitation model

This axisymmetric model used is shown in Figure 3-17. The nodes were released
beginning at the pole (left edge) along the matrix-granulate interface and the top surface
was displaced vertically to represent debonding along the interface. The volume of the
cavity that opened up was computed using numerical techniques and thus the change in
volume was computed. As the bottom edge is the plane of symmetry, these crack
lengths and volume change values were doubled to represent the crack opening at the
top and bottom of the granulate. The bottom and the left edge were constrained in the 2
and 1 directions respectively.
3.7.4 Analysis / Solution

The input deck was submitted to an equation solver (ABAQUS) which uses an iterative
approach which seeks to end the analysis either when a solution is obtained or when
some aspect of the model becomes more unstable. As the solution time approximately
squares with the number of degrees of freedom, it is important to use the optimum
number of elements in a model. Less number of elements will give an inaccurate
analysis whereas too many elements would take too long for a sensible solution to be
calculated.

131

Chapter 3

Experimental Techniques & Materials

3.7.5 Post processing

Once a successful analysis is carried out by the solver, the results can be viewed in
ABAQUS. Information such as total elastic strain energy for the whole model for
various crack lengths and nodal coordinates at various displacements are obtained in
this step.
3.8

Granulate characterisation

A piece of rubber for tyres is usually a mixture of two or more polymers, carbon black,
inorganic compounds such as calcium carbonate, silica and a relatively high number of
organic substances such as plasticizers and lubricants. In addition, other materials like
antioxidants and vulcanisation agents are also dissolved in the polymer matrix. In order
to incorporate tyre granulates of unknown composition into virgin matrix rubber
successfully in significant proportions it is paramount to know the material composition
of these granulates. Only if parameters such as the hardness or modulus are known, can
a conclusive study of the influence of the granulate on the matrix can be conducted.
With so many various ingredients present in tyre rubber, it is obvious that the
identification and quantification of each of its components will be a laborious task, and
will require a variety of analytical tools to be used.
3.8.1 ThermoGravimetric Analysis (TGA)

According to Fernandez-Berridi et al. (2006), thermal analysis has the potential to


provide quick information for characterisation and identification of an elastomer
composition and TGA is one of the most frequently employed techniques used in these
cases. It measures weight loss with temperature and provides information about the
material composition and its thermal stability.
The thermal analysis was carried out using a Hi-ResTM TGA-Q500, supplied by TA
Instruments that was of a greater resolution when compared to a conventional TGA.
Lower heating rates also resulted in greater accuracy. With conventional TGA the
heating rate of the sample once set cannot be changed during the test. With a high
resolution TGA, the heating rate is dynamic and varies constantly in response to
changes in the decomposition rate of the material. Thus when there is no weight change
taking place the heating rate becomes high and in case of a weight change the heating
132

Chapter 3

Experimental Techniques & Materials

rate drops. This ensures a maximum weight change resolution with a substantially
reduced experimental time.
The basic composition of an elastomer can be established using differential thermal
gravimetric analysis (DTGA). The DTGA curves are usually more sensitive to weight
loss than the parent TGA curves. The reason that TGA is a very effective technique is
that elastomer TGA and DTGA curves can be used as fingerprints for the
identification of many elastomers and blends. Typical temperature for these standard
peaks are [Fernandez-Berridi et al. (2006); Agullo and Borros (2002)]:
EPDM ~ 475 C
NR ~ 380 C
SBR ~ 420 C
Previous research by Williams and Besler (1995) also showed that NR always
decomposes at lower temperatures than SBR.
To carry out the analysis the following steps were undertaken:
a) Approximately 10mg of the sample was loaded onto the balance and heated
from room temperature to 600C at approximately 40C/min. The heating was
done with a continuous flow of nitrogen at 50 ml/min. This would decompose
the polymer content in the sample.
b) Upon complete decomposition of the polymer content, the weight change
reached a plateau (which usually happened before 600C). The sample was then
allowed to cool to 400C.
c) The sample was reheated from 400C but the purge gas was switched to oxygen.
This switching over results in the carbon black being oxidised in the sample and
thus the weight change would be due to the decomposition of the carbon black.
d) The run was complete when the weight change had reached a plateau at a much
lower value.
The whole run took about 90 minutes to complete. The resolution setting used was 4
which meant that when there was no weight change taking place the heating rate
remained at 40C/min. Upon detecting a weight change, the heating rate would drop
down to 0.01C/min until the end of the weight loss.

133

Chapter 3

Experimental Techniques & Materials

Figure 3-18 TGA plot showing the weight loss with temperature for SBR 70. The dotted
line is the DTGA curve.

In order to test the accuracy of this technique, granulates of known composition were
analysed using TGA. Figure 3-18 is a plot showing the TGA and DTGA curves for SBR
70 whose composition is shown in Table 3-1. DTGA curves were used to find out the
onset of weight loss. So under the inert atmosphere of nitrogen that was maintained
from 25C to 600C, there was a 7.4% weight loss because of volatiles that was
followed by a 49.9% weight loss which is due to the polymer decomposition. The
DTGA curve shows a corresponding peak at 420C which implies that the polymer is
SBR. The polymer degradation is complete by 600C and then the next step is cooling
to 400C followed by the oxidation step where the nitrogen gas is switched to oxygen.
According to Agullo and Borros (2002) this cooling step is crucial prior to the oxidation
step without which there would be an error in the estimated quantity of carbon black.
The oxidation step is carried on until the curve reaches a plateau. The residue in the
sample pan of the instrument is 6.5% of the overall weight which is normal for
compounds that contain mineral fillers and is probably a combination of some mineral
filler and ash. If the residue quantity is about 20% then it implies that the material is
134

Chapter 3

Experimental Techniques & Materials

highly filled and contains some inorganic fillers in addition to the carbon black making
it harder to determine the rubber formulation.
Table 3-3 Comparison of TGA analysis of SBR 70 with the actual formulation data.

Component

TGA data

Actual data

SBR 1500

49.9%

54.4%

Carbon Black N330(HAF)

36.2%

38.1%

Figure 3-19 TGA plot of granulates labelled compound X.

From Table 3-3 it is evident that the technique was reasonably accurate and could be
used to quantify the amount of polymer and carbon black present in the material. Figure
3-19 shows one such TGA run that was carried out on granulates obtained by grinding

135

Chapter 3

Experimental Techniques & Materials

tyre tread whose exact composition was not known. For convenience this compound
will be addressed as compound X.
As granulates derived from tyres usually contain elastomer blends, a plot different from
that seen for SBR 70 was expected. On comparison with Figure 3-18, an additional peak
was seen around 340C which is a characteristic of NR [Agullo and Borros (2002);
Williams and Besler (1995)]. In addition to this there was an SBR peak at 420C. The
initial weight loss at around 200C was due to volatiles such as oil and other additives.
The TGA results also showed that the material contained 27.3% carbon black which is
typical for tread compounds. The residue was 5.7% which meant that the material
contained only limited amounts, if any, inorganic filler. As a result the rubber was a
highly extendable soft composition. As it was possible to estimate the weight
percentage using the DTGA curves, it was possible to obtain a rough estimate of the
NR/SBR ratio. The TGA calculations revealed that compound X had roughly a 60/40
blend of NR/SBR.
3.8.2 Fourier-Transform Infrared Spectroscopy (FTIR)

Infrared spectroscopy is widely used for identification of polymers as it is a nondestructive method. When a molecule is subjected to radiation of a specific frequency
(300 - 4000 cm-1), the individual bonds or functional groups within the molecule can
absorb this energy and thus vibrate (localised vibrations). The localised vibrations are
stretching (symmetrically or asymmetrically), bending, rocking, twisting and wagging.
Figure 3-20 shows three of the vibrational modes: symmetrical and asymmetrical
stretching, and bending, in a methylene group. On absorbing energy, the bond alters
vibrational state from its lowest to the next highest. According to O'Keefe (2004) each
polymer has specific absorption bands and a specific pattern which are used in their
identification. The sample is analysed using photoacoustical FTIR (PA-FTIR) and then
the spectrum is compared against a reference spectra to identify the material.

136

Chapter 3

Experimental Techniques & Materials

Figure 3-20 - Diagram showing the main three types of vibrational modes of a methylene
group.

All samples were kept moisture-free by storing them in a desiccator for at least two days
prior to characterisation. A Nicolet FTIR 800 spectrometer fitted with a photoacoustic
sampler (MTech PAS Cell) and controlled by a 680D Spectral Workstation was used to
obtain spectra in the mid-infrared region. Spectra were obtained in the wavenumber
range 4000-400 cm-1, at 8 cm-1 resolution averaging 512 scans. Each sample was purged
with helium gas for 10 minutes to remove atmospheric CO2 and the sample spectra
adjusted against a carbon black background spectrum.
O'Keefe (2004) has discussed the infrared spectra of two dozen of the most important
commercial polymers according to which a typical SBR sample would show dominant
peaks at 2925 cm-1, 2854 cm-1 and 700 cm-1. The first two are due to methylene
stretches and the last one is due to out-of-plane aromatic C-H stretch. Some of the other
characteristic peaks occur at 968 cm-1 and 912 cm-1 because of the trans and vinyl
double bond out-of-plane C-H bending respectively. Peaks should also occur at
760 cm-1 and 1452 cm-1 due to aromatic C-H deformation and aromatic ring breathing
mode respectively. If the sample has NR in it, then characteristic vibration bands at 841
cm-1 and 1378 cm-1 can also be seen.

137

Chapter 3

Experimental Techniques & Materials

Figure 3-21 PA-FTIR scan of SBR 0 granulates.

Figure 3-21 shows the result of a PA-FTIR scan carried out on SBR 0 granulates. The
peaks that are characteristic of a SBR material are present in this scan which confirms
that these are SBR 0 granulates. Thus with the help of a reference spectra it is possible
to differentiate between polymers. However when it comes to scanning carbon black
filled rubber granulates complications arise due to the presence of carbon black.
Waddell and Parker (1992) have explained that this is because carbon black absorbs
infrared radiation strongly and scatters the incident radiation extensively.
In case of PA-FTIR carbon black is run as a background and as filled granulates usually
contain more than 25% carbon black by weight, it is difficult to discern the peaks that
can help characterise the material.

138

Chapter 3

Experimental Techniques & Materials

Figure 3-22 PA-FTIR scan of compound X.

Figure 3-22 shows the PA-FTIR scan carried out on filled granulates whose
composition was also unknown. If compared with Figure 3-21 it is evident that the
peaks were less sharp and so it was more difficult to identify them. However certain
peaks that were characteristic of NR were seen at 837 cm-1 and 1377 cm-1 which made it
likely that these were filled NR granulates. However, it is difficult to rely on this scan to
confirm the composition and besides it was not possible to identify any other
constituents. Similarly PA-FTIR scans on the SBR 70 and compound X granulates were
not informative. An alternative strategy was to run a scan of rubber sample with about
35-40 wt % carbon as background and then run the granulates. However, the signal to
noise ratio was poor which rendered the observed spectrum unusable.
3.8.3 Pycnometry

A pycnometer is a vessel whose volume is precisely known and so usually it is used for
density () determination. If the mass of the sample (m) is known then density is
calculated by:

139

Chapter 3

Experimental Techniques & Materials

m
V

eq. ( 3-2)

In this study an automated gas pycnometer (AccuPyc 1330Micromeritics Ltd.) was


used to determine the density of the rubber granulates. It detects the pressure change
due to displacement of gas by a solid object. The gas used was helium as it can readily
diffuse into small pores.
Table 3-4 Results of density measurements for rubber granulates.

Granulate composition

Average density - g/cc

SBR70

1.2

SBR 0

1.05

Compound X

1.08

The procedure followed for measurement of granulate density was as follows:


1. The chamber used to hold the sample (V = 1cm3) was weighed using a mass
balance.
2. Granulates were loaded into the chamber up to 2/3rd of its capacity.
3. The weight of chamber + sample was noted (thus the sample weight was known)
and then the chamber was inserted into the holder in the instrument. It was
covered and the device was sealed with a lid.
4. The instrument was set to run 10 runs which implied that at the end of 10 gas
purges for each sample, an average density value was obtained.
The average density of various granulates were established using this technique and
these are tabulated in Table 3-4.

140

Chapter 3

Experimental Techniques & Materials

Thus using TGA, the compound X was understood to be a blend of NR and SBR with
60% NR and 40% SBR by weight. Using pycnometry, the average density of compound
X was calculated as 1.08 g/cc.
3.8.4 Particle size measurement

In order to carry out the particle size analysis of rubber granulates, a Mastersizer 2000
from Malvern Instruments, was used which is capable of analysing particle sizes over a
large range (0.02-2000 m). J. Allcocks and Sons Ltd who ground the SBR 70 and
SBR 0 sheets for the purpose of this study, had screened the particle sizes using sieves
with different hole sizes as described in section 3.2. The maximum size used for our
study ranged from 650 m as these could not pass through the 30 mesh screen. The
granulates were hence labelled as 30 mesh.
The Mastersizer 2000 works on the principle of laser diffraction which is more correctly
known as Low angle laser light scattering (LALLS) [Rawle.A. (2006)]. This method
relies on the basic principle that the angle of diffraction is inversely proportional to
particle size. It is also non-destructive and is highly reproducible.
The equipment consists of three main components
a) A laser that acts as the source of light of a fixed wavelength. Usually HeliumNeon gas lasers are used because of their better signal to noise and stability with
respect to temperature.
b) The system houses a detector for the size estimation. This usually is a silicon
slice that is photosensitive and there are usually about 16-32 discrete detectors.
c) A carrier that will pass the sample through the laser beam. In order to achieve
optimum measurement it is vital that the dispersion is good.
Before running an analysis it is essential to ensure that the recirculating column is clear
of any particles. It is therefore important to run water through the system as a
background run. In order to estimate the size of rubber granulates, a combination of
water and little quantity of a detergent (teepol) was used as the carrier to help disperse
the rubber granulates. This also prevented granulates from agglomerating. The carrier +
granulates suspension was added slowly into the dispersion chamber which contained
water upto 3/4th of its capacity. Before addition of suspension, the dispersion speed was

141

Chapter 3

Experimental Techniques & Materials

set to 2000 rpm in order to ensure complete dispersion of granulates. For every
specimen 3 runs were carried out in order to obtain an average.

Figure 3-23 Particle size analysis for SBR 70 30 mesh granulates.


d (0.1) 106.62 m

d (0.5) 336.59 m

d (0.9) 734.46 m

10% of the particles were smaller than 106 m, 50% were smaller than 336 m and
90% were smaller than 734 m.
To check how this technique compared with a conventional sieve analysis, the 30 mesh
SBR 70 granulates were analysed using the mastersizer. Figure 3-23 shows that. 90% of
the particles were smaller than 734.46 microns. Hence in terms of a sieve analysis these
granulates would definitely pass through a 20 mesh screen (~ 840m) but 50% would
not pass through a 30 mesh screen. Hence these granulates would come under the 30
mesh category which suggests that this technique is a valuable way of conducting a
particle size analysis.
The same procedure was used to analyse compound X granulates and the results are
shown in Figure 3-24. 90% of the granulates are smaller than 234m which implies that
these will definitely pass through a 60 mesh screen (~ 250m). Hence these granulates
are approximately 80 mesh.

142

Chapter 3

Experimental Techniques & Materials

Figure 3-24 Particle size analysis for compound X.


d (0.1) 69.55 m

d (0.5) 130.66 m

d (0.9) 233.9 m

Based on all the characterisation techniques covered above, it is thus possible to analyse
rubber granulates of unknown composition and extract the following information:
1. Elastomer type present. If there is a blend present, then the constituent parts of
the blend types can be identified as well.
2. It is possible to quantify the main constituents present in the granulates. Thus,
the weight percentage of elastomer(s), carbon black, oil can be estimated in the
final compound.
3. The density of granulates can be estimated.
4. Particle size analysis can be done to give an idea of the granulate size.
The characterisation techniques discussed above are useful but only give the user an
approximate idea of the granulate composition, size and density. Literature research
shows that it is possible to get much more extensive information using elemental
analysis and chemical and spectroscopic techniques that can point out the exact
constituents of rubber granulates. This was beyond the scope of this work. The nature of
this study was such that it was important to incorporate rubber granulates of a known
composition. In order to study the interface between the virgin matrix and the rubber
granulate and examine strength properties, it was essential to know the exact properties
of the granulate material. This would not be possible without conducting a detailed
chemical analysis. It is for this reason that rubber sheets of known formulations (SBR

143

Chapter 3

Experimental Techniques & Materials

70 and SBR 0) were mixed, moulded and sent to J. Allcocks and Sons Ltd for
granulation.

144

Chapter 4

Intrinsic Flaw Size

4
4.1

Intrinsic Flaw Size

Introduction

Chapter 2 describes how flaws that are inherent to the rubber are distributed randomly
within them. These intrinsic flaws weaken the material. It is widely established (as
discussed in section 2.5) that the incorporation of rubber granulates in rubber materials,
reduces their strength. One of the objectives of this work was to establish if this
reduction in strength resulted from an increase in the intrinsic flaw size (IFS) of the
composite (rubber + granulate) upon addition of granulates into the rubber matrix. The
purpose of this chapter is to investigate this and to discuss the findings of the work
related to measuring IFS. It was therefore imperative to develop a technique for the
estimation of IFS in rubber materials and to study the change in flaw size due to the
addition of granulates. As outlined in chapter 3, elastomers used in this study were SBR
70 (SBR with 70 phr carbon black) and SBR 0 (SBR with no carbon black). This
allowed a comparison between heavily filled and unfilled matrix to be made.

Figure 4-1 SEM scans to compare fracture surfaces of SBR 0 with no granulates (left)
and with 100 phr 30 mesh SBR 0 granulates incorporated (right).

One possible technique that might reveal the IFS would be to use microscopic
examinations of fracture surfaces for materials both with and without granulates

145

Chapter 4

Intrinsic Flaw Size

incorporated. Figure 4-1 shows freeze fractured surfaces of samples with and without
granulates that were examined. It is clear that the fracture surface morphology is
difficult to interpret in terms of a measure of IFS. Microscopic images revealed that it is
difficult to identify specific features that give an indication of the IFS. From these
fractured surfaces, as shown in the figure above, it may be possible however to identify
regions where granulates are incorporated. As a microscopic examination is of limited
use in identifying the IFS, it is necessary to use an alternative approach described in this
chapter.
In order to estimate the IFS, cuts were made onto the edges of tensile test specimens and
the resulting tensile strengths were noted. Section 4.3.3 includes a discussion of the
variation in tensile strength with increasing initial pre-cut size. Researchers who have
reported flaw size values in the past namely, Choi and Roland (1996), Lake and Lindley
(1965) and Lake (1972) have worked with specimens using edge cuts that are made
throughout the entire thickness of the specimen as shown in Figure 4-2. The work on
IFS discussed in this chapter is based on cuts that were made across the corner of one of
the long edges of the rubber strip specimens. The typical geometry used is shown in
Figure 4-3.
This study was carried out using corner cuts due to three reasons. Firstly, the corner of
the test piece is a probable location for a flaw to originate secondly typical flaws
commonly encountered will have similar length and breadth dimensions which is also
the case with a corner cut and is not the case for the edge cut where one side is typically
much greater than the other (Figure 4-3) hence making a corner cut more relevant to this
study. Thirdly, it was much easier to administer a really small cut using a corner cut
therefore making it possible to study strength data over a large range of flaw sizes. It
was too difficult to study the effects of small cuts using edge cuts as the cut cannot be
made easily with uniform depth. Similar to earlier work by Busfield et al. (1999), finite
element analysis (FEA) techniques were used in this study to help understand and
compare corner cuts with edge cuts. FEA was used to model both crack geometries in
tensile test piece (i) an edge crack shown in Figure 4-2 and (ii) a corner cut shown in
Figure 4-3. The model used for edge cut analysis adopts a plane stress 2D analysis. As
explained in section 2.11.1 this means that the model has an insignificant dimension in

146

Chapter 4

Intrinsic Flaw Size

one direction and is loaded in the plane of the body. The corner crack required the use
of a 3D model. Section 4.4.1 discusses the findings of this analysis.
The main purpose of this chapter is to examine the failure of matrixgranulate
composites and it aims to lay the groundwork for the chapters to follow. The work
presented in this chapter is a part of the work that has been published by the author.
Please see appendix 1.

147

Chapter 4

Intrinsic Flaw Size

a)

b)

c)

Figure 4-2 Tensile test piece with an edge cut a) Schematic b) is a undeformed half FEA
model of the same and c) is half FEA model deformed to extension ratio of 2.

148

Chapter 4

Intrinsic Flaw Size

a)

b)

c)

Figure 4-3 - (a) Schematic representing a tensile test specimen with a corner cut (b) 3D
half symmetry FEA model deformed to extension ratio of 2 (c) undeformed half FEA
model.

149

Chapter 4

4.2

Intrinsic Flaw Size

Experimental procedure for IFS

Stress strain behaviour of dumbbell shaped specimens were determined using tensile
tests described in section 3.4. These were used to avoid slippage at the grips. They were
stretched at a constant strain rate of 0.77 min-1. The value of tensile strength or stress at
which the specimen failed, was measured for each specimen. For the corner cut
specimens, the initial cut length c was also measured using an optical microscope. The
edge of the razor cut was clearly visible under an optical microscope on the final test
pieces after fracture. The distance measured was the perpendicular length between the
corner and the razor cut edge.
4.3

Mechanical characterisation stress strain behaviour

Section 2.11.4 in chapter 2 shows that based on the Rivlin (1956) approach the
following equation can be applied for a uniaxial tensile test:
W 1 W

+
= 2
2
I


2
1

eq. ( 2-32)

Where, is the engineering stress (the force required to pull the specimen divided by
the original cross-sectional area of the unstrained sample).
4.3.1 Characterising the stress versus strain behaviour of SBR 0.
Mooney (1940) used mathematical reasoning to derive the following strain energy
density function:
W = C10 ( I 1 3) + C01 ( I 2 3)

eq. ( 2-27)

C10 and C01 are material coefficients that are used to characterise the material. The
Mooney SEF was used to characterise SBR 0 as it has been used in the past successfully
for unfilled rubber vulcanisates.
From equation 2-27 it is clear that partial derivatives W / I 1 and W / I 2 equal C10
and C01 respectively. Thus equation 2-32 can therefore be rewritten as:
2C

= 2C10 + 01
2

eq. ( 4-1)

150

Chapter 4

Intrinsic Flaw Size

0.7
y = 0.38x + 0.30

-2
/ ( - ) / MPa

0.6
0.5
0.4
0.3

Curve fitting region

0.2
0.1
0
0

0.2

0.4

0.6

0.8

1/
Figure 4-4 - Plot to determine the coefficients for Mooney SEF of SBR 0.

A plot of reduced stress (

) versus 1/ should then give us a straight line of slope


2

2C01 and intercept 2C10. The coefficients used to characterise SBR 0 were determined
using experimental data obtained from tensile tests carried out on SBR 0 specimens.
Figure 4-4 is a Mooney plot for SBR 0 where the reduced stress,

is plotted
( 2 )

against 1/. A straight line fitted the data well for moderate and large strains. With these
new coefficient values equation 4-1 is now rewritten as:

0.38
= 0.3 +
2

eq. ( 4-2)

Thus comparing equations 4-2 and 4-1 the Mooney coefficients for SBR 0 were
determined as:
C10 = 0.15 MPa and C01 = 0.19 MPa

With these new coefficient values equation 2-27 is now rewritten as:

151

Chapter 4

Intrinsic Flaw Size

W = 0.15( I 1 3) + 0.19( I 2 3)

eq. ( 4-3)

Figure 4-5 compares the experimental tensile test data of SBR 0 with the theoretical
behaviour predicted by the Mooney SEF. Over the entire range of strains the fit was
excellent.

Experimental data for SBR 0


Mooney function data fit

1.8
1.6
Stress, / MPa

1.4
1.2
1
0.8
0.6
0.4
0.2
0
1

1.5

2
2.5
Extension ratio,

3.5

Figure 4-5 - Stress versus extension ratio curves for SBR 0 measured in tension and the
Mooney function used to fit the experimental data.

4.3.2 Characterising the stress versus strain behaviour of SBR 70.

Gregory (1979) has shown that for carbon black filled elastomers there exists a simple
relationship between reduced stress (which is the expression on the left hand side of
equation 2-32) and the first invariant I1 only. This implies that for filled elastomers,
W / I 1 is much larger than W / I 2 . Using this approach Yeoh (1990) proposed a new

SEF which makes a simplifying assumption that W / I 2 is equal to zero. As SBR 70

152

Chapter 4

Intrinsic Flaw Size

was filled with carbon black, it was appropriate to use this Yeoh SEF which has the
following form:

W = C10 ( I 1 3) + C 20 ( I 1 3) 2 + C 30 ( I 1 3) 3

eq. (2-35)

Here C10, C20 and C30 are the constants required to establish the material properties at a
given temperature and strain rate. If this equation is partially differentiated with respect
to I1, then it becomes:

W
2
= C10 + 2C 20 (I 1 3) + 3C 30 I 1 3)
I 1

eq. ( 4-4)

Assuming that W / I 2 is small and that W depends only on I1 allows equations 2-32
and 4-4 to be combined and rewritten as:

2
= 2C10 + 4C 20 (I 1 3) + 6C 30 I 1 3)
2

eq. ( 4-5)

This is a quadratic equation in (I 1 3) . To fit the three coefficients the stressextension


ratio behaviour was re-plotted as reduced stress

against (I 1 3) for SBR 70.


( 2 )

As shown in Figure 4-6, a best fit regression analysis was used to derive the coefficients
from the tensile test data. As was expected and has been shown by previous researchers,
the degree of fit was good for moderate to large strains but was less accurate at small
strains. As the analysis is principally interested in the behaviour at break, this was
thought suitable.
The second degree polynomial that was used to fit these data is:

2
= 2.65 + 0.22(I 1 3) + 0.026 I 1 3)
2

eq. ( 4-6)

Therefore, the coefficients SBR 70 were determined as being:


C10 = 1.32 MPa, C20 = 0.055 MPa and C30 = 0.004 MPa

Figure 4-7 shows good agreement between experimental tensile test data for SBR 70
and the predicted behaviour using the Yeoh SEF. These coefficient values could now be
used to characterise SBR 70 in all the models so that the elastomeric behaviour could be
reliably predicted at moderate to large strains.

153

Chapter 4

Intrinsic Flaw Size

4.5
4

y = 0.026x2 + 0.22x + 2.65

-2
/ ( - ) / MPa

3.5
3
2.5
2

Curve fitting region

1.5
1
0.5
0
0

8
I1-3

10

12

14

16

Figure 4-6 - Plot to determine the coefficients for Yeoh SEF of SBR 70.

20
18

Stress, / MPa

16

Experimental data

14

Yeoh function data fit

12
10
8
6
4
2
0
1

1.5

2.5

3.5

Extension ratio,
Figure 4-7 - Stress versus extension ratio curves for SBR 70 measured in tension and the
Yeoh function used to fit the experimental data.

154

Chapter 4

Intrinsic Flaw Size

Thus for all cases where the rubber was filled, the Yeoh SEF was used to characterise
the behaviour in the FEA model and for unfilled rubber the Mooney SEF was used.

Table 4-1 - Stress at break values for the material types used in this study.

Nomenclature

Stress at Break (b) / MPa

Standard Deviation / MPa

M0

2.14

0.14

M0-G0(21%)

1.86

0.13

M0-G0(47%)

1.79

0.64

M0-G70(17%)

2.06

0.27

M0-G70(41%)

4.23

0.14

M70

27.44

1.22

M70-G70(14%)

22.70

0.63

M70-G70(35%)

17.53

1.84

M70-G0(17%)

17.16

1.43

M70-G0(41%)

11.23

1.66

4.3.3 Tensile strength versus cut lengths

Tensile strength or stress at break (b) is the ratio of the force required to break the
specimen to the original cross-sectional area of the unstrained sample. Section 3.3
explains in detail the range of samples and material combinations tested. Details of the
155

Chapter 4

Intrinsic Flaw Size

actual combinations of materials that make up the matrix and granulates being given in
Table 3-2. A minimum of 5 specimens for each combination were stretched uniaxially
with no intentional flaw/cut introduced in order to obtain their tensile strengths so that
an average value could be taken. The tensile strengths of various materials tested are
tabulated in Table 4-1.
30

M70
M70-G70(14%)

Tensile strength, b / MPa

25

M70-G70(35%)

20
15
10
5
0
0

200

400

600

800

1000

1200

1400

Cut length, c / microns


Figure 4-8 - Effect of precut length on stress at break M70-G70(X) (SBR 70 matrix and 30
mesh SBR 70 granulates).

The remaining 25-30 specimens of each type were tested with corner cuts of different
sizes introduced into them using a sharp razor. For each material the stress at break, b
values were plotted against their respective initial cut lengths, c. Figure 4-8 is one such
plot of all M70-G70(X) specimens (where SBR 70 is the matrix and the granulates, if
incorporated, were also SBR 70) tested against their respective cut lengths (c). There is
a decrease in strength (b) with increase in c, similar to that reported by Berry (1961)
and Hamed (1983) and a decrease in b as the amount of granulates incorporated was
increased as has also been shown previously by Burgoyne et.al. (1976), Han and Han
(2002) and Gibala et al. (1999). This was also seen for M0-G0(X) [Figure 4-9] and
M70-G0(X) [Figure 4-11].

156

Chapter 4

Intrinsic Flaw Size

However, the exact opposite was seen in case of M0-G70(X) [Figure 4-10] where b
values increased as quantity of granulates incorporated was increased. The rise in b was
quite significant when 100 phr granulates were incorporated as increase in b was twice
that seen in case of no granulates. This indicated that these SBR granulates that
contained 70 phr of carbon black when incorporated into an unfilled SBR matrix,
reinforced the matrix. For M70-G0(X), the decrease in strength with incorporation of
larger amounts of granulates was greatest suggesting that the matrix-granulate
interaction was weakest for this combination of filled matrix and the unfilled granulates.

2.5

M0

Tensile strength, b / MPa

M0-G0(21%)
M0-G0(47%)

1.5

0.5

0
0

200

400

600

800

1000

1200

1400

Cut length, c /microns


Figure 4-9 - Effect of precut length on stress at break M0-G0(X) (SBR 0 matrix and 30
mesh SBR 0 granulates)

157

Tensile strength, b / MPa

Chapter 4

Intrinsic Flaw Size

4.5

M0-G70(41%)

M0-G70(17%)
M0

3.5
3
2.5
2
1.5
1
0.5
0
0

200

400

600

800

1000

1200

1400

1600

Cut length, c / microns


Figure 4-10 - Effect of precut length on stress at break of M0-G70(X) (SBR 0 matrix and
30 mesh SBR 70 granulates).

30

Tensile strength, b / MPa

M70
M70-G0(17%)

25

M70-G0(41%)

20
15
10
5
0
0

200

400

600
800
1000
Cut length, c / microns

1200

1400

1600

Figure 4-11 - Effect of precut length on stress at break of M70-G0(X) (SBR 70 matrix and
30 mesh SBR 0 granulates).

158

Chapter 4

4.4

Intrinsic Flaw Size

Intrinsic flaw size (IFS)

4.4.1 Strain energy release rate for edge and corner cut specimens.

Rivlin and Thomas (1953) showed that for a tensile test specimen with an edge cut of
length c as shown in Figure 4-13a:
T = 2K()Wc

eq. (2-4)

Where W is the strain energy density in the bulk of the test piece (away from the cut)
when the cut grows and K() is a function of strain .
Greensmith (1963) proposed that this strain dependence is:
K ( ) = (1 + )

1 / 2

eq. (2-6)

Busfield et al. (1999) has reported that in case of a tensile test specimen with a corner
cut:
T = 2 K ' ( )Wc

eq. (2-10)

where K ' () has a similar relationship to K() found in equations 2-4 and 2-6 but which
is less by a factor of approximately 2 for all strains. This implies that a corner cut is less
effective in raising stress value as compared to an edge cut.
Using FEA, models of tensile test specimens with edge (Figure 4-2) and corner cut
(Figure 4-3) were prepared in an IDEAS preprocessor and solved using ABAQUS.
This exercise was carried out to compare the difference in tearing energies with strain
for a rubber tensile test specimen with an edge cut with one with a corner cut. As
explained in section 2.8.1 the strain energy release rate is defined as:
U
T =

A l

eq. (2-1)

As detailed in section 3.8, for finite element models the strain energy release rates were
computed using the energy balance approach. Figure 4-12 is a plot that compares
tearing energy values (T) versus crack lengths (c) for both these models and obtained
using the node release or virtual crack extension technique that is also explained in
section 3.8. It was evident from this plot that for a given crack length, the amount of
energy available for the crack to propagate was almost twice for a specimen with an

159

Chapter 4

Intrinsic Flaw Size

edge cut as compared to a specimen with a corner cut. Thus from this analysis and work
done by Busfield et al. (1999) in the past, it was understood that for a corner cut to be as
effective as an edge cut in terms of causing a tear within a specimen, it would have to be
approximately twice the length of the edge cut.
Researchers such as Choi and Roland (1996), Lake and Lindley (1965) and Lake (1972)
in the past have reported IFS in the range of 50m for rubber specimens but they tested
specimens with edge cuts. They concluded from their observations that the effective
flaw size was equivalent to an effective edge razor cut.
In this study as corner cuts were made on tensile test specimens before straining them,
the effective flaw size is more likely to be equivalent to a corner razor cut and hence
expected to be notably larger flaw size values. The difference between the cut length
measurements on the fracture surface is shown in Figure 4-13.

SBR 70 - 100% Strain

Edge cut
Corner cut

Tearing energy, T / kJm-2

0.6
0.5
0.4
0.3
0.2
0.1
0
0

100

200

300

400

500

Crack length, c /microns


Figure 4-12 - Plot to compare the amount of energy available to propagate a crack for a
M70 tensile test specimen with an edge cut to that with a corner cut at 100% strain.

160

Chapter 4

Intrinsic Flaw Size

Figure 4-13 - Specimen with a cut along A-A cross-section (a) Fracture surface with edge
cut (b) Fracture surface with corner cut, c indicates the length of the cut.

4.4.2 Flaw size equivalent to corner razor cut

Further to section 2.9, Griffith (1921) suggested that a flaw/crack would propagate in a
stressed material and lead to fracture if the propagation brought about a reduction in
elastic strain energy, U that is greater than the surface free energy of the newly formed
fracture surfaces and hence fracture takes place:

U 1 A
Gc
c 2 c

eq. ( 4-7)

where A is the surface area of the specimen, c denotes the length of radius of the
fracture and the amount of energy required to propagate the fracture by unit area is
denoted by Gc. The factor of 1/2 takes into account the two newly formed surfaces.
It is explained in section 2.9.1 that if W is the strain energy per unit volume in the bulk
of the a tensile test piece with an edge cut of depth c, then the reduction in stored elastic
energy (U) is kc2tW. Here k is a dimensionless function of strain. Thus,

161

Chapter 4

Intrinsic Flaw Size

U
= 2kctW
c

eq. ( 4-8)

For edge cut geometry where the area of a single surface of the crack was ct:
A
= 2t
c

eq. ( 4-9)

Substituting equations 4-8 and 4-9 in 4-7 gives:


2kcW Gc

eq. ( 4-10)

For a linear elastic material where k at infinitesimal strains the strain energy density
can be expressed as:
W =

2
2E

eq. ( 4-11)

where E is the Youngs modulus. W satisfies equation 4-11 when the stress is equal to
the stress at break, b.
Thus using the above information and equation 4-11, equation 4-10 can now be
rewritten as:

c b 2
E

1/ 2

G E
= GC or b = c
c

eq. ( 4-12)

Berry (1961) confirmed that for a linearly elastic material if c is the cut length and b is
the stress at break, then at small strains:

b c 0.5

eq. ( 4-13)

However for these elastomers, it is seen that an empirical fit has a different power
relationship which is likely to result from the material nonlinearity at the large strains
encountered at break.
A double log plot of the stress at break, b versus cut length, c an example of which is
shown in Figure 4-14 allows the IFS to be calculated, as the data fits reasonably well on
to a straight line on this plot. In order to obtain the IFS of the material, this line was
extrapolated to the value of the tensile strength of specimens without a precut.

162

Chapter 4

Intrinsic Flaw Size

No pre-cut

With pre-cuts

1.45
Log (Tensile strength, b / MPa)

y = 1.438

1.4
1.35
1.3

y=

IFS = 131 m

1.25
-0.3
97x
+

1.0
88

1.2
1.15
1.1
1.05
1

-1

-0.8

-0.6
-0.4
-0.2
Log (Cut length, c / mm)

Figure 4-14 - Effect of precut length on stress at break for M70 (SBR 70 without
granulates). The point of intersection is the corner cut equivalent IFS for this material.

As discussed in chapter 2, a similar technique has been used previously by Gent (1978)
and Hamed (1983) to estimate the IFS but in those cases the cuts introduced on
specimens were edge cuts. As explained earlier, the intrinsic flaw sizes that are
investigated in this work will be flaws of equivalent size to a corner razor cut.
Figure 4-14 shows such a double log plot for a pure SBR 70 compound without
granulates incorporated (M70) where the slope of the straight line was -0.4. This and the
similar deviations observed in case of the materials tested is comparable to the small
strain linear elastic value of -0.5 with the difference being easily attributed to non linear
material behaviour at high strains.
Table 4-2 lists the corner cut equivalent IFS values thus calculated for the different
material combinations tested. The lowest corner cut equivalent IFS value was 70m for
M0 (SBR 0 with no granulates) and the highest was 470m for M70-G0(41%) (SBR
with 70 phr carbon black and 41% SBR 0 granulates by volume). For all types, the
corner cut equivalent IFS increased with an increase in the quantity of granulates
incorporated into the matrix. This could be due to the tendency of these granulates to

163

Chapter 4

Intrinsic Flaw Size

cluster as they are incorporated into the matrix and the resulting intergranulate interface
is likely to be weak. However, this has not yet been observed directly. The reduction in
the strength of the matrix could therefore be attributed to an increase in corner cut
equivalent IFS, when granulates were added.
Table 4-2 - IFS values for the different materials tested.

Corner cut equivalent

Standard

IFS/m

deviation/m

M0

70

1.2

M0-G0(21%)

75

1.5

M0-G0(47%)

125

2.2

M0-G70(17%)

82

2.4

M0-G70(41%)

150

2.3

M70

130

1.3

M70-G70(14%)

245

2.6

M70-G70(35%)

315

2.7

M70-G0(17%)

300

3.5

M70-G0(41%)

475

4.5

Polymer type

The corner cut equivalent IFS versus quantity of granulates data is plotted in Figure
4-15. This shows that the IFS values for M70-G70(X) (where both the matrix and
granulates are filled) and M70-G0(X) and M70-G0(X) materials (where either the matrix
or granulates are filled) were higher than M0-G0(X) (neither matrix nor granulates are
filled) and this indicated that the addition of carbon black also caused an increase in the
corner cut equivalent IFS. Busfield et al. (1999) and Lake and Lindley (1965) have

164

Chapter 4

Intrinsic Flaw Size

reported an increase in edge cut equivalent IFS values due to presence of carbon black.
The increase in the IFS with granulate loading is quite sharp for M70-G0(X) when
compared to other samples in Figure 4-15. This is probably due to the mismatch in
modulus between the filled SBR matrix and the unfilled SBR granulates.
Gent and Lindley (1958) conducted some studies on cavitation in rubber-like solids and
observed that these solids contained a size distribution of cavities within them of which
the larger cavities grew into visible cracks when the stress (negative hydrostatic
pressure) reached a value of 5E/6. Their work thus showed that the stress required for
cavitation was directly related to the Youngs modulus of the material and in the
process laid down the foundation for similar work that followed which is discussed in
section 2.10.2. It is possible that such cavitation phenomenon might be taking place in
case of our materials as well.
Figure 4-16 shows the dependence of b on c on linear axes for all M70-G70(X)
specimens, with all the data points represented by a single curve. Albeit the initial b
(with no precut) was quite different for the three materials, above the respective critical
cut lengths, the b values for all the materials were similar at given cut lengths. All the
data points could be represented by a single curve. This implied that for the same
matrix, the crack propagation was independent to the volume fraction of granulates.
This master curve can be used to deduce the intrinsic flaw size for materials without
precuts introduced, from a measure of their tensile strength. Thus, as shown in Figure
4-16, the IFS of M70 (SBR 70 without granulates) was estimated to approximately
140m as its b was known from tensile tests. This correlated well with the IFS value
obtained for the same material using the double log axes techniques (130m). Similar
observations were made for the case of M0-G0(X) specimens.
In Figure 4-17, the b of M0-G0(47%) (SBR 0 + 47% SBR 0 granulates) was
extrapolated to the curve and the IFS was approximated as 115m. This was again close
to the value of 125m obtained by the double log plot. Both techniques allowed us to
determine corner cut equivalent IFS values with an accuracy of 20%. However, it was
thought that the method of estimation of IFS using the double log method was more
robust as it involved extrapolation of a straight line (to the strength with no precut)
compared to the extrapolation of a curve in the second technique.

165

Chapter 4

Intrinsic Flaw Size

M70-G0(X)

500

M70-G70(X)

450

M0-G70(X)

IFS / microns

400

M0-G0(X)

350
300
250
200
150
100
50
0
0

20

40

60

80

100

120

Quantity of granulates / phr


Figure 4-15 - Comparison of IFS values for various materials tested.

30

M70 with precuts

b ~ 27 MPa

25
20
No granulates

Tensile strength, b / MPa

M70 with no precut

15
10
5

IFS ~ 140 m

0
0

200

400

600

800

1000

1200

1400

1600

Cut length, c / microns


Figure 4-16 - b versus cut lengths curve for M70-G70(X) specimens. b of M70 without a
cut is extrapolated to the curve to obtain the corner cut equivalent IFS.

166

Chapter 4

Intrinsic Flaw Size

M0 with precuts

b ~ 1.8 MPa

1.5
100 phr granulates

Tensile strength, b / MPa

M0-G0(47%) without a cut

0.5

IFS ~ 115 m

0
0

200

400

600
800
1000
Cut length, c / microns

1200

1400

Figure 4-17 - b versus cut lengths curve for M0-G0(X) specimens. b of


M0-G0(47%) without a cut is extrapolated to the curve to obtain the corner cut equivalent
IFS.

4.5

Conclusions

Tensile tests were carried out on filled and unfilled SBR compounds with a range of
different volume fractions of granulates incorporated in them. The stress at break (b)
values, were plotted against the respective precut lengths (c) and they were found to
decrease with increasing c. A technique for estimation of the corner cut equivalent
Intrinsic Flaw Sizes (IFS) of the composites was developed. These values were found to
increase with increasing quantity of granulates and this could explain the decrease in
strength of an elastomer when ground rubber was mixed into it. As now it is evident that
addition of granulates increases the microscopic flaw size thus initiating failure, it is
now important to identify the crack growth mechanism most likely to cause catastrophic
failure in real test materials.

167

Chapter 5

Matrix-Granulate Interface

5
5.1

Matrix-Granulate Interface

Introduction

The previous chapter describes a technique to estimate the size of the single largest flaw
that is present in a rubber specimen. This technique was used to estimate the IFS of
various matrix-granulate combinations and it was observed that there was an increase in
the flaw size of the rubber matrix upon the incorporation of rubber granulates which
continues to increase as the volume fraction of granulates increases.
Poor adhesion between the filler and the matrix leading to dewetting of fillers has been
suggested as a cause for this deterioration of mechanical properties by Burford and
Pittolo (1984). If the bonding between the matrix and granulate is weak then the
interface between them can be a site for a flaw to originate. It is therefore imperative to
develop a technique that can quantify this interfacial strength. As the volume fraction of
granulates in the matrix increases, the inter granulate separation reduces. In local areas
within the matrix, the granulates agglomerate as a result of which they come in contact
with each other. Hence whilst examining the strength of such composites, it is important
to study not just the interface between matrix and granulates but also the extent of the
interfacial strength between granules.
There is also a possibility that once a flaw is initiated, the crack could propagate in three
different ways either through the granule, in the matrix or along the interface. The crack
propagation behaviour was studied using FEA models and the direction along which the
crack grows can be examined using fracture mechanics. 2D models of matrix with
granulate embedded in it were prepared using I-DEAS v.11 and run using ABAQUS
v.6.6-1. To simplify the analysis, the granulates were assumed to be spherical in shape
and evenly dispersed throughout the matrix.
The work described in this chapter aims to examine the following key issues:
a) What are the likely sites for a flaw to occur in a matrix-granulate composite and
do these flaws act as crack initiators?

168

Chapter 5

Matrix-Granulate Interface

b) Why does the incorporation of increasing quantities of granulates result in a


composite material with a lower tensile strength?
c) Once a flaw is initiated, how does the crack propagate and is it possible to
estimate the time required for the crack to grow from one point to another?
5.2

Experimental procedure

Section 3.4 describes the tensile test procedure that was used to obtain the uniaxial
stress-strain data to compute the strain energy density (W) of the matrix material at a
specific strain.
5.2.1 Measurement of tearing energy
The technique used to measure the tearing energy is described in section 3.5. Sheets
with the same composition and curing conditions as the matrix and granulates were
prepared separately. Granulates were prepared using ambient grinding until the required
size reduction was obtained. In order to replicate the cure state found in the granulate
materials, sheets with the same composition as the granulates underwent two cure
cycles.
Rectangular strips with dimensions of 100 x 20 x 2 mm were cut from these sheets and
along the centre line of these strips a cut of length 30 mm was introduced to make a
trouser test piece as described in section 3.5. The two legs were then clamped in the
grips of an Instron tensile test machine and pulled apart at a constant strain rate of 0.77
min-1. The average force required to propagate the tear was used to compute the tearing
energy using equation 2-2. For each material type, atleast 5 strips were tested and the
tearing curves were found to be reproducible.
The work presented in this chapter is a part of the work that has been published by the
author. Please see appendix 1.
5.2.2 Measurement of interfacial strength
As detailed in section 3.6, this was a 2 stage process. The sheet with the same
composition as granulates was prepared by subjecting it to two curing cycles (each
curing cycle160C, 15 minutes). This sheet was then kept aside for 3 days. The sheet

169

Chapter 5

Matrix-Granulate Interface

with the same composition as the matrix was mixed by incorporating all the ingredients
using the two roll mill but not cured. The two sheets were then compressed together
with a thin strip of PTFE coated tape at one end so that once the curing cycle (160C for
15 minutes) was finished the two free ends could be gripped by the clamps in the
Instron tensile test machine. A flat interface was thus created between the two sheets
where the sheet representing granulates was double cured and the sheet representing the
matrix was single cured.
Quantification of interfacial strength was done by cutting out five separate strips which
were then pulled apart at a constant strain rate of 0.77 min-1 to separate the strips across
the interface. The peel energy, P, was calculated using the average force required to
propagate the tear along the interface.
The interface between granulates was also obtained using similar procedure as
explained above. However, as there was no matrix this time, the sheets with the same
composition as granulates were mixed, cured (160C, 15 minutes) and kept aside for 3
days. After this period, these sheets were then cured together in a press again for 160C
for 15 minutes with the thin PTFE layer in between at one end. This resulted in a flat
interface between two double cured sheets. Once again in order to obtain the interfacial
strength between granulates, five strips were cut and each of these were gripped and
pulled apart a constant rate to separate the strips across the interface. The peeling curves
were reproducible.
5.3

Tearing and peeling energy

Rivlin and Thomas (1953) have shown that tearing energy is independent of the
geometry of the specimen. But it will depend on the thickness of the leg of the
specimen. Figure 5-1 is therefore a normalised plot that compares the force required to
propagate a tear through a specimen of SBR 70 and SBR 0. As expected, it becomes
clear from the plot that once a crack is initiated, it will propagate much more quickly
through SBR 0 specimens than SBR 70.
A similar comparison between single cured SBR 70 and double cured SBR 70 is shown
in Figure 5-2. The ability of SBR 70 to resist crack growth reduces when it is subjected
to a double curing cycle.

170

Chapter 5

Matrix-Granulate Interface

SBR 70
SBR 0

Trouser tear

10

SBR 70

-1
F /t - Nmm

8
6
4
2

SBR 0

0
0

50

100
Extension / mm

150

200

Figure 5-1 Plot comparing the tearing force for single cured SBR 70 and SBR 0
specimens at a strain rate of 0.77 min-1.

SBR 70

12
Single cured

-1
F /t - Nmm

10
8

Double cured

6
4
2
0
0

50

100

150

200

Displacement / mm
Figure 5-2 Comparison between force required to propagate a tear through a single
cured SBR 70 (representing matrix) and double cured SBR 70 (representing granulate).

171

Chapter 5

Matrix-Granulate Interface

Equation 2-2 was used to compute the tearing energy of the trouser tear test and is
repeated here for convenience:

T=

2 F
wW
t

eq. ( 2-2)

where w is the width of each leg, ~10 mm, t is the thickness of each leg, ~2mm and W is
the elastic strain energy density in the legs at the maximum extension.
To compute W, the stress-strain data of the material is required. Using Figure 5-1 the
average force required to propagate a tear in SBR 70 was obtained. The extension ratio
in the leg of the test piece that corresponds to this average stress was obtained from a
stress-strain plot of a new specimen of SBR 70. The corresponding value of W is the
area under this plot up to that extension ratio.

Strain Energy Density/ 10 Jm

-3

SBR 70

Single cured

50
45
40
35
30
25
20
15
10
5
0
1

Extension Ratio

Figure 5-3 Plot used to obtain the value of strain energy density, W at any strain for SBR
70 in order to compute the tearing energy.

Calculating the area under the curve of a stress-strain plot of SBR 70 for all strains
allows a strain energy density versus extension ratio plot to be constructed as shown in

172

Chapter 5

Matrix-Granulate Interface

Figure 5-3. These values are then substituted into equation 2-2 to compute the tearing
energy.

SBR 0

Single cured
Double cured

2.5

F/t - Nmm-1

2
1.5
1
0.5
0
0

20

40

60
80
Extension / mm

100

120

Figure 5-4 - Comparison between force required to propagate a tear through a single
cured SBR 0 (representing matrix) and double cured SBR 0 (representing granulate).

Similarly Figure 5-4 shows how the SBR 0 single cured compares with SBR 0 double
cured. The W versus relation for SBR 0 double cured is shown in Figure 5-5. W versus
curves were also calculated for SBR 70 double cured systems (to represent the filled
granulates) and SBR 0 single cured systems (to represent the unfilled matrix).
Using equation 2-2 the tearing energy for the materials with the same composition as
matrix and granulates were determined.
Figure 5-6 shows representative peeling curves measured along the interface between
matrix and granulates where both had the same composition. Evident from the plot is
that the force required to peel the filled matrix-granulate combination is about 3 times
higher than that required to peel the unfilled matrix-granulate combination.

173

Chapter 5

Matrix-Granulate Interface

SBR 0
3

-3
6

Strain Energy Density / 10 Jm

Double cured

2.5
2
1.5
1
0.5
0
1

1.5

2.5

3.5

Extension Ratio

Figure 5-5 - Plot used to show the relation between W and extension ratio for
SBR 0 double cured (representing granulate) in order to compute the tearing energy.

SBR 70
SBR 0

-1
F /w - Nmm

Matrix-Granulate Interface

5
4.5
4
3.5
3
2.5
2
1.5
1
0.5
0

M70 - G70

M0 - G0

50

100

150

200

250

300

350

Extension / mm
Figure 5-6 Normalised plot that compares the peeling force between filled matrix-filled
granulate and unfilled matrix-unfilled granulate combinations.

174

Chapter 5

Matrix-Granulate Interface

The granulate-granulate interface was studied using Figure 5-7. Here too it is seen that
the interface between filled granulates is stronger than that between unfilled granulates
which is particularly weak.
The peeling energy was computed using the following equation:
P=

2 F
tW
w

eq. ( 2-3)

where w is the width of each leg, ~20 mm, t is the thickness of each leg, ~2mm and W is
the elastic strain energy density in the legs.

Granulate-Granulate Interface

2.5

SBR 70 + SBR 70
SBR 0 + SBR 0

-1
F /w - Nmm

2
1.5
G70 - G70

1
0.5
G0 - G0

0
0

50

100

150
200
Extension / mm

250

300

Figure 5-7 - Normalised plot that compares the peeling force between filled
granulate-filled granulate and unfilled granulate-unfilled granulate combinations.

Similar to the tearing energy calculations, the average peeling force, F, was obtained
from the normalised plot as shown in Figure 5-6. In order to obtain the corresponding
extension ratio, a stress-extension ratio curve of a new specimen with the same
composition as that of the matrix was used. The strain energy density, W, was calculated
in the same manner as described for tearing energy. The various W versus curves (as

175

Chapter 5

Matrix-Granulate Interface

shown in Figure 5-3 for SBR 70 single cured system) that had already been plotted for
determining the tearing energies were now also used to determine the peeling energies
for the different material types.
It was not possible to prepare a peel specimen for the M70-G0 matrix-granulate
combination. This was because the sheet representing the granulate composition (in this
case SBR 0) when subjected to a second cure cycle was very soft and so during curing,
the sheet deformed extensively and once removed from the mould it was not possible to
prepare a system with a sheet representing the matrix (SBR 70 single cured) and a sheet
representing the granulate (SBR 0 double cured) cured together.
5.3.1 Matrix and granulate interface

The trouser tear and peel test results for the types where matrix and granulates were of
same composition (M70-G70(X) and M0-G0(X)) are tabulated in Table 5-1. It was
found that for both materials the tear resistance of single cured compound was greater
than the tear resistance of double cured compound, which was then in turn greater than
the interfacial peel energy.

Table 5-1 - Energy table to study interface between matrix and granulates of same
composition [M70-G70(X) and M0-G0(X)].

Peel Energy,P/kJ/m2
Type

Tear Energy,T/kJ/m2
Single cured

Double cured

(matrix)

(granulates)

M0-G0(X)

0.8

3.7

2.6

M70-G70(X)

8.7

20

16

Table 5-2 shows that for M0-G70(X), as granulates were of a much higher modulus than
the matrix, the tear resistance of the doubly cured compound was the highest followed

176

Chapter 5

Matrix-Granulate Interface

by the tear resistance of single cured compound and interfacial peel energy being the
lowest.
In case of M0-G0(X), the peel energy was only about 20% of the tear energy of the
single cured material. Similarly with M70-G70(X) materials, the peel energy was only
about 40% of the tear energy of the single cured material. These low values indicated
that the interfaces between the matrix and granulates were weaker than the bulk
materials and the adhesion values were low. This agreed with previous findings of
Burford and Pittolo (1984). The tear energy for the double cured material that was
representative of the composition of granulates was consistently lower than single cured
systems. The tear resistance of the filled SBR materials (M70-G70(X)) was much higher
than their unfilled counterparts (M0-G0(X)). See Table 5-1. This was obviously due to
the reinforcement of SBR by the carbon black.

Table 5-2 - Energy table to study interface between matrix and granulates of different
composition [M0-G70(X)) and M70-G0(X)].

Peel Energy, P/kJ/m2


Type

Tear Energy,T/kJ/m2
Single cured

Double cured

(matrix)

(granulates)

M0-G70(X)

2.9

3.7

16

M70-G0(X)

20

2.6

*Unable to prepare a peel specimen in which SBR 0 underwent a double cure cycle as the
material was too soft and deformed too greatly in the second cure cycle.

Chapter 4 showed that the addition of carbon black caused an apparent increase in the
corner cut equivalent IFS of the virgin matrix (Table 4-2, chapter 4) which might be
supposed to decrease the strength. However, the increase in strength resulting from the
additional reinforcement more than compensates for this.

177

Chapter 5

Matrix-Granulate Interface

As seen in Table 5-1, the interfacial strength of filled SBR was about 10 times greater
than that of unfilled compound. This could be attributed to the increased viscoelastic
dissipation for a filled material when compared to an unfilled one. Comparison of
results in Table 5-1 and Table 5-2 shows that the interfacial strength of M0-G70(X) was
approximately 3.5 times that of M0-G0(X) which clearly indicated that interaction
between an unfilled single cured material (representing the matrix) and filled double
cured material (representing granulates) was much better than that observed with
M0-G0(X). This also probably explains the improved b contrasted with the unfilled
behaviour when a higher quantity of granulates were incorporated into the matrix (seen
in Figure 4-9 in chapter 4).
5.3.2 Granulate and granulate interface

The peel test results that were used to compare the interfacial strengths between
granulates are shown in Table 5-3. In case of M0-G0(X) and M70-G0(X) granulates
(SBR 0), the peel energy was only about 6% of the tear energy of the unfilled
granulates. For granulates used in M70-G70(X) and M0-G70(X) (SBR 70), the peel
energy was only about 9% of the respective tear energy.
Hence it could be seen that in all cases the interface between granulates was the weakest
followed by the interface between the matrix and granulates. This was probably the
reason that as more granulates were incorporated into the matrix, the composite got
weaker (as seen in Figures 4-7, 4-8 and 4-10 in chapter 4).

Table 5-3 - Energy table to study interface between granulates

Granulates type

Peel Energy, P/kJ/m2

Tear Energy, T/kJ/m2

G0-G0

0.2

2.6

G70-G70

1.4

16

178

Chapter 5

5.4

Matrix-Granulate Interface

Matrix-granulate model

5.4.1 Use of energy balance approach to compute strain energy release rates

This model, which was prepared by Tavakkoli (2006), is a 2D model that uses plane
stress 4 node elements in section 3.8. The shape of the granulate was assumed to be a
sphere and thus the granulate was modelled as a sphere with a diameter of 650 microns.
Using two independent particle size technologies (i.e. mastersizer and sieving) the
granulates used for various experiments throughout this study were determined as 30
mesh size. A granulate that comes under the 30 mesh size label corresponds to an
average diameter of 650 m even though there will clearly be a particle size
distribution.

Figure 5-8 FEA models of M0-G0(47%) held at 300% strain showing the crack
propagating through the granulate (tearing) [Tavakkoli (2006)].

179

Chapter 5

Matrix-Granulate Interface

The granulate was assumed to be distributed uniformly within the matrix. A crack was
introduced in the model that was made to propagate in one of two different paths. The
first path would simulate tearing through the granulate as illustrated in Figure 5-8. The
second crack path is illustrated in Figure 5-9 which simulates separation taking place at
the matrix-granulate interface.

Figure 5-9 - FEA models of M70-G70(35%) held at 300% strain showing the crack
propagating at the matrix-granulate interface (peeling) [Tavakkoli (2006)].

The models were held at 300% strain and the cracks are shown progressing either
through the centre of the granulate or along the interface.
As discussed in chapter 4, for all models with unfilled SBR as the matrix, the Mooney
SEF was used to represent the rubber like behaviour in the FEA model. Similarly for
models where the matrix is filled SBR, the Yeoh SEF was used. Using coefficients

180

Chapter 5

Matrix-Granulate Interface

obtained from the stress-strain plots, the models behaviour was characterised. The
procedure and the curve fitting is explained in section 4.3 (chapter 4).

Peeling - Path 2
Tearing - Path 1

M0-G0(47% )

Elastic strain energy, U /mJ

2.5
2
1.5
1
0.5
0
0

0.2

0.4
0.6
Crack length, c / mm

0.8

Figure 5-10 Change in total elastic strain energy with crack length for M0-G0(47%) at
300% strain [Tavakkoli (2006)].

Figure 5-10 shows that for a model with SBR 0 as the matrix containing 47% SBR 0
granulates by volume which is held at a fixed displacement of say 300% strain, there is
a drop in the total elastic stored energy with an increase in the crack length irrespective
of the route that the crack follows. This is expected because an increase in the crack
length should always result in a release of the elastic strain energy.
As seen in Figure 5-11, a similar observation can be made for the case of the model
with SBR 70 as matrix containing 35% volume fraction of SBR 70 granulates by
volume which also is held at 300% strain.

181

Chapter 5

Matrix-Granulate Interface

Figure 5-12 was plotted using the energy balance approach on the data from Figure 5-10
and shows how the elastic strain energy for M0-G0(47%) model (held at 300% strain)
decreases with increasing crack length.

Peeling - Path 2
Tearing - Path 1

M70-G70(35% )

Elastic strain energy, U / mJ

40
35
30
25
20
15
10
5
0
0

0.2

0.4

0.6

0.8

Crack length, c / mm
Figure 5-11 - Change in total elastic strain energy with crack length for M70-G70(35%) at
300% strain [Tavakkoli (2006)].

The difference in these strain energy values (dU) for two models held at a fixed
displacement causes the crack tip area A to be extended by a small area (dA). The strain
energy release rate can thus be computed as:
Strain energy release rate =

dU
dA

eq. ( 5-1)

This is the energy balance approach which was discussed in chapter 3. For example, the
calculation of strain energy release rate due to path 2 for M0-G0(47%) at a particular
crack length was computed using the change in internal energy dU for this path over
consecutive points in Figure 5-11 and the change in crack area, dA that accompanies
this crack growth.

182

Chapter 5

Matrix-Granulate Interface

Figure 5-12 plots the result of this calculation which shows that when a crack grows
following path 1, which simulates tearing across the granulate, the strain energy release
rate (tearing energy) is higher than the case (peeling energy) where the crack grows
through the path which simulates the peeling phenomenon across the matrix-granulate
interface. The strain energy release rate data for tearing through granulate, T is only
available for crack length values of 0.65mm, the granulate diameter in the model.

M0-G0(47% )

Strain energy release rate / kJm-2

Tearing through the granulate


Peeling at the interface

2.5
2
1.5
1
0.5
0
0

0.2

0.4
0.6
0.8
Crack length, c / mm

1.2

Figure 5-12 A plot comparing the energy required for crack propagation for two
different routes in the matrix-granulate model strained at 300%. In this case the matrix is
SBR 0 and it contains 47% SBR 0 granulate by volume [Tavakkoli (2006)].

As the crack starts to grow along the matrix-granulate interface, the energy release rate
shows a small dip and then increases with increasing crack length until it reaches a
maximum which indicates the energy release is a maximum. The decline in the peeling
energy then with further increase in crack length is due to the approach of the far side of
the granulate. Figure 5-12 also shows pictures of different models at different stages of
crack growth.

183

Chapter 5

Matrix-Granulate Interface

The energy balance approach was also used for the case of M70-G70(35%) and thus
Figure 5-13 was obtained using the elastic strain energy versus crack length data shown
in Figure 5-11. The trend was similar to that for M0-G0(47%) shown in Figure 5-12.

M70-G70(35% )

Peeling at the interface

45
Strain energy release rate / kJm-2

Tearing through the granulate

40
35
30
25
20
15
10
5
0
0

0.2

0.4
0.6
0.8
Crack length, c / mm

1.2

Figure 5-13 A plot comparing the energy required for crack propagation for two
different routes in the matrix-granulate model strained at 300%. In this case the matrix is
SBR 70 and it contains 35% SBR 70 granulate by volume [Tavakkoli (2006)].

5.4.2 Crack propagation path

The strain energy release rates for each different crack propagation paths were
compared for each material type. As discussed in the previous section, the tearing
energy required to grow a crack was always higher than the respective peeling energy at
all crack lengths for a given strain. However, the weak bonding between the matrix and
the granulate means that we have to actually compare the crack growth rate versus peel
or tearing energy relations to understand which path is most likely. This is obvious as
the tear path is much shorter than the peel path and the crack growth in tearing is always

184

Chapter 5

Matrix-Granulate Interface

perpendicular to the applied force. Simplistically therefore one might expect the cracks
to always go through the granulate.
5.4.3 Relation between crack growth rate and strain energy release rate

Figure 5-14 Cut growth rate, dc/dt as a function of tearing energy, T for unfilled SBR.
() tensile test pieces, (+) pure shear test pieces and the line is according to eq. ( 5-4) [Lake
and Lindley (1964)].

185

Chapter 5

Matrix-Granulate Interface

Lake and Lindley (1964) studied the crack growth in rubber under static conditions and
the results of the tests are plotted in Figure 5-14 which is a double log plot of crack
growth rate, dc/dt against tearing energy, T. It was possible to fit the tensile test data
points on a straight line which had the following form.

T
dc
=
dt
Tu

eq. ( 5-2)

where and are material constants and in order to make T/Tu dimensionless, Tu=1.0
kJm-2 is included.
To see which path is the most favourable it was necessary to use both the measured peel
energy versus peel rate and tearing energy versus tear rate to calculate the time it would
take for a crack to propagate through both paths.
The crack growth rate

dc
is calculated during the peel or tear test from the test machine
dt

cross head separation rate, x& using the following relationship:


dc
x&
=
dt 2

eq. ( 5-3)

Tearing tests were done at various cross head speeds on strips of SBR 0 and SBR 70
that were double cured to replicate the same composition as the granulates. The peeling
tests were done on sandwich specimens of:
a) SBR 70 single cured + SBR 70 double cured
b) SBR 0 single cured + SBR 0 double cured
The tests were carried out at various cross head speeds ( x& ) so that for every cross head
speed, the corresponding tearing energy, T and peeling energy, P values were
calculated. The in equation 5-5 is the extension ratio that corresponds to the average
tearing or peeling stress for each case.

186

Chapter 5

Matrix-Granulate Interface

-2.5

y = 3x - 7.9
y = 3x - 3.3

-2.7

Log dc /dt - m/s

-2.9
y = 3x - 9.2

-3.1

y = 4.5x - 5.6

-3.3
-3.5
-3.7

SBR 0 + SBR0 - Peeling

-3.9

SBR 0 Double cured - Tearing

-4.1

SBR 70 + SBR 70 - Peeling

-4.3

SBR 70 Double cured - Tearing

-4.5
-0.5

0.5

1.5

2.5

Log (Strain energy release rate - kJ/m2)


Figure 5-15 Double log plot of crack growth rate versus strain energy release rate (i.e.
tearing or peeling) for M0-G0 and M70-G70 type materials.

Figure 5-15 shows on a double log plot of crack growth rate versus strain energy release
rate for a range of different test types. In case of peeling between M0-G0 and M70-G70,
the strain energy release rate was actually peeling energy, P. It is evident that the
peeling of the SBR 0SBR 0 double cured interface (M0-G0) has the lowest strain
energy release rate. Comparatively, the tearing of a double cured SBR 70 material
(G70) has the highest strain energy release rate which agrees with the trend shown by
the experimental findings that are tabulated in Table 5-1.
Whilst preparing the M70-G70 peel test specimens for this exercise it was seen that the
interfacial bonding between the single cured and double cured sheets was dependent on
the pressure applied during moulding. It is possible that the interface might have been
strengthened by increasing the pressure still further.
In each case, the double log plot reveals a power law trend similar to Figure 5-14. Thus
depending on the crack growth path, it should be possible to express the crack growth
rate as:

187

Chapter 5

Matrix-Granulate Interface

dc
= f (T ) for tearing through granulate and
dt
dc
= f ( P) for peeling across matrix-granulate interface
dt

eq. ( 5-4)

The coefficients and will vary according to the crack growth path and can be
obtained from the intercept and the slope of the relevant type in Figure 5-15. The
coefficients derived from the peel and tear tests are given in Table 5-4.
Table 5-4 Coefficients obtained for crack growth materials in order to establish the
power law relationships.

Type

Test type

Equation

M0-G0

Peel

5 x 10-4

dc
= 5 10 4 P 3
dt

G0

Tear

3 x 10-6

4.5

dc
= 3 10 6 T 4.5
dt

M70-G70

Peel

1 x 10-8

dc
= 1 10 8 P 3
dt

G70

Tear

7 x 10-10

dc
= 7 10 10 T 3
dt

P and T are in kJ/m2 and dc/dt is in m/s

5.4.4 Interpreting experimental and FEA data to predict crack growth time

The relationship between crack growth rate and strain energy release rate was
established for the granulate materials and the matrix-granulate interfaces of M0-G0 and
M70-G70 in the previous section. This is combined here with the strain energy release
rate versus crack length data calculated earlier by FEA to predict the time required for a
crack to grow from one point to another within these materials either along the interface
or through the granulate.
The strain energy release rate versus crack length plots shown in Figure 5-12 and Figure
5-13 show the T or P values (depending on the crack path) at specific values of c. The

188

Chapter 5

Matrix-Granulate Interface

measured crack growth release rate can be taken for each T or P value using equation
5-6 with the appropriate coefficients. It is therefore possible to generate a crack growth
rate, dc/dt versus crack length, c plot such as that shown in Figure 5-16.

SBR 0 Double cured - Tearing through G0

0.0004
0.00035

dc /dt - m/s

0.0003
0.00025
0.0002
0.00015
0.0001
0.00005
0
0

0.0001 0.0002 0.0003 0.0004 0.0005 0.0006 0.0007


c /m

Figure 5-16 Plot that shows the crack growth rate variation with crack length as the
crack propagates through a SBR 0 double cured strip (G0).

This figure is now re-plotted as Figure 5-17 this time with

dt
on the y axis (which is
dc

the reciprocal of the crack growth rate shown on the y axis in Figure 5-16) against c.
The time taken for a crack to grow from length c1 to c2 on the granulate will be the area
under this curve from c1 to c2.
For example the time taken for the crack to grow from c1 = 6.6 x 10-6 m to
c2 = 6.2 x 10-4 m is the area under the curve (in the inset) between these two values of c
which is 7.02 seconds.

189

Chapter 5

Matrix-Granulate Interface

SBR 0 Double cured - Tearing through G0

300% Strain

140000
120000

dt /dc - s/m

100000
80000
60000
40000
20000
0
0

0.0001

0.0002

0.0003

0.0004

0.0005

0.0006

0.0007

c /m

Figure 5-17 Time required for crack to propagate from one point to another across the
G0 granulate can be estimated using this plot. The integral shown being the time taken by
the crack to grow from c1 = 6.6 x 10-6 m to c2 = 6.2 x 10-4 m as 7.02 seconds.

This same approach was used to estimate the time required for crack to propagate from
one point to another along the M0-G0 interface as well.
A different data set gives the crack growth rate versus crack length plot for an
interfacial crack in M0-G0. The time taken by a crack to progress along the matrixgranulate interface can be calculated using a dt/dc versus c plot as shown in Figure 5-18.
The time taken for the crack to grow from c1 = 6.4 x 10-6 m to c2 = 4.5 x 10-4 m along
the M0-G0 interface is the area under the curve (in the inset) between these two values
of c and is estimated as 1.88 seconds.
In this manner the time required for crack to grow through G70 and along the M70-G70
interface was also predicted using Figure 5-19 and Figure 5-20 respectively. The results
are tabulated in Table 5-5.

190

Chapter 5

Matrix-Granulate Interface

SBR 0 + SBR 0 - Peeling at the M0-G0 Interface

300% Strain

14000
12000

dt /dc - s/m

10000
8000
6000
4000
2000
0
0

0.0002

0.0004 0.0006 0.0008


Crack length, c / m

0.001

0.0012

Figure 5-18 - Time required for crack to propagate from one point to another across the
M0-G0 interface can be estimated using this plot. The plot inset shows the area under the
curve from c1 to c2. The estimated time for crack to grow from c1 = 6.4 x 10-6 m to c2 = 4.5 x
10-4 m is 1.88 seconds.

The time required for complete peeling along the M0-G0 interface is 2.06 seconds and
this is lower than the 7.02 seconds required for the crack to progress diametrically as a
tear from one end of the G0 granulate to the other. Similar observation is made when
the interfacial peeling of M70-G70 is compared to tearing through G70. These cases
confirm that complete unzipping along the matrix-granulate interface is quicker and
hence more likely than crack propagation through the granulate. Clearly in each case the
interfacial peel energy is lower than the tearing energy and so it is easier for failure to
occur by separation along the matrix-granulate interface rather than due to crack
propagation through the granulate in the matrix-granulate composite.
With a knowledge of the matrix-granulate combination and the path that the crack is
likely to take, once initiated, it is possible to predict the time required for the crack to

191

Chapter 5

Matrix-Granulate Interface

propagate completely which would ultimately result in failure using the approach
outlined above.

SBR 70 Double cured - Tearing through G70

300% Strain

1200000

dt /dc - s/m

1000000
800000
600000
400000
200000
0
0

0.0002

0.0004

0.0006

0.0008

c /m
Figure 5-19 - Time required for crack to propagate from one point to another across the
G70 granulate can be estimated using this plot. The plot inset shows the area under the
curve from c1 to c2. The estimated time for crack to grow from c1 = 6.6 x 10-6 m to c2 = 6.2 x
10-4 m is 55.08 seconds.

192

Chapter 5

Matrix-Granulate Interface

SBR 70 + SBR 70 - Peeling at the M70-G70 Interface

300% Strain

120000

dt /dc - s/m

100000
80000
60000
40000
20000
0
0

0.0002

0.0004

0.0006

0.0008

0.001

Crack length, c / m
Figure 5-20 - Time required for crack to propagate from one point to another across the
M70-G70 interface can be estimated using this plot. The plot inset shows the area under
the

curve

from

c1

to

c2.

Time

taken

for

crack

to

propagate

from

c1 = 6.4 x 10-6 m to c2 = 4.5 x 10-4 m is estimated to be 15.3 seconds.

Table 5-5 Comparison of time required for crack/flaw to grow.

Type

c1 / m

c2 / m

Time taken / sec

G0

6.2 x 10-4

7.01

G70

6.2 x 10-4

55.08

M0-G0

1.0 x 10-3

2.06

M70-G70

1.0 x 10-3

15.30

193

Chapter 5

5.5

Matrix-Granulate Interface

Conclusions

A technique was developed to quantify the interfacial strength between the matrix and
granulate. Using techniques such as trouser tear and peel test it was deduced that the
interface between the matrixgranulate was weak. The interface between granulates
were studied as well and it was concluded that the interface between granulates was
weaker than that between matrix-granulate and this explained the drop in strength of the
overall composite material as the volume fraction of granulates increases.
FEA models were used to study the path along which flaws/cracks were likely to
progress, once initiated. This was done using the energy balance approach that
estimated the energy available for crack to proceed either along the matrix-granulate
interface or through the granulate. The analysis results agreed with the experimental
findings that the cracks were likely to proceed along the interface rather than through
the granulates.
Experimental peeling and tearing results were interpreted successfully along with results
obtained by analysis of 2D plane stress models to predict the time required for a
crack/flaw once initiated, to grow from one point to another, either along the matrixgranulate interface or through the granulate.

194

Chapter 6

Volume Change

6
6.1

Volume Change

Introduction

In chapter 4 it was observed that the incorporation of rubber granulates into a virgin
matrix resulted in a reduction in strength associated with an increase in the Intrinsic
Flaw Size (IFS). In chapter 5 the weak bonding between the elastomer matrix and the
granulate was evaluated. It was seen that a flaw/crack progressed along the interface at a
quicker rate compared to through the granulate. However, the precise nature of this
intrinsic flaw is not known. The focus of the work undertaken in this chapter is to
investigate the change in flaw size with strain and the reduction in strength resulting
from a weaker interface using volume change experiments.
The first set of volume change experiments under tensile strain were carried out on
carbon black filled rubber and without granulates incorporated. In Section 6.3 the
results of these experiments are discussed. In the absence of granulates no significant
volume changes were observed under strain. However, the incorporation of granulates
led to experimentally measured volume changes with strain that were substantially
greater.
Section 6.4 describes the micro-structural Finite Element Analysis (FEA) models that
were used to study the cavitation process. The findings of this study reveal how the
change in volume might result from a net increase in the flaw size with increasing
strain. The actual flaw sizes were derived using the FEA results. This approach is
discussed in section 6.5.
The work presented in this chapter has submitted for publication in Journal of Polymer
Science: Part B Polymer Physics. Please see appendix 2. A part of the work has been
accepted for publication in the proceedings of the European conference on constitutive
models for rubber (ECCMR), 2007. Please see appendix 3.
6.2

Experimental procedure for density determination

Hydrostatic weighing described in chapter 3, based on Archimedes principle, was used


to measure volume change. When a solid body is immersed in a fluid, the body will

195

Chapter 6

Volume Change

experience an upward force (buoyancy) which will be equal to the weight of the fluid
that is displaced by the solid. This upward force will therefore vary directly with the
volume of the solid body and the density of the fluid.
This buoyant force (FBuoy) can be expressed in terms of the volume of the solid (Vsolid)
and density of fluid (fluid) as:

FBuoy = Vsolid fluid g

eq. ( 6-1)

The negative sign implies that the immersed solid will experience an apparent loss of
weight due to buoyancy.
The apparatus used to carry out this study is shown in Figure 3-8 with four elastomer
rings tested simultaneously on the rig. This was suspended by a platinum wire which
was a part of the analytical balance that was placed on a vibration free table. The initial
dimensions of these rings were known and the whole experiment was carried out with
the rig and rings setup immersed in distilled water. A little quantity of detergent was
added to the water to reduce the risk of air bubbles being trapped and facilitate the
complete wetting out of the rig and the sample surfaces. The water temperature was
monitored at the time of each reading using a thermometer that was kept immersed in
the water.
The stretching (to a desired strain) and relaxing to 0% strain cycle was repeated three
times to derive a set of 3 mass readings for both the stretched and unstretched states at
each extension. The change in volume (V) at a particular extension was then calculated
from the difference in these mean weights between the unstretched and stretched states
as:
V m 1

=
VO W VO

eq. ( 6-2)

where, m is the change in mass between stretched and unstretched positions, w is the
density of water and V0 is the initial volume of rubber. So that the accuracy of these
readings were not compromised, necessary steps were taken to ensure that there was no
weight change due to water absorption and throughout the experiment the fixtures were
kept immersed until all the measurements were completed.

196

Chapter 6

Volume Change

300

M70-G0(41%)
M70-G70(35%)
M70-G70(14%)

V/V x 105

250

M70
M70-dilation

200

D ev70-0

150

Deviation by
M70-G0(41% )
from dilatation
calculated for
M70 at 300%
strain.

100
50

1
V
E
=
1
VO
3B

0
1

1.5

2.5
3
Extension ratio

3.5

4.5

Figure 6-1 - Experimental values of the fractional increase in volume observed for all the
materials with SBR 70 as matrix with various quantities of granulates incorporated
plotted against extension ratio. These are compared with the dilatational values expected
from hydrostatic stress for M70 calculated using equation 2-18.

6.3

Experimental volume change trends

The solid curve in Figure 6-1 was computed using the equation below which was
introduced earlier as equation 2-18. It takes into account the strain induced anisotropy of
rubber at high strains.

E 1
V
=
1
V
3B

eq. (2-18)

This equation was used to predict the volume change that would take place in our
materials due to dilatation effects alone under simple tension. The plot showing the
change in volume with increasing extension for all the combinations that have SBR 70
as the matrix is shown in Figure 6-1. There is a good agreement between the solid curve
(calculated using equation 2-18) and the experimental volume change data points for
M70 (the filled material without granulates) up to 200% strain. In this case, the volume
change under strain results from the dilatation behaviour alone suggesting that no
197

Chapter 6

Volume Change

cavitation occured. The introduction of granulates into the matrix causes an increase in
the volume change above this theoretical curve especially at higher strains. At more
modest strains (< 50%), the change in volume is similar for all materials irrespective of
granulate volume fraction. However, at larger strains, as the volume fraction of the
granulates increases, the volume change over and above the contribution from dilatation
alone also increases. This implies there is cavitation occurring in the granulate filled
rubber.

M0-G70(41%)
M0-G0(47%)
M0-G0(21%)
M0
M0-dilatation

1200
1000

D ev0-70

V/V x 105

800
600

Deviation by
M0-G70(41%) from
dilatation calculated
for M0 at 300% strain
(D ev0-70 ).

400
200

V
E 1
=
1
VO 3B

0
1

1.5

2.5
3
Extension ratio

3.5

4.5

Figure 6-2 - Experimental values of the fractional increase in volume observed for SBR 0
as matrix with various quantities of granulates incorporated plotted against extension
ratio. These are compared with the dilatational values expected from hydrostatic stress for
M0 calculated using equation 2-18.

Figure 6-2 shows a family of curves that compare the change in volume for materials
with SBR 0 as a matrix. The behaviour is broadly similar to the case with the SBR 70
matrix. Within the range of experimental error, there is a good agreement between the
calculated volume change for SBR 0 using equation 2-18 and the actual experimental

198

Chapter 6

Volume Change

values for M0. With the unfilled matrix though, the measured volume changes are
greater with strain.
It is clear that a mismatch in modulus between the matrix and granulate results in
greater volume changes.
6.4

Cavitation model

Figure 6-3 The crack profile is shown as the model is strained with nodes released along
the interface. The void is represented by the shape outlined by ABCO.

Micro-structural models of the cavitation process were used to estimate how large the
flaw size at the granulate-virgin rubber interface must be to create a given volumetric

199

Chapter 6

Volume Change

strain. The criterion adopted to calculate the debond at the interface is the critical work
of adhesion. The approach adopted here was to model a range of flaws over a wide
range of strains to evaluate how the volume of a cavity would vary with strain and flaw
size. It would then be possible to estimate from the experimental data of volume change
with strain, how large the average flaw size must be at a specific strain. Similar to the
models in chapter 5, it was assumed that granulates were evenly dispersed throughout
the matrix. It was also assumed for ease of modelling that granulates were spherical in
shape.
Different axisymmetric granulate models were prepared to take into account the
different volume fractions of material that was used in the volume change experiments.
The model is shown under strain in Figure 6-3 (and again in the unstrained state in
Figure 3-16 in chapter 3). The model is axisymmetric with debonding taking place
along the matrix-granulate interface at the poles of the granulate and aligned with the
strain.
The approach was to simulate a flaw opening up most readily as the vacuole formed at
the pole with the flaw size increasing at the pole of the granulate in the model. This
approach was adopted because the findings of chapter 5 and by Kumar et al. (2007)
showed that the granulate-virgin rubber interface was the most likely site for debonding
to arise. For each flaw size, the model was stretched under tension to a range of strains.
The volume of the cavity at various extension ratios was calculated as described below.
The deformed and undeformed coordinates of these nodes were obtained from
ABAQUS and the volume of the cavity was calculated using Pappus centroid theorem.
One of the elements that has nodes n with coordinates (xn,yn) and n+1 with coordinates
(xn+1,yn+1) along the curve AB is deformed and is shown in Figure 6-4. The area under
this deformed element consists of a triangle A1 and a rectangle A2 where,
A1 = 0.5(yn-yn+1)(xn+1-xn)
A2 = (xn+1-xn)(yn+1-yB)

eq. ( 6-3)

This area under the deformed element with nodes n and n+1 was rotated through an
angle 2 about the vertical symmetry axis, AO. The volume of the annulus was
calculated using Pappus theorem,

200

Chapter 6

Volume Change

x x

V1 = 2. A1 xn + n +1 n
3

x x

V2 = 2. A2 xn + n +1 n
2

eq. ( 6-4)

Total volume of the void = V1 + V2.

Figure 6-4 - A schematic showing the upper half of the crack profile in the axisymmetric
model and how the crack volume was calculated. Volume of the strip under each element
edge along the crack front was considered for this calculation.

The total volume under all the elements at a particular strain was obtained by the
summation of volumes under each element present along the crack face AB. The same
approach was adopted to calculate the volume above each element present along the
crack face BC to the horizontal line, OB, as well. Thus the volume of ABC was a sum
of the volumes under crack face AB and above crack face BC.
The models were prepared and meshed using I-DEAS v.11 and the analysis was carried
out using ABAQUS v 6.6.1. Once again the elastomer behaviour at large strains was
modelled using appropriate strain energy functions (SEF). As explained in chapter 4, for

201

Chapter 6

Volume Change

filled materials such as M70, the Yeoh SEF (equation 2-35) was appropriate over the
strain range considered here:
W = C10 ( I 1 3) + C 20 ( I 1 3) 2 + C 30 ( I 1 3) 3

eq. (2-35)

and conversely, where the matrix was unfilled, a Mooney SEF (equation 2-27) was
appropriate over the strain range considered:
W = C10 ( I 1 3) + C01 ( I 2 3)

eq. (2-27)

Experimental

250

175 m

200
(V/V ) x 10

150 m

150
100

125 m

50

100 m
75 m
50 m

0
1

1.5

2.5
3
Extension ratio

3.5

Figure 6-5 - Experimental volume change experienced by M70-G0(41%) due to cavitation


superimposed on the fractional volume change due to debonding in the FE model for the
same material combination. The values on the right in the figure above indicate the
modelled debond length values (Db).

6.5

Debonding in the model

To investigate the cavitation process under strain, several different finite element microstructural models with different intrinsic flaw sizes, Db, were modelled. Figure 6-5 is a

202

Chapter 6

Volume Change

plot of the volume change observed against the extension ratio for M70-G0(41%). Not
surprisingly increasing strain always results in an increase in the size of the void. The
specific flaw being indicated beside each curve and as expected larger flaw sizes
produce larger volume changes under strain. Also on this figure (shown by the thick
line) is an experimentally measured volume change under strain. This experimental
curve is obtained by subtracting the calculated volume change due to hydrostatic effects
alone from the experimentally measured volume change values. The value used at
extension ratio of 4 is highlighted by the arrow in Figure 6-1 as Dev70-0. In this manner
the difference is calculated for each extension ratio and the resultant curve is therefore
the volume change from cavitation effects alone.

80

102 m

70

Experimental

(V/V ) x 10

60
50
85 m

40
30
20

68 m

10

51 m
34 m

0
1

1.5

2.5

3.5

Extension ratio
Figure 6-6 - Experimental volume change experienced by M70-G70(35%) due to cavitation
superimposed on the fractional volume change due to debonding in the FE model for the
same material combination.

203

Chapter 6

Volume Change

500
120 m

450

(V/V ) x 105

400

Experimental

350

103 m

300
250
200
150

86 m

100

68 m

50

51 m
34 m

0
1

1.5

2.5
3
Extension ratio

3.5

Figure 6-7 - Experimental volume change experienced by M0-G0(47%) due to cavitation


superimposed on the fractional volume change due to debonding in the FE model for the
same material combination.

Figures 6-6 and 6-7 are similar plots for material combinations M70-G70(35%) and
M0-G0(47%) respectively. A comparison of the three plots for three different
combinations indicates that Figure 6-5 shows unusual behaviour for the specific flaw
size model for M70-G0(41%) as, the cavitation size reaches a maximum at an extension
ratio of 3 and then decreases with increasing strain.
For all three plots as shown in Figures 6-5, 6-6 and 6-7, it is seen that for a given
material, the experimental curve intersects the family of curves representing volume
change for various Db values at various points with increasing extension. These points
of intersection represent the average flaw size for a particular material type at that
specific strain. Figure 6-8 plots how this flaw size varies with strain for each
combination.

204

Chapter 6

Volume Change

Average Flaw size / m

180

M70-G0(41%)

160
140
120

M0-G0(47%)

100

M70-G70(35%)

80
60
40
20
0
1

1.5

2.5
3
Extension ratio

3.5

Figure 6-8 - Plot obtained from the intersections in figures 6-5, 6-6 and 6-7. It shows how
the intrinsic flaw size (IFS) also increase with strain for materials with a large granulate
volume fraction.

18

86 m

16

(V/V) x 105

14

Experimental

12
10

69 m

8
6
4

52 m

35 m

17 m

1.5

2.5

3.5

Extension ratio
Figure 6-9 - Experimental volume change experienced by M70-G70(14%) due to cavitation
superimposed on the fractional volume change due to debonding in the FE model

205

Chapter 6

Volume Change

30

(V/V ) x 105

25

86 m

20
Experimental

15

69 m

10
52 m

35 m
17 m

0
1

1.5

2.5
3
Extension ratio

3.5

Figure 6-10 - Experimental volume change experienced by M0-G0(21%) due to cavitation


superimposed on the fractional volume change due to debonding in the FE model.

60

(V/V ) x 105

50
138 m

40

Experimental

30

121 m

20

104 m

10

86 m
69 m
52 m
35 m

0
1

1.5

2.5
3
Extension ratio

3.5

Figure 6-11 - Experimental volume change experienced by M70-G0(17%) due to cavitation


superimposed on the fractional volume change due to debonding in the FE model.

206

Chapter 6

Volume Change

Figures 6-9, 6-10 and 6-11 show plots of volume change versus extension ratio for
different matrix-granulate combinations which are M70-G70(14%), M0-G0(21%) and
M70-G0(17%) respectively but with lower volume fraction of granulates where
experimental results were interpreted with FEA results. These show similar trends to
materials with larger volume fractions of granulate but as expected the volume changes
with strain are smaller.

140
M70-G0(17% )

Average Flaw size / m

120
100
80

M0-G0(21% )

60

M70-G70(14% )

40
20
0
1

1.5

2.5
3
Extension ratio

3.5

4.5

Figure 6-12 Plot obtained from figures 6-9, 6-10 and 6-11. It shows how the intrinsic flaw
size (IFS) also increase with strain for materials with a small granulate volume fraction.

Figure 6-12 compares average flaw sizes of all three material combinations and how
they change with extension when lower volume fractions of granulates are incorporated.
Similar to Figure 6-8, the average flaw sizes increase with extension for all three
combinations and the effect of higher volume fraction of granulates on the average flaw
size for each combination is shown by the plot in Figure 6-13.

207

Chapter 6

Volume Change

Average debonded length / m


at 300% Strain

200
SBR 70 + SBR 0

180
160
140
120

SBR 0 + SBR 0

100

SBR 70 + SBR 70

80
60
40
20
0
0

20

40

60

80

100

120

Quantity of granulates / phr


Figure 6-13 - Plot showing that addition of granulates into matrix rubber increases the
average debond length at 300% strain.

Figures 6-8, 6-12 and 6-13 reveal that for M70-G70 and M0-G0, an increase in strain
produces an increase in the average debond length. The debond lengths for these two
material types are quite similar and are much lower than the debond length values of
M70-G0.
It thus seems that with homogeneous material types where the matrix and granulate
have broadly the same modulus (M70-G70 and M0-G0) that the debonding is smaller
than that for an inhomogeneous matrix-granulate composite such as M70-G0.
With the third case, where the matrix is filled (SBR 70) and granulate is unfilled
(SBR 0), the average debond lengths are comparatively higher than their homogeneous
counterparts (Figure 6-13). The initial flaws of M70-G0(41%) and M70-G0(17%) have
also been reported as 475 and 300 m respectively in chapter 4 and by Kumar et al.
(2007). Thus it seems that such a composite wherein there is a significant mismatch in
modulus between the filled matrix and unfilled granulate, results in a significant stress

208

Chapter 6

Volume Change

gradient at the interface between the matrix and granulate which in turn results in a very
large apparent intrinsic flaw size (IFS).

Table 6-1 Comparison between IFS values [Kumar et al. (2007)] and average debond
lengths (from Figure 6-13) for the homogeneous matrix-granulate composites.

Nomenclature

Intrinsic Flaw Size


(IFS) / m
at break

Average Debonded
length / m
at 300% Strain

M0-G0(21%)

75

73

M0-G0(47%)

125

110

M70-G70(14%)

245

75

M70-G70(35%)

315

100

Table 6-1 lists the IFS values and average debond lengths for the homogeneous
matrix-granulate composites. The methodology adopted to estimate these IFS values is
outlined in chapter 4 which involved pulling the specimens to break and the value thus
obtained represents the single largest flaw in the material just prior to break that can
result in failure. Average debond lengths that have been tabulated are the values
obtained here by straining the models to 300%.
In case of M0-G0 average debond lengths and IFS values are broadly similar, which is
to be expected as the M0-G0 material type breaks at around 300% strain. However, this
is not the case with M70-G70 where the IFS is much greater than the average debond
length at 300% strain. This is also not unexpected as these samples break at much
higher strains and if the trend of increasing flaw size with strain were calculated using
this technique to the breaking strain it is likely that a similar match would result.

209

Chapter 6

Volume Change

Thus with the homogeneous matrix-granulate composite the trend between the average
debond length and IFS can be explained. With an inhomogeneous matrix-granulate
composite (Table 6-2), however, there is a discrepancy as the IFS is much larger than
the average debond lengths. This will need to be looked into in further detail.
The data and results seen in this chapter can be interpreted in the following manner:
when granulates are incorporated into the matrix, local stresses around the matrixgranulate interface tend to increase and this in turn leads to debonding at the
matrixgranulate interface. An increase in volume change is due to this debonding
which results in formation of vacuoles. When higher quantities of granulates are mixed
into the matrix, there is an increased granulate-granulate interaction which creates
additional regions of poor adhesion. This leads to further debonding and eventually
greater volume changes with strain.
The work presented in this chapter suggests that even for randomly dispersed
granulates, incorporated into a polymer matrix, there are areas with a large number of
cavitation sites as a result of the weak interface. These flaws increase in size with strain.
Modulus mismatch between the granulate and the matrix clearly results in much larger
flaws and weaker materials. This also needs to be studied in greater detail. The
technique presented in this chapter can be adopted to examine how tailoring the
interface can be pursued by a range of physical or chemical means to improve the
granulate-virgin rubber interface.
Table 6-2 - Comparison between IFS values [Kumar et al. (2007)] and average debond
lengths (from Figure 6-13) for the inhomogeneous matrix-granulate composites.

Nomenclature

Intrinsic Flaw Size


(IFS) / m
at break

Average Debonded
length / m
at 300% Strain

M70-G0(17%)

300

127

M70-G0(41%)

475

180

210

Chapter 6

6.6

Volume Change

Prediction of crack growth time

Figure 6-14 shows the variation in interfacial peel energy with increasing crack length
for M0-G0(47%). This plot was constructed using the energy balance approach which
has been discussed in chapters 2 and 5. Two observations can be made from this plot

For a given strain, the peel energy increases with increasing crack length.

Greater strains required a higher energy.

M70-G70(35%) shows a similar trend (Figure 6-15).

M0-G0(47% ) - Axisymmetric model

300% Strain
200% Strain

0.2

100% Strain

Peel energy, P / kJm-2

0.18
0.16

50% Strain

25% Strain

0.14
G

0.12
0.1
0.08
0.06
0.04
0.02
0
0

0.002

0.004

0.006

0.008

0.01

crack length, c / mm
Figure 6-14 Plot showing how the peel energy varies with growing crack along the
M0-G0 interface.

Chapter 5 shows that a power law relationship exists between the crack growth rate,

dc
dt

and the peeling energy, P for our materials. This relationship was established using
experimental data. In case of the M0-G0 interface, the relationship was found to be:
dc
= 5 10 4 P 3
dt

eq. ( 6-5)

211

Chapter 6

Volume Change

and similarly for M70-G70 interface this was established as:

dc
= 1 10 8 P 3
dt

eq. ( 6-6)

M70-G70(35% ) - Axisymmetric model

25% Strain
50% Strain
100% Strain

200% Strain
300% Strain

Peel energy, P / kJm-2

2.5
G

2
1.5
1
0.5
0
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

crack length, c / mm
Figure 6-15 - Plot showing how the peel energy varies with growing crack along the
M70-G70 interface.

As described in chapter 5, it is therefore possible to construct the plots shown in Figure


6-16 and Figure 6-17. The inset plot in Figure 6-16 shows that the time required for a
crack to propagate from 56 microns to 120 microns along the interface is the area under
the curve which is approximately 0.12 minutes. This is low compared to the 16.9
minutes required for a crack to propagate approximately the same distance along the
M70-G70 interface as shown in the inset of Figure 6-17. Thus clearly the M70-G70
interface is stronger than the M0-G0 interface and is able to delay the crack growth
more effectively. In this manner it is possible to calculate the time required for a crack
to progress from one point to another along the interface.

212

Chapter 6

Volume Change

300% Strain

M0-G0(47% ) - Axisymmetric model

dt/dc - min / mm

2000
1500
1000
500
0
0

0.02

0.04
0.06
0.08
0.1
crack length, c / mm

0.12

0.14

Figure 6-16 Plot used to compute the time required for a crack to progress from one
point to another along the interface of the M0-G0 model held at 300% strain.
300% Strain

M70-G70(35% ) - Axisymmetric model

700000

dt /dc - min/mm

600000
500000
400000
300000
200000
100000
0
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

Crack length, c / mm
Figure 6-17 - Plot used to compute the time required for a crack to progress from one to
point to another along the interface of the M70-G70 model held at 300% strain.

213

Chapter 6

Volume Change

Due to the weakening of the SBR 0 rubber when subjected to double cured cycle, it was
not possible to prepare a peel specimen of M70-G0 combination. Hence a relationship
between crack growth rate and peel energy could not be obtained.
6.7

Conclusions

Hydrostatic weighing was used to measure the volume changes that take place in
stretched samples of granulate filled virgin rubber. Deviations from expected volume
changes due to dilatation were seen that increased with volume fraction of granulate and
strain. FEA techniques were used to analyse micro-structural models of the elastomeric
materials in order to examine the origin of this deviation.
Experimental and FEA results were merged in order to carry out a systematic study of
the debonding that occurs at the matrix-granulate interface. It is clear that for the case of
modulus matched materials (with the same matrix and granulate modulus), the extent of
debonding was much less than for mismatched materials. Modulus mismatch gave rise
to higher debonding and thus a larger effective intrinsic flaw size over the entire range
of strains.
Crack growth rate versus peel energy relationships were established for the
homogeneous matrix-granulate combinations used in this study. Using these
relationships it would be possible to predict the time required for the matrix-granulate
composite material to fail due to crack/flaw propagation along the interface.

214

Chapter 7

Conclusions & Future work

7
7.1

Conclusions & Future work

Conclusions

The behaviour of rubber granulate filled elastomers has been investigated. Several
different techniques have been developed to measure the changes in behaviour, the
interfacial mechanics and the cavitation processes present as a result of the
incorporation of rubber granulates into cured rubber sheets. These techniques can now
systematically be adopted in more general studies that seek to improve the properties.
Flaw size measurements on both carbon black filled and unfilled rubber with and
without granulates incorporated into them were carried out using a technique that
involved extrapolating tensile strength, b versus cut length, c (corner cut) data for each
of the materials under consideration. The tensile strength values when plotted against
the respective precut lengths were found to decrease with increasing cut length. With
the flaw sizes which were actually corner cut equivalent flaw sizes, the trend seen was,
that the addition of rubber granulates into rubber matrix caused an increase in the flaw
size. The change in size of these flaws gives an insight into the failure initiation of
rubber matrix-rubber granulate composites.
A test to investigate interfacial strengths between matrix and granulates and between
granulates was developed. From this it was concluded that without suitable surface
modification of the granulate materials, the interface was weak. The composition of the
matrix rubber and the granulate rubber (that was incorporated into the matrix) was
known. It was therefore possible to prepare a flat interface between two sheets where
one had the same composition as the matrix rubber and the other with the same
composition as the granulate rubber. The comparison of the energy required to separate
these sheets was compared with the energy required to propagate a tear across the sheet
with the composition of the granulate rubber. This simple yet informative technique not
only showed that the interface between the matrix and granulate was weak but also gave
an indication of the extent to which it was weaker. FEA techniques confirmed the
experimental observation that a flaw once initiated was more likely to follow a path
along the matrix-granulate interface rather than through the granulate. Using FEA it was

215

Chapter 7

Conclusions & Future work

also possible to estimate the time taken for the composite material to fail
catastrophically and the times were found to vary depending on whether the crack
progressed along the interface or through the granulate.
Using these techniques not only could we conclude that the incorporated granulates
increased the size of the defects that was pre-existent in the matrix but also that the
interface between the matrix and granulate was a likely site for defects to arise and was
a suitable path for these defects or flaws to grow.
This study showed that the interface between granulates was important because as the
volume fraction of granulates increases, the incidence of intergranulate contact would
also increase. The strength of this interface was measured using the same technique
described above. It showed that the granulate-granulate interface was much weaker than
the matrix-granulate interface. This implied that with such situations, in addition to the
poor interfacial strength of matrix-granulate, additional weakening would result due to
the poor bonding between granulates that would ultimately result in further reduction of
the strength properties of the overall composite. This therefore provides insight into the
reason for the reduction in strength properties of matrix-granulate composite materials.
The work on cavitation showed that matrix-granulate composite materials when strained
exhibited significant changes in volume when compared to materials without granulates
incorporated. This indicated that cavitation was occurring in the composite materials.
The experimental and FEA results, when combined, indicated that the flaw size
increases in a characteristic way with strain for materials where the matrix and
granulates have a similar modulus, whereas a modulus mismatch between the matrix
and the recycled granulate results in much larger volume changes and hence greater
flaw size which also appears to increase with strain.
Thus the final chapter of this work emphasised the importance in practical applications
of matching the modulus of recycled granulate materials to that of the new virgin
material in the matrix. A novel technique for examining, small changes in the interfacial
bonding mechanisms under strain was also outlined.
It was also seen that the incorporation of a greater volume fraction of granulates led to
further debonding at the matrix-granulate interface. This basically corroborated earlier
findings that had shown that with granulates coming in closer proximity to one another,

216

Chapter 7

Conclusions & Future work

the interface between granulates became important and there was a reduction in the
strength of the composite.
Based on the findings of this study it becomes clear that before granulates can be
successfully and commercially reincorporated it will be essential to improve the
adhesion between the matrix and granulates as well as between the granulates by either
increasing the effective surface area or by reducing the granulate size or by surface
modification and chemical activation of granulates.
Characterisation of rubber granulates was also undertaken in the initial stages of this
study and was quite challenging. Techniques such as FTIR, TGA, pycnometry and
particle size analysis were used to obtain relevant information about both the rubber
granulates used in this study as well as rubber granulates secured from industrial
sources. These techniques were used to determine the size and density of rubber
granulates, the type of base polymer as well as to quantify the volume fraction of the
polymer present in the granulate. In case of granulates that contained elastomer blends,
it was possible to identify each of the different elastomer materials as well.
7.2

Future work

The present study looked at the crack initiation and propagation mechanism. All
polymers contain intrinsic flaws and it is now clear that matrix-granulate composite
materials are weak because of the weak interface along which it is easy for a flaw to
propagate. Also, the greater the difference in the modulus between the matrix rubber
and granulate rubber, the weaker is the overall composite. The results of this study
therefore make it easier to understand the reason for drop in strength of rubber sheets
that have recycled rubber granulates embedded in them.
However, there is definitely scope for further work which can take the study further and
also probably tackle some of the issues that were not looked into in this study. Some of
the key aspects on which further research can be based are:

This study was entirely based on non crystallising SBR. Although the
comparison of results between filled and unfilled matrix-granulate combinations
made this an interesting study, it will be useful to examine the flaw size values
and interfacial strengths in case of NR materials and with blends of NR and SBR

217

Chapter 7

Conclusions & Future work

as well as other matrix-granulate combinations. The volume change aspect of


this study would be complicated for NR compounds due to the strain induced
crystallisation effects.

In this study it was seen that there was a drop in the tensile strength of the
material when granulates were added into the matrix. It will be interesting to see
how the fatigue strength and wear properties vary with the addition of rubber
granulates into the matrix rubber. For practical industrial applications fatigue
and wear data are essential.

Granulates used in this study were approximately of a maximum size of 650 m


in diameter. Future work could involve studying the effects of incorporating
granulates of different size ranges varying from very fine granulates of 100 mesh
(~ 150 m) to coarser ones 10 mesh (2000 m).

The techniques developed in this work can be used to evaluate the effectiveness
of surface modification techniques. The latter can be done and then the effect of
modification on flaw size, interfacial bonding can be examined. Perhaps it may
then be possible to incorporate greater quantities of granulates. Eventually all
rubber products will have to be recycled so very high volume fractions are
desired to use up all the old rubber stock.

As well as by increasing the strength of the interface it is possible to modify the


interfacial mechanics by changing the surface area to volume ratio of the
granulates. This can most easily be done by reducing the size of the granulates.
Therefore a detailed study is now required on how granulate size alters the
mechanical properties such as strength at equivalent volume fractions.

To be of commercial value it is vital that granulates derived from tyres be of a


known composition. This can be done using detailed FTIR investigations,
chemical methods etc which was not possible during the timeframe of this study.

FEA was used at various stages in this study. However, it will be useful to use
detailed FEA to especially examine the interface.

218

References

References
Adhikari, B., De, D. and Maiti, S., Reclamation and Recycling of Waste Rubber.
Progress in Polymer Science (Oxford), 25, 909-948, 2000.
Agullo, N. and Borros, S., Qualitative and Quantitative Determination of the Polymer
Content in Rubber Formulations. Journal of Thermal Analysis and Calorimetry, 67,
513-522, 2002.
Ahmed, R. and Klundert, A., Waste Rubber Recycling ,WEDC Conference ,Conference
Proceeding , (Colombo, Sri Lanka), 1994.
Ball, J. M., Discontinuous Equilibrium Solutions and Cavitation in Non-Linear
Elasticity. Philos.Trans.R.Soc.London.Ser.A., 306, 557-611, 1982.
Bauman, B. D., High-Value Engineering Materials From Scrap Rubber. Rubber World,
212, 1995.
Berry, J. P., Fracture Processes in Polymeric Materials. II. The Tensile Strength of
Polystyrene. Journal of Polymer Science, 50, 313-321, 1961.
Bhowmick, A. K., Fractographic Investigation of Tensile and Fatigue Fracture Surfaces
of Rubber Containing Inhomogeneous Inclusions. Journal of Materials Science Letters,
5, 1042-1044, 1986.
Blow, C. M., Rubber Technology and Manufacture; 1971, Butterworth & Co Ltd,
London, 1 ed.
Brown, D. A. and Watson, W. F., Tech Service: Novel Concepts in Environmentally
Friendly Recycling. Rubber World, 225, 28-32, 2001.
Burford, R. P. and Pittolo, M., Characterization and Performance of Powdered Rubber.
Rubber Chemistry and Technology, 55, 1233-1249, 1982.

219

References

Burford, R. P. and Pittolo, M., Fracture Morphology of Rubber Compounds Containing


Recycled Rubber Powder. Journal of Materials Science Letters, 2, 422-424, 1983.
Burford, R. P. and Pittolo, M., The Mechanical Properties of Rubber Compounds
Containing Soft Fillers. Part 1 Tensile Properties and Fracture Morphology. Journal of
Materials Science, 19, 3059-3067, 1984.
Burgoyne, M. D., Leaker, G. R. and Krekic, Z., Effect of Reusing Ground Flash and
Scrap Rubber in Parent Compound. Rubber Chemistry and Technology, 49, 375-378,
1976.
Busfield, J. J. C., The Prediction of the Mechanical Performance of Elastomeric
Components using Finite Element Analysis, 2000, Ph.D. Thesis, Queen Mary,
University of London .
Busfield, J. J. C., Davies, C. K. L. and Thomas, A. G., Aspects of Fracture in Rubber
Components. Progress in Rubber and Plastics Technology, 12, 191-207, 1996.
Busfield, J. J. C., Jha, V., Liang, H., Papadopoulos, I. C. and Thomas, A. G., Prediction
of Fatigue Crack Growth Using Finite Element Analysis Techniques Applied to ThreeDimensional Elastomeric Components. Plastics, Rubbers and Composites, 34, 349-356,
2005.
Busfield, J. J. C., Thomas, A. G. and Ngah, M. F., Application of Fracture Mechanics
for the Fatigue Life Prediction of Carbon Black Filled Elastomers ,Conference
ProceedingDorfmann, A. and Muhr, A., 1999.
Chang, R. J. and Gent, A. N., Effect of Interfacial Bonding on the Strength of Adhesion
of Elastomers - 1. Self-Adhesion. Journal of Polymer Science: Polymer Physics Edition,
19, 1619-1633, 1981.
Chatterley, T., Impact of the End of Life Vehicles Directive on the Tyre Industry.
MIRA New Technology, 2004.

220

References

Choi, I. S. and Roland, C. M., Intrinsic Defects and the Failure Properties of Cis-1,4Polyisoprenes. Rubber Chemistry and Technology, 69, 591, 1996.
Christensen, R. G. and Hoeve, C. A. J., Comparison Between Theoretical and
Experimental Values of the Volume Changes Accompanying Rubber Extension. J
Polym Sci Part A-1 Polym Chem, 8, 1503-1512, 1970.
De, S. K., Isayev.A.I. and Khait, K., Rubber Recycling; 2005, CRC Press Taylor &
Francis Group,
Department of Trade and Industry, Tyre Recycling. 2001.
Dierkes, W., Solutions to the Rubber Waste Problem Incorporating the Use of Recycled
Rubber. Rubber World, 214, 25-31, 1996.
Evans, A. and Evans, R., Waste Tyre Materials: Definitions & General Uses by Size.
2006.
Fagan, M. J., Finite Element Analysis : Theory and Practice; 1992, Harlow : Prentice
Hall, 1992,
Fedors, R. F. and Landel, R. F., Dependence of Relative Volume on Strain for an SBR
Vulcanizate. Rubber Chemistry and Technology, 43, 887-896, 1970.
Fernandez-Berridi, M. J., Gonzalez, N., Mugica, A. and Bernicot, C., Pyrolysis-FTIR
and TGA Techniques As Tools in the Characterization of Blends of Natural Rubber and
SBR. Thermochimica Acta, 444, 65-70, 2006.
Fesus, E. M. and Eggleton, R. W., Recycling Rubber Products Sensibly. Rubber World,
203, 23-26, 1991.
Flory, P. J., Thermodynamic Relations for High Elastic Materials. Trans.Faraday Soc,
57, 829, 1961.

221

References

Fukahori, Y. and Seki, W., Molecular Behaviour of Elastomeric Materials Under Large
Deformation: 1. Re-Evaluation of the Mooney-Rivlin Plot. Polymer, 33, 502, 1992.
Gee, G., The Interaction Between Rubber and Liquids. X. Some New Experimental
Tests of a Statistical Thermodynamic Theory of Rubber-Liquid Systems. Trans.Faraday
Soc, 42, B033-B044, 1946.
Gee, G., Studies in Rubber Elasticity: 2. Volume and Energy Changes in the Elongation
of Elastomers. Polymer, 28, 386-392, 1987.
Gee, G., Stern, J. and Treloar, L. R. G., Volume Changes in the Stretching of
Vulcanized Natural Rubber. Trans.Faraday Soc, 46, 1101-1106, 1950.
Gent, A. N. Strength of Elastomers, Science and Technology of Rubber, Eirich, F. R.,
1978, New York, Academic Press, ed., 419-454
Gent, A. N., Detachment of an Elastic Matrix From a Rigid Spherical Inclusion. Journal
of Materials Science, 15, 2884-2888, 1980.
Gent, A. N. and Hwang, Y. C., Internal Failures in Model Elastomeric Composites.
Journal of Materials Science, 25, 4981-4986, 1990.
Gent, A. N. and Lindley, P. B., Internal Rupture of Bonded Rubber Cylinders in
Tension. Proc.R.Soc.London Ser.A, 249, 195-205, 1958.
Gent, A. N., Lindley, P. B. and Thomas, A. G., Cut Growth and Fatigue of Rubbers. I.
The Relationship Between Cut Growth and Fatigue. Journal of Applied Polymer
Science, 8, 455-466, 1964.
Gent, A. N. and Park, B., Failure Processes in Elastomers at or Near a Rigid Spherical
Inclusion. Journal of Materials Science, 19, 1947-1956, 1984.

222

References

Gent, A. N., Campion, R. P. and American Chemical Society.Division of Rubber,


Engineering With Rubber : How to Design Rubber Components; 1992, Munich : Hanser
Publishers,
Gibala, D. and Hamed, G. R., Cure and Mechanical-Behavior of Rubber Compounds
Containing Ground Vulcanizates. Part I - Cure Behavior. Rubber Chemistry and
Technology, 67, 636-648, 1994.
Gibala, D., Hamed, G. R. and Zhao, J., Tensile Behavior of an SBR Vulcanizate
Containing a Single Rubber Particle. Rubber Chemistry and Technology, 71, 861-865,
1998.
Gibala, D., Laohapisitpanich, K., Thomas, D. and Hamed, G. R., Cure and Mechanical
Behavior of Rubber Compounds Containing Ground Vulcanizates. Part II - Mooney
Viscosity. Rubber Chemistry and Technology, 69, 115-119, 1996.
Gibala, D., Thomas, D. and Hamed, G. R., Cure and Mechanical Behavior of Rubber
Compounds Containing Ground Vulcanizates: Part III. Tensile and Tear Strength.
Rubber Chemistry and Technology, 72, 357-360, 1999.
Gough, J., Muhr, A. H. and Thomas, A. G., Material Characterisation for Finite Element
Analysis of Rubber Components. Journal of Rubber Research, 1, 222, 1998.
Greensmith, H. W., Rupture of Rubber X. The Change in Stored Energy on Making a
Small Cut in a Test Piece Held in Simple Extension. Journal of Applied Polymer
Science, 7, 993-1002, 1963.
Gregory, M. J., The Stress/Strain Behaviour of Filled Rubbers at Moderate Strains.
Plastics and Rubber: Materials and Applications, 4, 184-188, 1979.
Griffith, A. A., The Phenomena of Rupture and Flow in Solids. Philosophical
Transactions of the Royal Society of London, Series A, 221, 163-198, 1921.

223

References

Hamed, G. R., Effect of Crosslinking Density on the Critical Flaw Size of a Simple
Elastomer. Rubber Chemistry and Technology, 56, 244-251, 1983.
Han, S. C. and Han, M. H., Fracture Behavior of NR and SBR Vulcanizates Filled With
Ground Rubber Having Uniform Particle Size. Journal of Applied Polymer Science, 85,
2491-2500, 2002.
Hilyard, N. C., Tong, S. G. and Harrison, K., Influence of the Cure System on the
Properties of Vulcanizates Incorporating Whole Tyre Scrap Rubber Crumb. Plastics and
Rubber Processing and Applications, 3, 315-322, 1983.
Jacob, C., Bhowmick, A. K., De, P. P. and De, S. K., Utilization of Powdered EPDM
Scrap in EPDM Compound. Rubber Chemistry and Technology, 76, 36-59, 2003.
James, A. G. and Green, A., Strain Energy Functions of Rubber II. The Characterisation
of Filled Vulcanizates. Journal of Applied Polymer Science, 19, 2319, 1975.
Jones, H. C. and Yiengst, H. A., Dilatometer Studies of Pigment-Rubber Systems.
Industrial & Engineering Chemistry, 32, 1354-1359, 1940.
Khasanovich, T. N., Note on the Article by Hewitt and Anthony: ``Measurement of the
Isothermal Volume Dilation Accompanying the Unilateral Extension of Rubber''.
Journal of Applied Physics, 30, 948-949, 1959.
Kim, J. K., Morphological Study of Deformation of Acrylonitrile/Butadiene Rubber
Loaded With Crumb Rubber During Stretching. Plastics, Rubber and Composites
Processing and Applications, 27, 362-368, 1998.
Kim, J. K., Fracture Behavior of Crumb Rubber-Filled Elastomers. Journal of Applied
Polymer Science, 74, 3137-3144, 1999.
Kim, J. K. and Burford, R. P., Study on Powder Utilization of Waste Tires As a Filler in
Rubber Compounding. Rubber Chemistry and Technology, 71, 1028-1041, 1998.

224

References

Klingensmith, W. and Baranwal, K., Recycling of Rubber: an Overview. Rubber World,


218, 41-46, 1998.
Kohler, R. and O'Neill, J., New Technologies for the Devulcanisation of Sulfur-Cured
Scrap Elastomers ,Conference Proceeding , (Louisville,KY), 1996.
Kumar, P., Fukahori, Y., Thomas, A. G. and Busfield, J. J. C., Recycled Rubber: The
Rubber Granulate - Virgin Rubber Interface. Rubber Chemistry and Technology, 80 ,
24-39, 2007.
Kuznetsova, O. P., Zhorina, L. A. and Prut, E. V., Blends Based on Ground Tire
Rubber. Polymer Science,Ser.A., 46, 151-159, 2004.
Lake, G. J. and Lindley, P. B., Cut Growth and Fatigue of Rubbers. II. Experiments on a
Noncrystallising Rubber. Journal of Applied Polymer Science, 8, 707-721, 1964.
Mathew, G., Singh, R. P., Lakshminarayanan, R. and Thomas, S., Use of Natural
Rubber Prophylactics Waste As a Potential Filler in Styrene-Butadiene Rubber
Compounds. Journal of Applied Polymer Science, 61, 2035-2050, 1996.
McCrum, N. G., Buckley, C. P. and Bucknall, C. B., Principles of Polymer Engineering;
1988, Oxford : Oxford University Press, 1988, International student ed.
Mooney, M., A Theory of Large Elastic Deformation. Journal of Applied Physics, 11,
582-592, 1940.
Mullins, L. and Tobin, N. R., Carbon-Black Loaded Rubber Vulcanisates: Volume
Changes in Stretching. Trans.Inst.Rubber Ind., 33, 2-10, 1957.
Myhre, M. and Mackillop, D. A., Rubber Recycling. Rubber Chemistry and
Technology, 75, 429-474, 2002.
O'Keefe, J. F., Identification of Polymers by IR Spectroscopy. Rubber World, 230, 2732, 2004.

225

References

Obata, Y., Kawabata, S. and Kawai, H., Mechanical Properties of Natural Rubber
Vulcanizates in Finite Deformation. J Polym Sci Part A-2 Polym Phys, 8, 903-919,
1970.
Oberth, A. E. and Bruenner, R. S., Tear Phenomena Around Solid Inclusions in
Castable Elastomers. Journal of Rheology, 9, 165-185, 1965.
Ogden, R. W., Volume Changes Associated With the Deformation of Rubber-Like
Solids. J.Mech.Phys.Solids, 24, 323-338, 1976.
Payne, E., Reclaim Rubber Usage and Trends. Rubber World, 210, 4, 1994.
Penn, R. W., Volume Changes Accompanying the Extension of Rubber.
Trans.Soc.Rheol., 14, 509-517, 1970.
Phadke, A. A., Chakraborty, S. K. and De, S. K., Cryoground Rubber-Natural Rubber
Blends. Rubber Chemistry and Technology, 57, 19-33, 1984.
Pittolo, M. and Burford, R. P., The Mechanical Properties of Rubber Compounds
Containing Soft Fillers - Part 2 Tear Properties. Journal of Materials Science, 19, 33303336, 1984.
Rawle.A., Basic Principles of Particle Size Analysis. Malvern Instruments - Technical
Paper, 1-8, 2006.
Reichert, W. F., Hopfenmueller, M. K. and Goeritz, D., Volume Change and Gas
Transport at Uniaxial Deformation of Filled Natural Rubber. Journal of Materials
Science, 22, 3470-3476, 1987.
Riggle, D., A Finer Grind for Rubber Recyclers. Biocycle, 36, 42-&, 1995.
Rivlin, R. S. Large Elastic Deformations, Rheology: Theory and Applications, Eirich, F.
R., 1956, New York, Academic Press, ed.,

226

References

Rivlin, R. S. and Thomas, A. G., Rupture of Rubber. I. Characteristic Energy for


Tearing. Journal of Polymer Science, 10, 291-318, 1953.
Roland, C. M., The Barrier Performance of Latex Rubber. Rubber World, 208, 15,
1993.
Schippel, H. F., Volume Increase of Compounded Rubber Under Strain. Industrial &
Engineering Chemistry, 12, 33-37, 1920.
Sekhar, N. and Van Der Hoff, B. M. E., Cavity Formation on Elongation in Filled
Elastomers. Journal of Applied Polymer Science, 15, 169-182, 1971.
Shulman, V. L., Tyre Recycling. Rapra review reports, expert overviews covering the
science and technology of rubber and plastics, 7, 2004.
Slater, S., UK Used Tyre Market 2004. 2006.
Stringfellow, R. and Abeyaratne, R., Cavitation in an Elastomer: Comparison of Theory
With Experiment. Materials Science & Engineering A: Structural Materials: Properties,
Microstructure and Processing, A112, 127-131, 1989.
Sunthonpagasit, N. and Duffey, M. R., Scrap Tires to Crumb Rubber: Feasibility
Analysis for Processing Facilities. Resources, Conservation and Recycling, 40, 281299, 2004.
Swor, R. A., Jensen, L. W. and Budzol, M., Ultrafine Recycled Rubber. Rubber
Chemistry and Technology, 53, 1215-1225, 1980.
Tavakkoli, B., Interfacial Mechanics of Recycled Granulate Rubber Incorporated into
Virgin Rubber, 2006, MSc. Thesis, Queen Mary, University of London.
Thomas, A. G., Fracture of Rubber ,Physical basis of Yield and Fracture ,Conference
Proceeding , (Oxford), 1966.

227

References

Treloar, L. R. G., The Physics of Rubber Elasticity; 1975, Oxford : Clarendon Press, 3rd
ed.
Treloar, L. R. G., Dilation of Rubber on Extension. Polymer, 19, 1414-1420, 1978.
Waddell, W. H. and Parker, J. R., Direct Characterization of Tire Materials by
Photoacoustical Infrared and Proton Induced X-Ray Emission Spectroscopy. Rubber
Chemistry and Technology, 65, 836-850, 1992.
Warner, W. C., Methods of Devulcanization. Rubber Chemistry and Technology, 67,
559-566, 1994.
White, J. R., De, S. K., Rubber Technologist's Handbook; 2001, Rapra Technology
Limited, Shrewsbury, 1 ed.
Williams, P. T. and Besler, S., Pyrolysis-Thermogravimetric Analysis of Tyres and Tyre
Components. Fuel, 74, 1277-1283, 1995.
Yeoh, O. H., Characterization of Elastic Properties of Carbon-Black-Filled Rubber
Vulcanizates. Rubber Chemistry and Technology, 63, 792-805, 1990.
Yeoh, O. H. and Fleming, P. D., New Attempt to Reconcile the Statistical and
Phenomenological Theories of Rubber Elasticity. Journal of Polymer Science, Part B:
Polymer Physics, 35, 1919-1931, 1997.

228

Appendix1

Appendix 1
Refereed journal papers published by the author as part of this research

RECYCLED RUBBER: THE RUBBER GRANULATE - VIRGIN


RUBBER INTERFACE
P. Kumar, Y. Fukahori, A.G. Thomas and J.J.C. Busfield1
Department of Materials, Queen Mary, University of London, Mile End Road, London, E1 4NS, UK.

Corresponding author, Tel: +(44)20-7882-5570; e-mail: j.busfield@qmul.ac.uk

ABSTRACT
Reusing granulates derived from old tyre stock and other sources in high tech
engineering applications is still considered a high risk option. In addition to ensuring
that the granulates are correctly identified, it is important to know how the incorporation
of these materials alters the intrinsic flaw size of a finished product and to see how
much the strength of the interface between these materials and the virgin materials
compares to the basic strength of the virgin stock and the granulates. This paper
explores possible techniques that can examine both properties so that an informed
evaluation of the effect of reincorporating granulates can be established in practice.

Published in: Rubber Chemistry and Technology, vol. 80, issue 1, 2007, p 24-39

229

Appendix2

Appendix 2
Refereed journal papers published by the author as part of this research

VOLUME CHANGES UNDER STRAIN RESULTING FROM THE


INCORPORATION OF RUBBER GRANULATES INTO A RUBBER
MATRIX

P. Kumar, Y. Fukahori, A.G. Thomas and J.J.C. Busfield

Department of Materials, Queen Mary, University of London, Mile End Road, London, E1 4NS, UK

ABSTRACT
The strength of an elastomer is in part determined by the size of the intrinsic flaws that
are present. It has been observed that the incorporation of rubber granulates into a virgin
matrix results in a reduction is strength and this has previously been attributed by
Kumar et al.(2007) to an increase in the Intrinsic Flaw Size (IFS). The precise nature of
this intrinsic flaw is the subject of this investigation. Fundamental questions concerning
the change in flaw size with strain and the reduction in strength resulting from a weaker
interface have been investigated using volume change experiments. Initial experiments
on carbon black filled rubber with no granulates incorporated, have shown no
significant volume change under strain. This contrasts with granulate filled materials,
whose experimentally measured volume changes with strain were seen to be
substantially greater. Micro-structural Finite Element Analysis (FEA) has revealed how
this change in volume might result from a net increase in the flaw size with increasing
strain.
This work suggests that flaw size increases in a characteristic way with strain for
materials where the matrix and granulates have a similar modulus, whereas a modulus

230

Appendix2

mismatch between the matrix and the recycled granulate results in much larger volume
changes and hence greater flaw size which also appears to increase with strain. This
work emphasises the importance in practical applications of matching the modulus of
recycled granulate materials to that of the new virgin material in the matrix. This paper
introduces a novel technique for examining, small changes in the interfacial bonding
mechanisms under strain such as that caused by surface modification techniques.

Keywords: Rubber recycling, Debonding, Flaws, Interface, Finite Element Analysis

Submitted April 2007: Journal of Polymer Science Part B: Polymer Physics

Presented at: Tire Tech Expo, 13-15 March 2007, Kln Messe, Germany.

231

Appendix3

Appendix 3
Refereed conference proceedings presented and published by the author as part of
this research

CAVITATION IN GRANULATE FILLED RUBBER MATERIALS


P. Kumar, J.J.C. Busfield, Y. Fukahori and A.G. Thomas
Department of Materials, Queen Mary, University of London, Mile End Road, London, E1 4NS, UK

ABSTRACT
All elastomers contain flaws within them and these intrinsic flaws affect the strength of
the material. Rubber granulates when incorporated into virgin rubber results in
premature failure of the resulting material and it has previously been attributed to an
increase in the Intrinsic Flaw Size (IFS). The question remains, is it just an increase in
the flaw size or do these flaws open up when the matrix-granulate composite is strained
and is the failure associated to the poor bond between the matrix and granulate? Our
experiments on volume change show that for carbon black filled rubber with no
granulates incorporated, no significant volume change under tensile strain due to
cavitation is observed. In contrast, for granulate filled materials, volume changes are
measured experimentally and have also been modelled using Finite Element Analysis
(FEA). This has allowed us to deduce how the flaw size changes as a function of strain.
This work shows that flaw size increases similarly with strain for all the composite
materials where the matrix and granulates have same modulus whereas a modulus
mismatch between the matrix and the recycled granulate results in much larger volume
changes and hence greater flaw size which appear to also increase with strain. This
work emphasises the importance of matching the modulus of recycled granulate
materials to that of the new matrix.
To be published in: Proceedings of the Fifth European Conference on Constitutive
Models for Rubber, 4-7 September 2007. Ecole des Mines de Paris.

232

Appendix4

Appendix 4
Refereed conference proceedings presented and published by the author as part of
this research

INVESTIGATING THE BEHAVIOR AT THE RUBBER


GRANULATE-VIRGIN RUBBER INTERFACE

P. Kumar, J.J.C. Busfield and A.G. Thomas


Department of Materials, Queen Mary, University of London, Mile End Road, London, E1 4NS, UK

ABSTRACT
Reusing granulates derived from old tyre stock and other sources in high tech
engineering applications is still considered a high risk option. In addition to ensuring
that the granulates are correctly identified, it is important to know how the
reincorporation of these materials alters the intrinsic flaw size of a finished product and
to see how much the strength of the interface between these materials and the virgin
materials compares to the basic strength of the virgin stock and the granulates. This
paper explores possible techniques that can examine both properties so that an informed
evaluation of the effect of reincorporating granulates can be established in practice.

Published in: Tire Technology International, 22-24 February 2005, Kln Messe,
Germany, p. 74.

233

S-ar putea să vă placă și