Elastic And Mechanical Properties Of Expanded Perlite And ...

226
Elastic And Mechanical Properties Of Expanded Perlite And Perlite/Epoxy Foams Haleh Allameh Haery MIEngDes (UPM) This thesis is submitted for the degree of Doctor of Philosophy (Mechanical Engineering) University of Newcastle, Australia Faculty of Engineering and Built Environment School of Engineering February 2017

Transcript of Elastic And Mechanical Properties Of Expanded Perlite And ...

Page 1: Elastic And Mechanical Properties Of Expanded Perlite And ...

Elastic And Mechanical Properties

Of Expanded Perlite And

Perlite/Epoxy Foams

Haleh Allameh Haery

MIEngDes (UPM)

This thesis is submitted for the degree of

Doctor of Philosophy (Mechanical Engineering)

University of Newcastle, Australia

Faculty of Engineering and Built Environment

School of Engineering

February 2017

Page 2: Elastic And Mechanical Properties Of Expanded Perlite And ...

Statement of Originality

This is to certify that the thesis entitled “Elastic and Mechanical Properties of Expanded

perlite and Perlite/Epoxy foams” submitted by Miss. Haleh Allameh Haery contains no

material which has been accepted for the award of any other degree or diploma in any

university or other tertiary institution and, to the best of my knowledge and belief,

contains no material previously published or written by another person, except where due

reference has been made in the text. I give consent to this copy of my thesis when

deposited in the University Library, being made available for loan and photocopying

subject to the provisions of the Copyright Act 1968.

Haleh Allameh Haery

Page 3: Elastic And Mechanical Properties Of Expanded Perlite And ...

Statement of Collaboration

I hereby certify that the work embodied in this thesis has been done in collaboration with

other researchers. I have included as part of the thesis a statement clearly outlining the

extent of collaboration, with whom and under what auspices.

Page 4: Elastic And Mechanical Properties Of Expanded Perlite And ...

Statement of Authorship

I hereby certify that the work embodied in this thesis contains published papers of which

I am a joint author. I have included as part of the thesis a written statement, endorsed by

my supervisor, attesting to my contribution to the joint publications.

Haleh Allameh Haery

Page 5: Elastic And Mechanical Properties Of Expanded Perlite And ...

Dedication

I dedicate this thesis to my parents, who have always loved and supported me

unconditionally.

Page 6: Elastic And Mechanical Properties Of Expanded Perlite And ...

Acknowledgments

I would like to express my sincere gratitude to my supervisors, Professor Erich Kisi and

Doctor Thomas Fiedler, for their help and support throughout my period of PhD study. I

would like to take this opportunity to express my appreciation to my supervisor Erich,

whose patience, kind words, complete support and professional advice have been a source

of encouragement and a motivation to explore my ideas to their fullest depth. I would also

like to thank Doctor Jubert Pineda for his mentorship and constructive advice which has

helped me enormously.

I would like to take this opportunity to express my appreciation to the University of

Newcastle for granting me IPRS and Postgraduate Research scholarships during the four

years of my study.

I would also like to gratefully acknowledge the assistance from the staff in the electron

microscope and X-ray unit at the University of Newcastle, Australia.

Last, not the least, I want to thank my parents for their love, support and words of

encouragement throughout studies.

Page 7: Elastic And Mechanical Properties Of Expanded Perlite And ...

List of Publications and Awards

Awards

1. 2016 Three Minutes Thesis Finalist (Faculty of Engineering and Built

Environment); (University of Newcastle, Australia).

2. Runner Up - 2016 Three Minutes Thesis Faculty competition (Faculty of

Engineering and Built Environment); (University of Newcastle,

Australia).

3. Faculty of Engineering and Built Environment 2015 postgraduate research

prize, Mechanical Engineering (University of Newcastle, Australia).

4. Faculty of Engineering and Built Environment 2014 postgraduate research

prize, Mechanical Engineering (University of Newcastle, Australia).

5. Runner Up - 2015 Three Minutes Thesis Faculty competition (Faculty of

Engineering and Built Environment); (University of Newcastle,

Australia).

6. 2015 Three Minutes Thesis Finalist (Faculty of Engineering and Built

Environment); (University of Newcastle, Australia).

Conference (Poster)

Haleh Allameh-Haery, Erich Kisi, Thomas Fiedler, “Characterization of perlite-epoxy

foam cores for sandwich panels”, International Conference on Advances in Functional

Materials", 29 June-3 July, 2015, Long Island, NY, USA.

Page 8: Elastic And Mechanical Properties Of Expanded Perlite And ...

Conference Papers

H. A. Haery, and H. S. Kim, "Damage of hybrid composite laminates." SPIE Proceedings,

Volume 8793, Aerospace Composites, 87931F-87931F-13 (2013)

Journal Papers

1. H. Allameh Haery, H. S. Kim, R. Zahari, E. Amini, Tensile strength of notched

carbon/glass/epoxy hybrid composite laminates before and after fatigue loading.

Journal of Industrial Textiles, 2014, 44(2): p. 307 – 331.

2. H. Allameh Haery, and H. S. Kim, Damage in hybrid composite laminates.

Journal of Multifunctional Composites, 2013, 1(2): p.127-137.

3. H. Allameh Haery, E. Kisi, and T. Fiedler, Novel cellular perlite-epoxy foams:

effect of density on mechanical properties. Journal of Cellular Plastics, Doi:

10.1177/0021955x16652110, 2016.

4. H. Allameh-Haery, C.M. Wensrich, T. Fiedler, and E. Kisi, Novel Cellular

perlite-epoxy foams: effects of particle size. Journal of Cellular Plastics, Doi:

10.1177/0021955X16670528, 2016.

5. H. Allameh-Haerya, E. Kisi, J. Pineda, L. P. Suwal, T. Fiedler, Characterization

and prediction of Elastic properties of green compacts of Expanded perlite

particles. Journal of Powder Technology, 25 January 2017.

6. H. Allameh-Haery, E. Kisi, J. Pineda, L. P. Suwal, T. Fiedler, Elastic properties

of perlite-epoxy foams. (Under review).

Page 9: Elastic And Mechanical Properties Of Expanded Perlite And ...

Table of Contents

Statement of Originality .................................................................................................... 2

Statement of Collaboration ............................................................................................... 3

Statement of Authorship ................................................................................................... 4

Dedication ......................................................................................................................... 5

Acknowledgments ............................................................................................................. 6

List of Publications and Awards ....................................................................................... 7

Conference (Poster) ........................................................................................................... 7

Conference Papers ............................................................................................................. 8

Table of Contents .............................................................................................................. 9

List of Figures ................................................................................................................. 13

List of Tables................................................................................................................... 20

List of Symbols and Abbreviations .................................................................................... I

Abstract ............................................................................................................................. 1

1 Chapter One: Introduction......................................................................................... 4

2 Chapter Two: Literature Review ............................................................................... 9

Introduction ........................................................................................................ 9

The characteristics, manufacturing processes and applications of Perlite ....... 10

2.2.1 Introduction to perlite ................................................................................ 10

2.2.2 Expansion process ..................................................................................... 12

2.2.3 The physical properties of perlite .............................................................. 16

2.2.4 Chemical composition ............................................................................... 19

2.2.5 Applications .............................................................................................. 24

Page 10: Elastic And Mechanical Properties Of Expanded Perlite And ...

2.2.6 Concluding Remarks ................................................................................. 41

Foams ............................................................................................................... 42

2.3.1 Introduction to foams and syntactic foams ............................................... 42

2.3.2 Manufacturing syntactic foams ................................................................. 46

2.3.3 Particulate composites containing naturally occurring fillers ................... 50

Motivation and Problem Statement .................................................................. 59

Research Objectives and Research Significance .............................................. 60

3 Chapter Three: Methodology .................................................................................. 61

Introduction ...................................................................................................... 61

Material ............................................................................................................ 62

3.2.1 Expanded perlite particles ......................................................................... 62

3.2.2 Epoxy resin ............................................................................................... 63

Sample preparation ........................................................................................... 64

3.3.1 Preparation of expanded perlite (EP) particle samples ............................. 64

3.3.2 Preparation of solid perlite samples .......................................................... 67

3.3.3 Preparation of resin samples ..................................................................... 67

3.3.4 Preparation of EP/epoxy foam samples .................................................... 70

Experimental Setup and Tests .......................................................................... 73

3.4.1 Mechanical testing on the packed beds of EP Particles ............................ 73

3.4.2 Mechanical testing of EP/epoxy foams ..................................................... 73

3.4.3 Elastic wave tests on EP particles ............................................................. 75

3.4.4 Elastic wave tests on EP/epoxy foams and solid perlite ........................... 79

Microstructural analysis ................................................................................... 81

Damage observation ......................................................................................... 82

Page 11: Elastic And Mechanical Properties Of Expanded Perlite And ...

The theory of dynamic moduli measurement ................................................... 83

4 Chapter Four: Results .............................................................................................. 86

Introduction ...................................................................................................... 86

Elastic properties of sintered solid perlite ........................................................ 88

Properties of packed beds of EP particles ........................................................ 91

4.3.1 Structural characterisation of EP particles ................................................ 91

4.3.2 Measurement of elastic moduli of packed EP particle beds using quasi-static

mechanical tests ...................................................................................................... 96

4.3.3 Measurement of elastic moduli of packed EP particle beds using elastic

waves 98

Mathematical models for prediction for Elastic properties of porous bodies 107

4.4.1 Phani Models ........................................................................................... 109

4.4.2 Nielson Model ......................................................................................... 113

4.4.3 Minimum solid area (MSA) models ....................................................... 117

4.4.4 Gibson and Ashby Model ....................................................................... 121

Properties of epoxy resin ................................................................................ 124

Properties of EP/epoxy foams ........................................................................ 126

4.6.1 Microstructural characterisation of EP/epoxy foams .............................. 126

4.6.2 Volume fraction of the epoxy in EP/epoxy foams .................................. 128

4.6.3 Compressive response of EP/epoxy foams ............................................. 133

Damage analysis ............................................................................................. 141

Measurement of the elastic moduli of EP/epoxy foams using elastic wave speed

147

5 Chapter Five: Discussion ...................................................................................... 159

Page 12: Elastic And Mechanical Properties Of Expanded Perlite And ...

Introduction .................................................................................................... 159

Manufacturing method ................................................................................... 160

Young’s modulus of packed EP particles ....................................................... 163

Structural characterisation of EP/epoxy foams .............................................. 167

Compressive behaviour and compressive properties ..................................... 172

Damage mechanisms under compressive loading .......................................... 178

6 Chapter Six: Summary .......................................................................................... 182

Conclusions .................................................................................................... 182

Future Research .............................................................................................. 186

References ..................................................................................................................... 189

Page 13: Elastic And Mechanical Properties Of Expanded Perlite And ...

List of Figures

Figure 1.1. Schematic diagram illustrating the structure and composition of syntactic

foams. ................................................................................................................................ 4

Figure 2.1. (a) Unexpanded perlite (b) Expanded perlite. Both types were supplied by

Ausperl. ........................................................................................................................... 11

Figure 2.2. Horizontal expander and handling system [30]. ........................................... 15

Figure 2.3. Vertical expander and handling system [30]. ............................................... 15

Figure 2.4. Cellular structure formed internally in expanded perlite particles. .............. 17

Figure 2.5. US expanded perlite use by application....................................................... 24

Figure 2.6. Spherical gas tank farm in a petroleum refinery [116]. ................................ 32

Figure 2.7. SEM image showing EP microspheres consisting of one or more microcellular

bubbles [27]. .................................................................................................................... 37

Figure 2.8. (a) Closed cell Polyurethane foam [102], (b) Open cell Polyurethane foam [3].

......................................................................................................................................... 43

Figure 3.1. Durometer (Type D) for hardness test. ........................................................ 63

Figure 3.2. Tapping device for measuring tapped density of EP particles. .................... 65

Page 14: Elastic And Mechanical Properties Of Expanded Perlite And ...

Figure 3.3. Prepared mould (a) The different parts of the mould (b) Assembled mould.

......................................................................................................................................... 66

Figure 3.4. Solid perlite sample (sintered perlite powders). ........................................... 67

Figure 3.5. Mould for manufacturing epoxy resin samples for (a) compression and

flexural tests and (b) tensile tests. ................................................................................... 68

Figure 3.6. A hand-made device for separating expanded particles from unexpanded and

semi-expanded particles. ................................................................................................. 71

Figure 3.7. Formation of two phases by buoyancy. The top phase is packed perlite

particles and diluted binder and the bottom phase contained only diluted binder. ......... 72

Figure 3.8. Samples prepared using three particle size ranges; from the left, 1 - 2 mm, 2 -

2.8 mm and 2.8 – 4 mm. ................................................................................................. 72

Figure 3.9. Schematic representation of the experimental set-up for measuring wave

velocity in packed beds of EP particles........................................................................... 75

Figure 3.10. Examples of (a) P-wave and (b) S-wave signals in low density (0.12 g/cm3)

EP particle compacts and (c) P-wave and (d) S-waves in high density (0.3 g/cm3) EP

particle compacts. ............................................................................................................ 78

Figure 3.11. Schematic representation of the ultrasonic experimental set-up for measuring

wave velocity in solid perlite and EP/epoxy samples. .................................................... 79

Page 15: Elastic And Mechanical Properties Of Expanded Perlite And ...

Figure 4.1. Microscope images taken from the polished surface of sintered perlite. The

polishing marks are designated on the picture in order not to be confused with porosity.

......................................................................................................................................... 88

Figure 4.2. SEM images showing the external structure of an EP particle: (a) in the 1 - 2

mm size range; (b) in the 2 - 2.8 mm size range; (c) in the 2.8 - 4 mm size range. ........ 92

Figure 4.3. SEM images showing the internal structure of an EP particle in the: (a) 1 - 2

mm size range; (b) 2 - 2.8 mm size range; (c) 2.8 - 4 mm size range; (d) 2 - 2.8 mm size

range, illustrating natural reinforcement inside EP particle cells. .................................. 93

Figure 4.4. Constrained modulus of packed EP particle beds as a function of compact

density. ............................................................................................................................ 97

Figure 4.5. The (a) Compression wave and (b) Shear wave velocities versus compact

density. ............................................................................................................................ 99

Figure 4.6. Formation of platy and fine particles as a result of the brittle crushing of cell

walls. ............................................................................................................................. 100

Figure 4.7. (a) Percentage of debris at each EP compact density. (b) Cumulative particle

size distribution for entire compacts (EP particles and debris). .................................... 101

Figure 4.8. (a) Particle density versus particle size. This graph also includes debris density

versus debris size; (b) Inter-particle porosity (excluding debris) versus compact density;

(c) Inter-particle space filled by debris versus compact density. .................................. 105

Page 16: Elastic And Mechanical Properties Of Expanded Perlite And ...

Figure 4.9. (a) Young’s modulus (E) of packed beds of EP particles versus compact

density. (b) Poisson’s ratio of packed beds of EP particles versus compact density. In both

graphs, the upper and right hand scales allow the normalised moduli versus porosity to

be read from the same graphs........................................................................................ 106

Figure 4.10. (a) Normalised Young’s modulus versus porosity (based on experimental

moduli); (b) Normalised Poisson’s ratio versus porosity (based on experimental moduli);

(c) Normalised Young’s modulus versus porosity (based on modified moduli); (d)

Normalised Poisson’s ratio versus porosity (based on modified moduli). ................... 112

Figure 4.11. (a) Shape factor versus compact porosity for the Phani and Neilson models

applied to both the experimental moduli and modified moduli; (b) Modified Phani and

Nielson models for Young’s Moduli as a function of porosity; (c) Modified Phani and

Nielson models for Poisson’s ratio as a function of porosity........................................ 116

Figure 4.12. Hardness versus time curve used to determine the curing period of the diluted

epoxy. ............................................................................................................................ 125

Figure 4.13. Changes in the viscosity of epoxy + hardener diluted with acetone versus

acetone content. ............................................................................................................. 125

Figure 4.14. Micrographs taken from the cross-sections of EP/epoxy foams with a density

of 0.15 g/cm3 made with EP particles in the size ranges: a) 1 - 2mm; b) 2 - 2.8mm; and c)

2.8 - 4mm. ..................................................................................................................... 127

Page 17: Elastic And Mechanical Properties Of Expanded Perlite And ...

Figure 4.15. (a) Foam density as a function of applied pressure. (b) Particle density as a

function of compaction pressure. .................................................................................. 130

Figure 4.16. Volume fraction of epoxy versus density of the foam. ............................. 131

Figure 4.17. Volume fraction of epoxy binder in EP/epoxy foams of type 2. .............. 133

Figure 4.18. Typical stress-strain curves for the different EP/epoxy foam densities of: (a)

Type 1; (b) Type 2; (c) Type 3. ..................................................................................... 135

Figure 4.19. Properties of manufactured foams of type 1 ( ⃣ ), type 2 (◇) and type 3 (∆):

(a) Maximum stress versus foam density; (b) Effective modulus versus foam density; (c)

Modulus of toughness versus foam density. ................................................................. 137

Figure 4.20. Confined modulus of EP particles ( ⃣ ) and confined modulus of the foams

(◇) as a function of the density of the packed bed of EP particles. ............................. 138

Figure 4.21. Predicted elastic moduli of EP particles using the Voigt and Reuss models.

To quantify the particles’ contributions to the stiffness of the foam, the effective elastic

modulus of the EP/epoxy foams as a function of foam density is presented. ............... 140

Figure 4.22. Schematic representation of failure in EP/epoxy foams. .......................... 141

Figure 4.23. Macroscopic images showing (a) A sample of type 1 compressed to a strain

of 12% (b) A typical remnant of the samples after the test. .......................................... 145

Page 18: Elastic And Mechanical Properties Of Expanded Perlite And ...

Figure 4.24. SEM images showing (a) Uncrushed EP particles in the wedge-like fractured

side of a failed sample; (b) Cell walls of EP perlite particles fractured along a plane; (c)

Crushed cells in the central region of a uniformly compressed sample. ....................... 146

Figure 4.25. Typical stress-strain curve for different EP/epoxy foam densities. .......... 147

Figure 4.26. The longitudinal wave and shear wave velocities versus foam density. .. 149

Figure 4.27. Young’s modulus of EP/epoxy foams versus foam density determined from

the elastic wave speed (○) and mechanical tests (∆); and Young’s modulus of the EP

particles versus foam density determined from the elastic wave speed (◇). ............... 149

Figure 4.28. Volume fraction of the epoxy binder in the second series of EP/epoxy foams

versus foam density. ...................................................................................................... 151

Figure 4.29. (a) stress-strain curves and (b) Young’s modulus measured from the gradient

of the unloading path in Figure 4.29 (a) at different stress level in cyclic compressive tests

conducted on a sample with density of 0.26 g/cm3. ...................................................... 155

Figure 4.30. Poisson’s ratio of epoxy resin (­­­­), foam (○) and packed beds of EP

particles ( ⃣ ) versus the foam density. .......................................................................... 157

Figure 5.1. Young’s moduli of the packed EP particle beds measured using mechanical

tests and estimated using the Voigt and Reuss models. ................................................ 165

Page 19: Elastic And Mechanical Properties Of Expanded Perlite And ...

Figure 5.2. (a) Volume fraction of EP particles and (b) Porosity within inter-particle space

(by consideration of debris) and total porosity in type 2 EP/epoxy foams. .................. 169

Figure 5.3. A schematic representation of the internal structure of (a) low density and (b)

high density EP/epoxy foams. ....................................................................................... 171

Figure 5.4. Schematic representation of compressive stress-strain curves for a) an

elastomeric, b) an elastic-plastic, c) a brittle foam [11]. ............................................... 172

Figure 5.5. Schematic representation of the compressive stress-strain curve for EP/epoxy

foams. ............................................................................................................................ 175

Figure 5.6. Compressive strength (a) and compressive modulus (b) plotted against density

for currently available foams (Ashby et al., 2000) and results obtained for perlite-based

foams ( )....................................................................................................................... 177

Page 20: Elastic And Mechanical Properties Of Expanded Perlite And ...

List of Tables

Table 2.1. Typical physical properties of expanded perlite [23, 26, 39]......................... 18

Table 2.2.Chemical composition of perlite (Percent) [22]. ............................................. 20

Table 2.3. Mineral phases detected in perlites of different origins. ................................ 23

Table 2.4. Examples of perlitic sound insulation boards in which starch is used as binder

[69]. ................................................................................................................................. 30

Table 3.1. Chemical composition [198] .......................................................................... 62

Table 3.2. ASTM D-695, ASTM D-638 and ASTM D-790 for measuring the

compressive, tensile and flexural properties of epoxy resin. .......................................... 68

Table 4.1. Elastic properties of solid component of EP particles. For comparison, the

corresponding data for obsidian by Manghnani et al. [213] are presented. ................... 90

Table 4.2. Bulk density and particle density measurements of EP particles ................... 91

Table 4.3. Cell dimensions of EP particles of 1 - 2 mm, 2 - 2.8mm and 2.8 - 4 mm size

range. Each value is an average of 150 measurements. .................................................. 95

Table 4.4. Exponents of Eq. (4.3) and Eq. (4.4) for the three particle size ranges. ........ 96

Page 21: Elastic And Mechanical Properties Of Expanded Perlite And ...

Table 4.5. Physical parameters of mathematical models applied to the experimental and

modified moduli. The physical parameters of the modified moduli are given in

parenthesis. .................................................................................................................... 111

Table 4.6. Mechanical properties of cured epoxy resin with and without dilution by

acetone........................................................................................................................... 124

Table 4.7. Coefficients in Equations (4.21) and (4.22) for foam density Df and particle

density Dp . .................................................................................................................... 131

Table 4.8. Elastic properties of EP particles, epoxy resin and EP/epoxy foams........... 150

Page 22: Elastic And Mechanical Properties Of Expanded Perlite And ...

I

List of Symbols and Abbreviations

Item Description

EP Expanded perlite

Cp Compression wave velocity

Cs Shear wave velocity

CL Longitudinal wave velocity

𝜈 Poisson’s ratio

𝜈𝑐𝑟 Critical Poisson’s ratio

𝜈0 Poisson’s ratio at zero porosity

E Young’s modulus

E0 Young’s modulus at zero porosity

E* Constrained elastic modulus

G Shear modulus

G0 Shear modulus at zero porosity

K Bulk modulus

σcomp Compaction pressure

𝜌 Density

𝜌0 Density at zero porosity

ρc Composite density

ρp Particle density

ρg Density of the glass

ρtrue True density of the microspheres

ρe Epoxy density

Df foam density as a function of compaction pressure

Page 23: Elastic And Mechanical Properties Of Expanded Perlite And ...

II

DP Particle density as a function of compaction pressure

tw Wall thickness of particle

r Average radius of the microspheres

C Cell size

dp EP particle diameter

𝜌𝐸𝑃 EP particle density

𝜌𝑈𝑃 Density of fully dense perlite

𝜑 Angle of internal friction

a Packing geometry dependent parameter

P Porosity

Pcr Critical porosity

tapP Porosity of tapped packing state

greenP porosities of as-poured packing state

0n Constant related to pore morphology

En Pore structure dependent parameter (shape factor)

𝛽 Pore structure dependent parameter (shape factor)

S Shape factor as a function of porosity

CSA Composite sphere assemblage

b Constant related to particle stacking

𝑏′ Constant related to particle stacking

M Young’s, shear or bulk modulus

C Constant of proportionality

C Constant of proportionality

Ø Volume fraction of solid in the cell- struts

Page 24: Elastic And Mechanical Properties Of Expanded Perlite And ...

III

UpperPE Upper bound of Young’s modulus of EP particle

LowerPE Lower bound of Young’s modulus of EP particle

σhyd Hydrostatic stress

σdev Deviatoric stress

δ Kronecker delta

λ, µ Lamé constants

Page 25: Elastic And Mechanical Properties Of Expanded Perlite And ...

1

Abstract

Syntactic foams are composite materials made by reinforcing a resinous matrix with

hollow particles called microspheres. These materials are often used as the core material

for sandwich panels, where a combination of low density, high compressive strength, high

compressive deformation and high damage tolerance are required, e.g. in the aerospace,

automotive and marine industries. Syntactic foams have superior mechanical and thermal

properties but they are more expensive and denser than conventionally gas-blown foams.

The higher density and cost of syntactic foams is mainly due to the density and cost of

the synthetically made microspheres. This problem can be mitigated by using light-weight

naturally occurring particles which provide syntactic foams with similar properties but

which are significantly cheaper. In this study, the potential of expanded perlite particles

(EP particles) in manufacturing light-weight syntactic foams is investigated. Perlite is a

glassy volcanic rock of silicic composition and in its expanded form has a high porosity

(>95%), low density (~ 0.18 g/cm3) and offers excellent thermal and acoustical insulating

properties, chemical inertness, physical resilience, fire resistance and water retention

properties. These features, along with the fact that it is abundant and cheap make it a

suitable candidate for manufacturing syntactic foams.

In this study, the structural, microstructural, physical and mechanical properties of EP

particles were investigated. The elastic properties of packed EP particle beds were

characterised by the isotropic elastic moduli Poisson’s ratio and Young’s modulus,

calculated from elastic wave speeds along the axial (compaction) direction for a wide

range of compaction densities. It was observed that during compaction to achieve

different densities, some crushing of particles into smaller particles and platy debris

Page 26: Elastic And Mechanical Properties Of Expanded Perlite And ...

2

occurred. Consequently, analyses were conducted based on both the raw compaction

densities and densities modified by the removal of debris from consideration, on the

assumption that debris is non-structural. Based on the raw compaction densities, Young’s

moduli of the packed EP particles were found to vary in the range 31.4 - 371.3 MPa, while

based on the modified densities, they varied in the range 31.4 - 152.8 MPa for the compact

densities ranging from 0.1 to 0.375 g/cm3. Poisson’s ratio of packed EP particles did not

show a large variation with compact density in the range 0.1 - 375 g/cm3; Poisson’s ratio

was about 0.3. The equation for Poisson’s ratio is independent of density, hence the values

obtained based on the experimental and modified densities resulted in the same Poisson’s

ratios.

Four analytical models were applied to predict the elastic moduli of packed beds of EP

particles within the porosity range 84 - 95%. Models were assessed on their ability to

successfully predict the elastic moduli of these highly porous bodies from the properties

of solid perlite for both cases: using the raw compaction density and the modified density.

It was found that the Wang (Minimum Solid Area) model was able to estimate Young’s

modulus, while the Gibson and Ashby model was reasonable for the average behaviour

of both of the elastic moduli. The best agreement, however, was obtained through the

Phani model utilising our modified shape factor.

The impact of the research was broadened by dispersing EP particles in a matrix of epoxy

resin and the manufacture of light-weight EP/epoxy foams. Foams were fabricated with

three distinct particle size ranges and, within each size range, the samples covered a

density range 0.15 - 0.45 g/cm3. The effects of particle size and foam density variations

on the compressive strength, effective elastic modulus and modulus of toughness of the

EP/epoxy foams were investigated. The compressive properties of the EP/epoxy foams

Page 27: Elastic And Mechanical Properties Of Expanded Perlite And ...

3

showed a strong dependence on the foam density, but were almost independent of the

particle size. The compressive strength of the EP/epoxy foams was found to vary linearly

in the range 0.15 - 1.77 MPa with the foam densities ranging from 0.15 to 0.45 g/cm3.

However, the Young’s modulus and modulus of toughness of the EP/epoxy foams varied

parabolically in the ranges 33 - 227 MPa and 0.01 - 0.35 MPa, respectively. The elastic

properties of the EP/epoxy foams were characterised by adopting an isotropic model for

the medium and measuring the elastic wave speeds (i.e. longitudinal and shear wave) in

the axial direction (similar to the measurements of the particle beds). Quasi-static

compressive test results were compared with those obtained by the elastic wave tests.

Both were observed to follow the same qualitative pattern, however the Young’s moduli

measured using elastic waves were more than twice those obtained from the mechanical

tests. Poisson’s ratio showed an increasing trend, ranging from 0.17 to 0.34 over the foam

density range, and appeared to be influenced by the increase in contact surface area

between the particles and the matrix as the foam density increased.

Post-test scanning electron microscopy (SEM), coupled with photogrammetry during the

tests, were used to understand the behaviour of the foams under compressive load. The

observations revealed the presence of three different failure modes for all of the foams,

regardless of their particle size and density, however the strain to activate each mode was

different for each foam type. In addition, the observations showed that the formation of

wedge-like fragments in the foam samples under applied compressive stress were due to

the effects of pure shear (i.e. the deviatoric components of the applied stress). However,

the compressive deformation due to the effect of the hydrostatic components of the

applied stress was concentrated in the middle of the foam samples.

Page 28: Elastic And Mechanical Properties Of Expanded Perlite And ...

4

1 Chapter One: Introduction

Foams are cellular solids where cells with solid ligaments and membranes pack together

in three dimensions to fill space. These materials are widely found in nature, and have

been in use for more than 5000 years, for example, the wooden artefacts found in Egypt’s

pyramids or cork which has been used for the soles of shoes since Roman times [1]. These

materials inspired engineers to synthesise light-weight, yet strong and stiff, cellular

materials for various structural applications. Syntactic foams are a special class of cellular

materials which are made by embedding hollow microspheres in a polymeric matrix or

binder. They are identified as foams due to their cellular structure created by void space

in both the polymer resin matrix and the hollow microspheres. The schematic structure

of syntactic foams is presented in Figure 1.1.

Figure 1.1. Schematic diagram illustrating the structure and composition of syntactic foams.

Syntactic foams were initially developed in the 1960s as a buoyancy aid material for deep

submergence applications where high hydrostatic pressures are involved [2]. They are

generally used as the core material in sandwich plates and shells as an effective weight-

saving design option for various structural applications [3], and in areas where a

Page 29: Elastic And Mechanical Properties Of Expanded Perlite And ...

5

combination of low density, high compressive strength, high compressive deformation

and low moisture absorption are required, e.g. in naval, aeronautical, aerospace, civil and

automotive applications [4-6].

There are a wide variety of hollow microspheres available for use in syntactic foams.

They are commonly made from glass, ceramic, carbon or any polymeric materials. The

appropriate choice of microspheres can produce resinous foams with superior strength,

stiffness, damage tolerance and chemical resistance. Preference is generally given to

hollow glass microspheres due to their low density (~ 0.28 - 0.7 g/cm3 [7-9]), mechanical

strength, stiffness, the regularity of the surface and their price, which is lower than other

microspheres [8]. However, compared with conventionally gas-blown foams, syntactic

foams have higher density and cost of production. This is mainly due to the cost of

synthetically made hollow spheres. Cost and density are two main factors in the selection

of materials, especially when production of weight-sensitive structural components in

large quantities is involved. Hence, it would be worthwhile to find a light-weight material

which has the potential to provide syntactic foams with similar specific properties, but

which is significantly cheaper than other commercially available microspheres. One

promising material is expanded perlite particles. Perlite is a glassy volcanic rock of

rhyolitic composition normally comprised of 71-75% SiO2, 12.5-18% Al2O3, 4-5% K2O,

1-4% sodium and calcium oxides, lesser amounts of several metal oxides and 2-5% water

by weight [10]. Upon rapid controlled heating in the range 760 - 1100°C, the combined

water in perlite grains is vaporised, producing pressure that causes expansion of the perlite

to 4 - 20 times its original volume. The expanded form of perlite has a low density (~ 0.18

g/cm3 [11]) and offers excellent thermal and acoustic insulating properties, chemical

inertness, physical resilience, fire resistance and water retention properties [12]. In

Page 30: Elastic And Mechanical Properties Of Expanded Perlite And ...

6

addition, expanded perlite particles have a cellular structure mainly consisting of closed

cells (sealed bubbles) which provide them with progressive crushing characteristics as

opposed to the one-step crushing of microspheres [13]. This can be beneficial to

preserving the cellular structure of the perlite-reinforced plastic/foam for a longer service

life.

The present study focuses on the development of a light-weight foam material made by

dispersing expanded perlite particles (hence called EP particles) in a matrix of epoxy

resin. The structural, microstructural, and mechanical properties of EP particles and

EP/epoxy foams are investigated. Consequently, the potential of EP/epoxy foams as a

novel class of engineering materials, especially for light-weight structural applications,

are evaluated.

The layout of the remainder of the thesis, a significant portion of which is based on articles

published by the author [14-16] while undertaking this study, are organised as follows:

Chapter 2 delivers an overview of the existing literature related to this work in two main

sections. The first section gives a comprehensive overview of the physical properties,

chemical properties, expansion process and applications of perlite particles. The second

section outlines an overview of foams, syntactic foams and polymeric composites

reinforced with different naturally occurring minerals, including perlite particles.

Subsequently, based on gaps in the current literature, the motivation for conducting the

current project and the key objectives are highlighted.

Chapter 3 includes a description of the various raw materials used in manufacturing

EP/epoxy foams, fabrication techniques and testing procedures (i.e. quasi-static

mechanical tests and elastic wave tests) used for characterisation of the mechanical and

Page 31: Elastic And Mechanical Properties Of Expanded Perlite And ...

7

physical properties of solid perlite (i.e. virtually pore free), packed beds of expanded

perlite particles, and the novel EP/epoxy foams manufactured.

Chapter 4, firstly, outlines the experimental results related to tests performed on solid

perlite, e.g. microscopy and elastic wave tests. Secondly, the experimental results relating

to microscopy and mechanical and elastic wave tests, on packed beds of EP particles are

discussed. Thirdly, the experimental results relating to mechanical and elastic properties

of cured epoxy resin are presented. Fourthly, four mathematical models for the prediction

of elastic properties of packed EP particle beds are introduced. The predictive ability of

the models is evaluated and modifications are made to adapt to the morphology of the

packed EP particles. Fifthly, the behaviour of the foams under quasi-static compressive

loading, together with the consequences of different parameters (i.e. particle size and

foam density) on the response of the foams, are investigated. Finally, the results of elastic

wave tests on EP/epoxy foams are presented and a comparison with those obtained by the

quasi-static tests is made.

Chapter 5 discusses the significance of the results in a broad context, including an

evaluation of the advantages and disadvantages of the manufacturing method used in this

study and possible alternative methods. Additional factors are discussed, such as: i) the

Young’s modulus of packed EP particle beds measured using the different methods; ii)

the interior structure of EP/epoxy foams based on the experimental results; iii) the

compressive behaviour and properties of EP/epoxy foams; and iv) the damage

mechanisms during compressive tests. Moreover, a comparison is made with other

commercially available foams, on the basis of their compressive properties, compressive

behaviour and damage mechanisms.

Page 32: Elastic And Mechanical Properties Of Expanded Perlite And ...

8

Chapter 6 presents conclusions which may be drawn from the study. It also provides

several suggestions for future work, which may be taken as a guide by students and

researchers entering this area or wishing to pursue improved EP/epoxy foam manufacture

and/or its properties.

Page 33: Elastic And Mechanical Properties Of Expanded Perlite And ...

9

2 Chapter Two: Literature Review

Introduction

This chapter is divided into four parts. The first part introduces the perlite material and

provides comprehensive information about the process of expansion, its physical and

chemical properties, as well as current commercial uses of perlite in three areas:

construction, horticulture and industry. In the second part of this chapter, a review is

conducted on previously developed polymer matrix foams and particulate composites

which contained naturally occurring fillers, including perlite particles. The final two parts

provide the motivation and main objectives of the current project based on gaps in the

literature presented in the first and second parts.

Page 34: Elastic And Mechanical Properties Of Expanded Perlite And ...

10

The characteristics, manufacturing processes and applications of

Perlite

2.2.1 Introduction to perlite

Perlite is a glassy volcanic rock of silicic or rhyolithic composition, typically formed by

the hydration of obsidian. Perlite occurs naturally in the form of block type-lava domes,

restricted in area, which are formed by the extrusion of highly viscous magmas. Perlite

deposits are restricted to Tertiary age, or younger, deposits of rhyolithic composition.

Rhyolithic glasses are unstable and devitrify1 with age into microcrystalline aggregates

of quartz and feldspar, and transform to zeolites and other aluminosilicate minerals.

Therefore, the preservation of perlite is rarely found in rocks older than Tertiary age [17].

The name perlite is a derivation from the German word perlstein and is originally given

to “certain glassy rocks (hyaloliparites, hyalo-rhyolites) characterized with numerous

concentric cracks”, the ‘perl’, referring either to the resemblance of the broken-out

fragments to pearls or to the pearly lustre of the surfaces [18]. Petrologically, the term

‘perlite’ refers to volcanic glasses in which cooling strain resulted in a concentric or onion

structure of fracturing which may be visible to the naked eye or may only be observed

under microscopes. This structure of fracturing is also known as perlitic [19].

Perlite is distinguished from other natural water-bearing glasses (e.g. obsidian, pumice

and pitchstone) by the total water content of 2 to 5 wt% held within the glass structure,

by the presence of a pearly lustre and by its onion skin perlitic fractures. The term,

1. The process in which a glass (noncrystalline or vitreous solid) transforms to a crystalline solid.

Page 35: Elastic And Mechanical Properties Of Expanded Perlite And ...

11

however, should be expanded to include the less dense textures of perlite which are related

to the classical variety but rarely, or only microscopically, exhibit perlitic fractures, and

which might not exhibit a pearly lustre. Textures of perlite may be commonly found in

deposits ranging from the dense and classically fractured variety to vesiculated and

pumiceous varieties which may not exhibit a pearly lustre or megascopic perlitic fracture

[20]. Commercially, the term ‘perlite’ has been applied to any naturally glass of igneous

origin that will expand or ‘pop’ when heated quickly, forming a light-weight frothy

material. However, this property is also present in some glassy rocks that do not have

perlitic structure, and in some that are not primary in origin but are alteration products of

rhyolite or obsidian [21].

Upon heating within its softening range of 760°C to 1100°C [10], the combined water in

perlite grains is converted to steam and the perlite expands 4 - 20 times its original

volume. Hence, it is called ‘expanded perlite’ (EP). Figure 2.1 illustrates the unexpanded

and expanded forms of perlite.

(a) (b)

Figure 2.1. (a) Unexpanded perlite (b) Expanded perlite. Both types were supplied by Ausperl.

Page 36: Elastic And Mechanical Properties Of Expanded Perlite And ...

12

Expanded perlite has low thermal conductivity, high sound absorption, high resistance to

heat, chemical inertness, physical resilience and water retention ability. More than half of

the perlite produced is consumed in the construction industry as aggregate in insulation

boards, acoustical ceiling tiles, plaster and concrete [17]. It is also used in horticulture

and other applications like filtration, e.g. in the pharmaceutical and food industries, and

as fillers in various processes and materials [22]. There have been many studies about the

use of perlite in particular industrial and construction applications (e.g. [23], [24], and

[25]). However, few articles have been published which review perlite’s properties and

its applications. Singh [26] conducted a review on the multifarious uses of perlite as a

construction material. Barker and Santini [27] provided an overview of the world perlite

deposits and activities. Therefore, this section of the chapter aims to provide a

comprehensive overview on the characteristics and properties of perlite, as well as its

applications. Hopefully, the information provided has the potential to rapidly fill the void

in our understanding of perlite.

2.2.2 Expansion process

In general, the perlite expansion process involves rapid heating in a kiln to a softening

point, and then rapidly cooling. The size to which the rock is crushed before heating

depends on the composition of the perlite (e.g. water content), furnace design and the end-

use for which the expanded perlite is required. Generally, perlite particles that pass 8-

mesh (2.380 mm) and are retained on 30-mesh (0.595 mm) are suitable to be used as

concrete aggregate; perlite particles that pass 14-mesh (1.410 mm) and are retained on

50-mesh (0.297 mm) are suitable to be used as plaster aggregate; and fine particles that

pass 30-mesh (0.595 mm) are used for the lightest types of expanded product [28]. A

Page 37: Elastic And Mechanical Properties Of Expanded Perlite And ...

13

typical kiln feed contains perlite grains with diameters from 0.254 to 2.54 mm. It is

customary to eliminate grains smaller than 25 µm, as they tend to produce very fine dust

particles in the expanded product [21].

At the stage of heating perlite to the softening point, the combined water is vaporised,

producing pressure that causes expansion of the perlite to 4 - 20 times its original volume

leading to a significant reduction of its bulk density [29]. The heating may last from one

to three minutes, at temperatures varying between 760°C and 1100°C. It can be carried

out in either a rotary horizontal expander or a stationary vertical expander as both operate

on the same principle. Figures 2.2 and 2.3 show schematics of horizontal and vertical

expanders, respectively. Detailed descriptions of these two types of expander can be

found in [30].

The heating rate for the expansion process may be optimised. If the perlite granules are

heated too slowly, the combined water is removed through pores without much of a

‘popping effect’, leaving a slightly porous but relatively dense product. However, if the

rate of heating is too high, the expansion of the water may be so violent that the perlite

particles shatter and form numerous fines [31]. The optimum temperature at which perlite

expansion occurs may depend on its chemical composition. Variations in the composition

of the glass affect the softening point, the type and degree of expansion, the size of the

bubbles and the wall thickness between them, as well as the porosity of the product [21].

Also, water content in perlite is another factor affecting the optimisation. It has been

found that perlite containing excessive combined water is likely to shatter upon

expansion, whereas perlite containing much less water is likely to produce higher density

products [32]. A typical value of 3.2% to 3.7% combined water in perlite seems to work

best in most conditions [27]. King et al. [33] reported that the water in perlite above 1.2%

Page 38: Elastic And Mechanical Properties Of Expanded Perlite And ...

14

is loosely held, and therefore is easily removable. However, as dehydration proceeds, the

residual water (below 0.5%) is increasingly more firmly held, which is the effective

portion in producing the expansion of perlite upon heating. Some form of pre-heating is

generally used to reduce the effect of the initial thermal shock, and thus to reduce the

amount of shattered material. It also drives off some of the loosely held water, brings the

larger particles closer to their softening point, and thus reduces the time spent in the kiln

in order to improve the fuel efficiency of the expander [34]. It also can lower the density

of the final product, hence improve the insulating ability of the product. For coarse grades,

in particular, preheating is important as these grades offer less surface area for heat

transfer in a given amount of time, while intermediate and fine grades are not usually

preheated [30].

Although it is possible to achieve some flexibility in the properties of the finished product

properties through adjustment of furnace conditions and through preheating, it should be

noted that all particles do not expand alike [27]. The characteristics of the final product,

such as size, strength and density, depend on the initial granule size, granule water content

and granule composition, together with the design and operation of the kiln and the

heating rate [31]. When expanded perlite is cooled down, it is air classified to separate

any unexpanded particles and fines. Depending on the intended end-use, the expanded

perlite may be further size classified, surface treated or milled. For example, some

expanded perlite is milled down to -100 mesh (0.15 mm) for use as a filter aid, especially

in rotary pre-coat filtration [12]. Micronised perlite particles up to 2 μm can also be

produced by milling for anti-block filler for polymeric film or as a reinforcing filler for

polymers [35].

Page 39: Elastic And Mechanical Properties Of Expanded Perlite And ...

15

Figure 2.2. Horizontal expander and handling system [30].

Figure 2.3. Vertical expander and handling system [30].

Page 40: Elastic And Mechanical Properties Of Expanded Perlite And ...

16

2.2.3 The physical properties of perlite

Perlite textures may vary from classical (onion skin-like) to granular and pumiceous [27].

Classical, or onion skin perlite is the densest of the perlite types and is characterised by

well-developed concentric fractures, a pearly-to-resinous lustre and a grey to bluish black

colour. Granular perlite has a sugary or saccharoidal appearance, is microvesicular2 and

highly fractured, with colours ranging from white to grey. Granular perlite is lighter than

classical perlite, but denser than pumiceous perlite. Pumiceous perlite is extremely

lightweight, white to light grey, frothy and commonly friable. All perlite types are mined,

however the most common commercial one is in granular form [36]. It was found that the

grain morphology affects the expansion ratio. Pumiceous perlite has a higher expansion

ratio than granular perlite, but a lower expansion ratio than classical (onion skin-like)

perlite. The granular perlite has a well distributed microvesicular texture that allows water

vapour to escape easily, and thus only a small degree of perlite expansion occurs,

considerably smaller than that of pumiceous and classical perlites. However, in classical

perlite, the water vapour has difficulty in escaping through the successive onion skin-like

layers. Therefore, the sudden vaporisation results in increasing pressure, thus blowing up

the grain until it explodes and greatly expanding the grain [37]. The light-weight porous

material with a cellular interior structure that perlite transforms into upon heating above

760°C, is shown in Figure 2.4.

2. Vesicular texture refers to the presence of small cavities called vesicles, which were originally gas

bubbles in the liquid magma. This gas was initially dissolved in the magma but has come out of solution

because of the pressure decrease during eruption or as the magma rose before eruption. The texture is often

found in extrusive aphanitic, or glassy, igneous rock [Claudia Owen et al., 2010].

Page 41: Elastic And Mechanical Properties Of Expanded Perlite And ...

17

Figure 2.4. Cellular structure formed internally in expanded perlite particles.

Unexpanded ("raw") perlite has a bulk density ranging from 1.04 to 1.2 g/cm3, while

expanded perlite typically has a bulk density of about 0.032 - 0.4 g/cm3. However, the

true density of perlite does not change as significantly as expected [38]. Expanded perlite

ranges from a fluffy highly porous (85 - 95 vol% [11, 39]) to glazed glassy particles

having a low porosity [19]. Porosity3 provides expanded perlite with a volumetric and

surface absorption capability. The porosity is an advantageous feature for drainage,

aeration and retaining moisture and fertilizers in horticultural uses and landscaping [40].

However, the water absorption in thermal insulation applications is not desirable as the

heat conductivity increases when the pores are filled with water [23]. The expansion

process also creates a brilliant white colour in perlite, which is due to the reflectivity of

the cellular structure. The brilliant white colour is also beneficial in light coloured, visible

3 Porosity is determined as the average ratio between the volume of pores and the total volume of the perlite

grains [Topçu Işikdağ, 2006].

Page 42: Elastic And Mechanical Properties Of Expanded Perlite And ...

18

surface coatings. Colour is also of concern in filler applications, especially when the

colour of the end product is important [27]. The typical physical properties of expanded

perlite are given in Table 2.1.

Table 2.1. Typical physical properties of expanded perlite [23, 26, 39].

Typical Physical Properties

Colour White

GE brightness % 78

Refractive Index 1.5

Free Moisture, Maximum 0.5%

pH (of water slurry) 6.5 - 8.0

Specific Gravity 2.2 - 2.4

Wet density (kg/m3) 103

Bulk Density (Loose) 0.032 - 0.400 g/cm3

Mesh Size Available 4 - 8 mesh and finer

Softening Point 871 - 1093°C

Fusion Point 1260 - 1343°C

Unit weight 2.2 - 2.4 g/cm3

Specific Heat 0.20 kcal/kg°C

Thermal Conductivity at 75°F (24°C) 0.04 - 0.06 W/m·K

Solubility - Soluble in hot concentrated alkali and HF

- Moderately soluble (<10%) in 1N NaOH

- Slightly soluble (<3%) in mineral acids

(1N)

- Very slightly soluble (<1%) in water or

weak acids

Page 43: Elastic And Mechanical Properties Of Expanded Perlite And ...

19

2.2.4 Chemical composition

Perlite is chemically inert, having an approximate pH of 7, and is mainly composed of

silica, alumina and lesser amounts of several metal oxides (sodium, potassium, iron,

calcium and magnesium). Table 2.2 gives the composition of perlite ores from different

countries. As can be seen, although perlite from different origins exhibits considerable

variation in composition, it is characterised by a high silica content (≥ 65%) and alumina

content (11 - 18%) with about 7 - 8% of alkaline content [41]. Burriesci et al. [41] found

that there are no traces of carbonate in perlite ore based on thermogravimetric analysis

[41] and another alumina silicate, tuff [42], and comparison between their weight loss

curves. Considering the linearity and flatness of the weight loss curve in the range 700 -

900°C, a temperature range at which the decomposition of carbonate occurs, they

considered that no carbonate should exist in perlite as, if it did, its presence should have

shown a reduction in weight as was shown for tuff. Although this assumption is valid,

referring to the weight loss curve for perlite in [41], the flatness is not well established.

Only three points in this range were used for both curves which does not seem satisfactory

as every line with three points appears linear.

In addition to the above-mentioned compounds, trace amounts of some other elements

were detected in perlite ores, which are usually less than 2% by weight. Typically, these

elements include Arsenic, Boron, Beryllium, Barium, Chlorine, Chromium, Copper,

Gallium, Lead, Lanthanum, Manganese, Molybdenum, Nickel, Niobium, Neodymium,

Sulphur, Titanium, Thorium, Vanadium, Yttrium, Zirconium and Zinc [28-32].

Furthermore, the presence of trace amounts of rare earth elements (La, Ce, Pr, Nd, Sm,

Eu, Gd, Dy, Ho, Er, Yb, Lu) were detected in the analysis of the chemical composition

Page 44: Elastic And Mechanical Properties Of Expanded Perlite And ...

20

of perlite deposits from Trachilas (Greece) [43], Eastern Rhodope (Bulgaria) [44],

Nathrop (Colorado) and No Agua (New Mexico) [45].

Table 2.2.Chemical composition of perlite (Percent) [22].

Components New Mexico

[29]

Greece

[36]

China

[37]

Turkey

[38]

Iran

[39]

Thailand

[40]

SiO2 74.10 76.10 76.89 68.40 72.32 75.20

Al2O3 13.30 12.16 10.51 15.11 12.62 12.75

Na2O 3.20 3.62 0.80 1.62 2.96 2.85

K2O 4.60 3.03 8.25 4.05 5.02 5.12

CaO 0.60 1.21 0.12 1.72 0.66 0.38

Fe2O3 0.50 1.19 2.48 2.76 0.67 0.99

MgO 0.10 0.27 0.06 0.42 0.21 <0.10

TiO2 0.05 0.14 0.07 n.d. n.d. 0.14

MnO n.d. 0.05 0.04 n.d. 0.66 0.05

P2O5 n.d. 0.03 0.03 n.d. 0.13 <0.05

SO3 0.10 n.d. n.d. n.d. n.d. n.d.

H2O n.d. n.d. 0.31 4.92 n.d. 0.43

LOIa n.d. 2.19 0.64 n.d. 3.75 1.50

a. LOI, loss of ignition.

The chemical composition and physio-chemical properties of perlite (e.g. index of

refraction, specific gravity, etc.) are comparable to those of obsidian and pumice.

However, perlite has a higher water content (generally about 3 - 4%) [41]. Water has an

important role in the expansion process, not only by causing expansion of the grain during

vaporisation, but also by reducing the viscosity of the grain [37]. It is reported that water

Page 45: Elastic And Mechanical Properties Of Expanded Perlite And ...

21

in perlite is in two forms, viz. molecularly dissolved water and hydroxyl groups. The ratio

of these two species of water is variable in perlite [17]. King et al. [33] made an analogy

between the molal entropy increase between 0 to 298 K in perlite and a number of

substances containing water, either as water of crystallisation (e.g. in MgCl2.H2O) or as

water as hydroxyl groups (e.g. in Mg(OH)2). The molal entropy increase in perlite was

comparable to those compounds containing water of crystallisation. They did not

conclude that the water in perlite is held similarly to the water of crystallisation, however,

as they concluded that it is impossible that the large portion of the water in perlite is bound

as hydroxyl groups. Later, Stolper [46] found that the concentration of hydroxyl groups

increases rapidly with increasing total water content at low total water content (<2 wt%),

but its concentration levels off or even decreases at a total water content higher than 3%

by weight. On the other hand, the concentration of molecular water increases slowly at

low total water content and more rapidly at a total water content higher than 3% by

weight. Moreover, it was found that in water-bearing glasses having total water content

of 4% by weight, the dissolved water is equally divided between the two species of water

but in glasses with total water content of >4% by weight, more water is dissolved as

molecular water. Saisuttichai and Manning [47] investigated the perlite ore from northeast

of Lopburi (Thailand) and conducted Fourier transform infra-red spectroscopy on

samples containing 3.40 wt% water, which showed higher molecular water than hydroxyl

groups, which appeared to be consistent with Stolper’s findings. Roulia et al. [37] also

studied perlite grains containing 3.48 wt% water from Trachilas (Greece) and the results

showed consistency with Stolper’s findings as well. Therefore, the portion of hydroxyl

and molecular groups in perlite ores vary depending on the total water content, and it is

Page 46: Elastic And Mechanical Properties Of Expanded Perlite And ...

22

not correct to generally suppose a larger portion of molecular water than hydroxyl groups

without considering the total water content as some sources do (e.g. [21]).

Studies on the nature of the perlitic network have shown that perlite is mostly amorphous

in nature, with a small amount of crystalline inclusions (e.g. quartz, feldspar and biotite).

Herskovitch and Lin [48] studied crude perlite from Milos Island4 and reported it to have

89.4% by volume of the amorphous phase and 10.6% by volume of the crystalline phases

(feldspars, biotite, quartz, magnetite and chlorite). The presence of crystals in crude

perlite is correlated to its origin, as represented in Table 2.3. Crystalline silica (e.g. quartz,

cristobalite and tridymite), though it is present in small amounts, is of concern during

production and marketing. Crystalline silica is classified by the International Agency for

Research on Cancer (IARC) as a class 2A (probabale carcinogen) in humans when it is

inhaled but not when it is ingested or contacted. In most commercial perlite deposits,

however, the amount of crystalline silica is very low (<1-3 wt%). This can be even further

lowered during processing and use, e.g. dilution in end product [49]. Though crude perlite

comprises both amorphous and crystalline phases, the expanded perlite has been found to

be clearly amorphous. X-ray diffraction analysis conducted by Barth-Wirsching et al. [50]

and Chakir et al, [51] indicated that expanded perlite is amorphous. Roulia et al. [37] also

found that the expansion of perlite, affected by rapid thermal treatment, only takes place

in the amorphous phase of raw perlite and that crystallites do not expand and are rejected

as unexpanded waste product. Chemical analysis of EP has shown a similar composition

to crude perlite, but with a lower amount of combined water. It was found that the

4 Perlite was supplied by Habonim Industries, Moshav Habonim.

Page 47: Elastic And Mechanical Properties Of Expanded Perlite And ...

23

expansion of perlite results in a remarkable reduction in hydroxyl group content with a

smaller reduction in molecular water [47].

Table 2.3. Mineral phases detected in perlites of different origins.

Origin Minerals

Socorro, New Mexico[29] feldspar, cristobalite, zeolites, fluoride

Searchlight, Nevada [29] feldspar, quartz, biotite, magnetite, limonite, zeolites

Korea [43] plagioclase, biotite, opaque minerals, traces of hornblende,

apatite, zircon and sanidine

Kawalan, Yemen [44] k-feldspar, plagioclase, pyroxenes, chlorite, quartz, serpentine

Sardinia, Italy [26] α-quartz, mica, feldspatiods

Milos Island, Greece [23] feldspar, quartz, biotite

Lopburi, Thailand [32] plagioclase and alkali feldspar, biotite, opaque minerals

In addition, it was found that, upon expansion, fluorine may be lost from perlite,

presumably as HF, and it becomes increasingly available to water. The increase in the

availability of fluorine to water may limit the use of such perlite in horticulture

applications [47] as the detrimental effects of fluorine to plants are well-known [52]. It

was also found that, upon expansion, the availability of Ca (Calcium) and Mg

(Magnesium) is reduced, but that the extractable K (Potassium) is increased (especially

in highly potassic perlitic rhyolites). However, the increase in the water solubility of K

on expansion can be useful in horticultural applications (e.g. as a possible slow-release

source of potash) [47].

Page 48: Elastic And Mechanical Properties Of Expanded Perlite And ...

24

2.2.5 Applications

The low density, relatively low price and good insulating properties of perlite have led to

the development of many commercial applications. Figure 2.5 illustrates the

proportionate use of expanded perlite computed by the U.S. Department of the Interior

U.S. Geological Survey [92]. These applications can be classified into three general

categories: construction applications, horticultural applications and industrial

applications.

Figure 2.5. US expanded perlite use by application.

1. Includes acoustic ceiling panels, pipe insulation, roof insulation board, and unspecified formed

products.

2. Includes absorbents, laundries, paint texturisers, and other miscellaneous uses.

3. Estimated and reported data with specific use unknown

Page 49: Elastic And Mechanical Properties Of Expanded Perlite And ...

25

2.2.5.1 Construction applications

The low density, high heat insulating value, outstanding fire resistance and good sound

absorbing properties of expanded perlite give it a number of construction applications

mainly as formed products, concrete aggregate, masonry and cavity-fill insulation, and

plaster aggregate. Construction uses of perlite account for about 60% - 70% of the total

consumption in major markets. In the United States, the manufacture of formed products

accounts for about 53% of total perlite use (see Figure 2.5 ). However, such a large

demand for the use of perlite in formed products might not be seen in Europe due to

differences in construction methods. In the US, most homes are constructed on a timber

frame with extensive use of gypsum and insulating boards to form internal surfaces. There

is much less use of brick or other masonry products for external walls, in comparison with

Europe. Therefore, there is less use of light-weight insulating plasters, mortars and loose

fill insulation in the US, compared to Europe, which results in a higher percentage of

perlite use for formed products in the US [53]. In the following sections, the different

construction applications in which perlite is currently utilised, or is substituted for a

traditional material, are explained.

Gypsum perlite-based plaster

Perlite, instead of sand, can be mixed with gypsum and water for plastering on a lath in

order to provide a resilient wall or ceiling capable of withstanding stresses and thermal

changes. Indeed, the manufacturing of gypsum-perlite and prefabricated gypsum boards

has become one of the major uses of perlite. The plasters made of expanded perlite are

characterised by their light-weight, ease of application and savings in structural steel.

They are resistant to cracking under stress, highly fire resistant and have excellent sound

Page 50: Elastic And Mechanical Properties Of Expanded Perlite And ...

26

absorbing properties [21]. It is claimed that gypsum-plaster made of expanded perlite

aggregate weighs about one third of that conventionally made with sand, and has six times

the thermal insulating value [54]. Perlite can also enhance the workability of plasters and

mortars and can be used to create a texture on the finished surfaces [53].

Concrete

Concrete is a composite material and since around 75% of its volume is occupied by

aggregate, the performance of concrete is greatly affected by the properties of the

aggregate. The use of light-weight and porous aggregate as a constituent of concrete

enables the production of light-weight concretes [55]. Light-weight concretes are

economically beneficial, by reducing heat conductivity and unit weight. Reducing unit

weight is especially beneficial in reducing the damage by earthquakes in high buildings

[56]. The low density of EP particles make them suitable for the production of light-

weight concretes [57, 58]. At the same time, EP particles greatly improve thermal

insulation, reduce noise transmission, and they are rot, vermin and termite resistant [59].

The high water absorption capability of EP particles can provide water for the hydration

of cement in the later stages of curing, which is especially advantageous in concrete with

low water/cement ratios [60]. EP particles can also have pozzolanic activity5 and can be

used as a mineral admixture when finely ground [61]. The properties of light-weight

5 A pozzolan is a siliceous or siliceous and aluminous material which itself has little or

no cement-like value but in finely divided form and in the presence of water will react

chemically with calcium hydroxide (slaked lime) at ordinary temperatures and form

compounds exhibiting cement-like properties [61,62].

Page 51: Elastic And Mechanical Properties Of Expanded Perlite And ...

27

concrete, however, are usually different from normal concretes. Sengul et al. [62] made

light-weight concrete by partially replacing the natural aggregate (e.g. sand) with EP

particles, from 0% to100% with 20% increments. The replacement of natural sand by EP

particles resulted in a reduction in the unit weights by 8.4 - 80.8 % and a reduction in the

compressive strength of the concrete by 40 - 99.7% due to the lower strength of the perlite.

Topçu and Işikdağ [56] also reported a reduction in compressive and splitting tensile

strength of concrete by about 54% and 47%, respectively, for the replacement of the

natural sand by 60% EP particles. The accompanying air entrainment was also effective

in lowering the strength of the concretes. Not only the compressive strength, but also the

modulus of elasticity were adversely affected. For the replacement of natural sand by EP

in the range 20 - 60%, Sengul et al. [62] reported a reduction in the modulus of elasticity

by 34 - 83%. This is due to the fact that the modulus of elasticity of concrete is a function

of the modulus of elasticity of the constituents (i.e. hardened cement paste and concrete).

Concrete made with perlite aggregate may not be as strong as that made with pumice or

heavier materials, however it may be used where compressive strength above about 13.79

MPa is not required [21]. On the other hand, increasing the expanded perlite ratio

improves the thermal insulation of concrete by increasing the total porosity. Porosity is

one of the factors affecting the thermal conductivity of concrete. Introducing enclosed

pores reduces the thermal conductivity due to the low thermal conductivity of air [62].

These materials are beneficial in insulating spaces around heating, steam or coolant pipes;

providing insulating bases for ovens, furnaces; cold storage tanks; and wherever light

structural decks are desired [53].

Page 52: Elastic And Mechanical Properties Of Expanded Perlite And ...

28

Loose fill insulation

Another important construction application of perlite is as loose fill insulation. Perlite’s

outstanding low density, thermal insulation and fire resistance allow it to be widely used

as loose fill insulation in cavity walls and for filling cores, crevices, mortar areas and the

air holes of masonry blocks. There are several other materials that can alternatively be

used, such as fibre glass; expanded polystyrene foam, beads and panels; and vermiculite.

However, the valuable features of being non-settling and the free flowing behaviour of

perlite make it more preferred [53]. Moreover, the use of perlite as loose fill in masonry

construction has been proven to reduce the transmission of heat by as much as 50 - 70%

when filling the cavities of concrete blocks. As already mentioned, there is a relationship

between thermal conduction and density at various mean temperatures. Based upon the

reported data, the density of 32 - 173 kg/m3 is recommended for loose fill insulation [63].

However, when using perlite as loose fill insulation, its ability to absorb moisture should

be taken into account. Perlite moisture absorption can adversely affect the insulation value

of expanded perlite, or increase settling problems within a conventional wall construction

where perlite is used. Nevertheless, this problem has been alleviated by coating expanded

perlite particles or adding a moisture repellent to the mix before its application into the

wall. Coating the perlite particles with sodium silicate, for example, can improve the

moisture resistance by closing the open pores of perlite and thus reducing the water

retaining property [64]. Alternatively, it can be mixed with silicone (less than 5% by

weight), a natural resin or Bitumen [65, 66]. However, silicon polymers are expensive

and there are hazards associated with their use. Instead, the annealing of expanded perlite

was found to be effective in reducing perlite’s water absorption [67]. In the annealing

process, expaneded perlite is heated to a temperature sufficient to soften the perlite

Page 53: Elastic And Mechanical Properties Of Expanded Perlite And ...

29

surface (usually ranging from 426 to 537 °C [67]) and to heal many of the surface cracks

and fissures. Sealing the surface fissures results in a reduction of water absorption.

Ceiling tiles

The use of ceiling tiles has had tremendous growth over the last few decades owing to the

advantages it offers, like sound absorption, fire resistance, good thermal insulation, light

reflectance and structural support. The increase in the use of ceiling tile systems has

resulted in an increase in the use of perlite as a component of many kinds of ceiling tiles.

Standard ceiling tiles are typically made from a combination of mineral wool or slag wool,

fibreglass, binders (e.g. starch), and sometimes inorganic fillers (e.g. mica, wollastonite,

silica, calcium carbonate). Expanded perlite is added to the mixture to enhance the

acoustic insulation, thermal insulation and fire resistance properties of the tile. Depending

on the desired properties of the end product, the proportion of expanded perlite can vary

from 0% to75% of the total mixture by weight. Moreover, perlite can be used to create

different textures to the visible surface of the tile [53].

Starch is a well-known binder in ceiling perlitic tiles. The use of starch has been found

useful for acoustic or sound insulation as these tiles do not bleed through or discolour the

material if painted. Baig [68] used starch as a binder in making acoustic perlitic tiles and

produced tiles with acceptable properties, improved drainage time and lower

manufacturing costs. Denning et al. [69] also used starch in manufacturing insulation

boards and reported an unexpected water repellence property. The product was soaked

for three weeks and it was found that material had just slightly softened. It was also

suggested that the addition of a water-repellent substance to achieve greater water

repellence is possible. However, the addition of water repellent substances to acoustical

boards is limited to those that would not discolour the board. Examples of sound

Page 54: Elastic And Mechanical Properties Of Expanded Perlite And ...

30

insulation boards made using starch are given in Table 2.4. However, the problem with

the use of starch is that under certain environmental conditions, like increased temperature

and humidity, the board loses its dimensional stability. To deal with this problem, latex

binders were found to be useful. Kesky [70] and Deporter et al. [71] used latex binders in

manufacturing perlitic ceiling tiles and reported perlite insulation boards made with latex

binder show a higher MOR (Modulus of Rupture), higher breaking load, lower water

absorption and better dimensional stability than perlitic ceiling tiles made with starch.

Examples of latex binders, having glass transition temperature ranging from about 30° C

to about 110° C, and useful in ceiling tiles include polyvinyl acetate, vinyl acetate/acrylic

emulsion, vinylidene chloride, polyvinyl chloride, styrene/acrylic copolymer and

carboxylated styrene/butadiene [72].

Table 2.4. Examples of perlitic sound insulation boards in which starch is used as binder [69].

Example 1# 2# 3# 4#

Expanded perlite........percent 67.5 62.5 49.6 60.0

Paper pulp fibre.........percent 22.5 27.5 28.9 30.0

Starch.........................percent 10.0 10.0 20.0 10.0

Modulus of rupture ……kPa 357.84 488.84 409.55 536.41

Under floor insulation

Water repellent, dust-suppressed perlite produced in accordance with ASTM: C 549.81

is used for under floor insulation. Perlite under floor insulation can be employed under

floating concrete floors, asphalt floors and floating board floors (thickness 6 - 10 cm). It

is especially effective in the reduction of sound transmission in construction components,

Page 55: Elastic And Mechanical Properties Of Expanded Perlite And ...

31

from floor to floor, from floor to walls, from footsteps and from under floor piping

systems. The neutral pH of perlite also prevents the development of corrosion in piping

and electrical wiring that may be installed in the under floor area [63].

Fire resistant boards

Perlite’s outstanding properties of light weight and thermal insulation are used for

producing light-weight insulation panels for fire protection where substantial weight

saving is important, especially in the marine, aviation/aerospace and land/rail transport

industries. For example, light-weight insulating materials that have high thermal

insulation and fire resistance are suitable for components of vehicular interiors such as

cabins and cargo holds, partitions and fire doors, or for transporting combustible materials

[73]. The thermal insulation boards discussed in [74] and [73] are representatives of these

types of light-weight perlite-based insulation boards.

2.2.5.2 Horticultural applications

The chemical resistance and porosity of perlite are important features in horticulture

applications when soils are wet for prolonged periods or where drainage, aeration and

optimum moisture retention are required. The high brightness of finely ground perlite

applied to the surfaces of seed blocks reflects light to the underneath of seedlings and

helps with rapid and sturdy growth [27]. Perlite can be used as a soil amendment or alone

as a medium for hydroponics. Studies have shown that outstanding yields can be achieved

when perlite is used as a soil-less growing media in hydroponic systems. When used as

an amendment, perlite’s high permeability, and low water retention can be beneficial and

helps to prevent soil compaction. Other features of perlite advantageous for horticultural

Page 56: Elastic And Mechanical Properties Of Expanded Perlite And ...

32

applications are its neutral pH and the fact that it is sterile and weed-free. Perlite is also a

good carrier for fertilizer, herbicides and pesticides and for pelletising seed [75].

2.2.5.3 Industrial applications

Perlite has a broad range of applications in industry. These include the use of perlite as

fillers, filter aids, and in cryogenic and high temperature applications. Perlite is also used

as silica-alumina source for the synthesis of zeolites. These applications are briefly

described below.

Cryogenic applications

Expanded perlite particles contain countless glass-sealed particles, which accounts for its

excellent insulating properties. Expanded perlite is a very effective insulator at

temperatures below -100 °C and is widely used in insulated storage tanks containing

liquid hydrogen, helium, oxygen and other gases [76]. Liquefied gas storage tanks (Figure

2.6) are spherical, double-walled vessels in which the space between the inner and outer

wall is evacuated and filled with expanded perlite as the thermal conductivity of expanded

perlite under evacuated conditions is many times lower than under non-evacuated

conditions [53].

Figure 2.6. Spherical gas tank farm in a petroleum refinery [116].

Page 57: Elastic And Mechanical Properties Of Expanded Perlite And ...

33

Perlite is not as hygroscopic as other siliceous powders and thus it is easier to prepare for

vacuum applications. The low density and correspondingly high porosity (0.85 ≤ ѱ ≤

0.95) of expanded perlite are also two characteristic properties which make the material

suitable for vacuum insulation [77]. Heat transfer through an evacuated porous material

like perlite involves the simultaneous operation of the three transport mechanisms of

thermal radiation, solid conduction and gas conduction. The radiant heat that passes

through an EP particle originates at the hot boundary wall and from the neighbouring

particles. This radiant energy is of a diffuse type, and, as it moves through the particle, it

is scattered and partially absorbed. The absorbed radiation heats the particle. However,

the scattering effect due to the symmetrical shape of typical vessels (e.g. concentric

spheres or cylinders) is not of great importance in the thermal radiation problem of vessels

filled with insulating porous particles [67]. To reduce the heating effect of radiation, the

addition of opacifiers has been found to be effective. Alternatively, metallic powders can

be added to the insulating particles to scatter and absorb radiation. Kropschot and Burgess

[78] found that the addition of 60 wt% aluminium powders (600 µm) can reduce thermal

conductivity of perlite with a density of 64 kg/m3 by 50%.

Thermal energy is also transferred by conduction through the solid fraction of the EP

particles. However, the solid heat conductivity of a porous insulation material is usually

much smaller than the thermal conductivity of the pure solid fraction as the morphology

of the material influences the heat transport [79]. In addition, the particle size was found

to affect the solid conductivity of evacuated perlite beds. When the particle size decreases

at a constant density, thermal conductivity decreases. The contact resistance increases

with decreasing particle size because the greater strength and curvature of small particles

Page 58: Elastic And Mechanical Properties Of Expanded Perlite And ...

34

decreases the contact area and there are large numbers of contacts per unit length. In

addition, at low temperatures, phonons carrying most of the heat energy have a long wave

length with a long mean free path. If the particles become smaller than the mean free path

of phonons, the scattering and collisions of phonons at the walls of particles becomes

important and reduces the rate of heat transfer [80]. This effect is strengthened by the

internal structure of perlite particles which are cellular. Therefore, reduction of the

particle size increases the resistance to the flow of thermal energy inside each particle.

The morphology of perlite particles is also effective in the significant minimisation of gas

convection. Convection is suppressed inside the particle pores, as the emerging buoyancy

forces do not exceed friction [81]. Though gas convection is almost eliminated, depending

on the level of vacuum there is some residual gas contributing to heat transfer by gaseous

conduction. The small particle size of the powder limits the gaseous heat transfer in

evacuated powders to free molecular gas conduction as the mean free path of the gas

molecules, which describes the average distance a gas molecule travels between

consecutive collisions, is large relative to the distance between surfaces of the adjacent

particles. Aside from this, the reduction of pressure in high vacuum increases the mean

free path of the gas molecules. In such conditions, gas molecules rarely collide with each

other, thus the molecule to molecule conduction is eliminated. The energy is exchanged

between the surface and the colliding molecules as the gas molecules travel unhindered

between the confining walls of the adjacent solid particles. Therefore, the particle

structure imposes an upper limit on the mean-free path within the insulation [82].

In addition to perlite’s low thermal conductivity, its free-flowing nature, the high

strength/compaction resistance of the particles, the absence of shrinkage or slumping and

the ease of handling and installing are features that make expanded perlite ideal for

Page 59: Elastic And Mechanical Properties Of Expanded Perlite And ...

35

cryogenic applications. Perlite is also very effective in insulating non-evacuated low

temperature storage vessels for oxygen, nitrogen, liquefied natural gas, and similar

products. The effective insulating properties of evacuated perlite are also utilised in some

domestic refrigerators, in which evacuated plastic bags filled with perlite are used for

insulation [83]. Other applications include the insulation of industrial cold boxes, test or

processing equipment, and shipping containers and in food processing [53].

Fillers

One of the main industrial applications of expanded perlite is as fillers, in which its high

surface area, permeability, low density and inertness and reinforcing characteristics can

be very beneficial. Fillers are organic or inorganic particles added to materials (e.g.

plastics, composite materials and concrete) to reduce the consumption of more expensive

binder materials, and thus to improve the economics of the end product, in which case

they are also called extenders. They are also used to improve some properties of the

mixture by developing a beneficial chemical interaction with the host material; these are

called functional fillers [84]. It is useful to distinguish between the three classes of perlite

used in filler applications. One is produced by further milling and classifying expanded

perlite to produce broken particles with jagged interlocking structures. These are, for

example, used as insulating fillers in the manufacture of texture coatings, as anti-block6

6 Anti-block products are additives incorporated in plastic films to decrease the adhesion or blocking

between touching layers of plastic films during fabrication, storage or use. This can be accomplished by

slightly roughening the film surface by surface treatment with wax/polymers or by the inclusion of anti-

block filler products into the plastic films [21].

Page 60: Elastic And Mechanical Properties Of Expanded Perlite And ...

36

fillers in plastic films, and as reinforcement fillers in polymers [85]. Another class of

perlite used in filler applications is perlite hollow glass microspheres. Perlite

microspheres consist of one or a few inner cells, in contrast with the large number of cells

found in the larger, standardly produced expanded perlite particles. Perlite hollow glass

microspheres are produced under stringent quality control conditions utilising special

technologies. The feedstock for production of microspheres is crude perlite. Upon

expansion, glassy, spherical-shaped hollow particles are produced, as shown in Figure

2.7. Microspheres generally have a diameter smaller than 140 microns, with the average

between 30 and 70 microns. Depending on the desired properties, the microspheres can

be further surface treated with either silicone or silane, which can increase the

hydrophobicity (water repellency) of the microspheres. When perlite microspheres are

used as fillers, a number of important characteristics are imparted into the end-product.

These include improved crack resistance, higher impact, nailing and stapling resistance,

enhanced machinability and sanding features, finer texture with an increased workability,

reduced weight and shrinkage, and a shortened drying time.

Examples of applications in which perlite microspheres as fillers are desired include in

textured and acoustical coating mixes, adhesives and sealants, wall-patching compounds,

thermo-set castings, syntactic foam, sheet moulding (SMC) and bulk moulding

compounds (BMC), rotational moulding, stucco, fibre-reinforced product (FRP), spray

and hand lay-up, specialty coatings and block filler paints [27]. Lastly, the third class is

those which might be called basic or general light-weight perlite fillers, where traditional

expanded perlite with sizes ranging from 500 to 2000 microns is used as fillers in products

and mixes.

Page 61: Elastic And Mechanical Properties Of Expanded Perlite And ...

37

Figure 2.7. SEM image showing EP microspheres consisting of one or more microcellular

bubbles [27].

Filter aid

Perlite is chemically inert in many environments owing to the high silica content, usually

greater than 70%, and are adsorptive, which make them excellent filter aids [9]. Filter

aids are finely divided solids, chemically inert and light in weight. They help to control

flow and remove finely distributed colloidal suspensoids from liquids. They can be used

in two different ways, ‘precoating’ and ‘body feeding’. In the ‘precoating’ case, filter aids

are applied as a thin layer over the filter before the suspension is filtered. This will

enhance clarity and filter rate in the filtration process, and prevent gelatinous-type solids

from clogging the filter medium, which can cause its resistance to become excessive.

Furthermore, it facilitates the removal of filter cake at the end of the filtration cycle. The

second application method referred to as ‘body feeding’ involves the incorporation of a

certain amount of filter aid to the suspension before filtration. The addition of filter aids

increases the porosity of the slug, reduces the loading of undesirable particulate at the

filter medium, reduces the resistance of the cake during filtering, and hence maintains the

Page 62: Elastic And Mechanical Properties Of Expanded Perlite And ...

38

desired flow rate. Depending upon the specific separation involved, perlite products may

be used in precoating, body feeding, or both [31, 35].

The production of perlite filter aids involves careful selection of the ore, precise

monitoring of crushing and screening, expansion, light milling to break the closed

particles, followed by further screening and classification. This process produces

fragments of perlite particles with a precisely controlled particle size distribution to give

the desired clarity and flow rate in different applications. In most applications, perlite

filter aids are used to filter solid particles having one micron or larger size [86].

Perlite competes with other filter media like diatomite, diatomaceous earth and cellulose.

However, it has found extensive niches in many applications due to the advantages it

offers, like rapid filtration of viscous liquids, particle size distribution, high surface area,

high porosity, inertness in most mineral and organic acids, low density compared to other

filter media, and relative cheapness [53].

Perlite is particularly suitable for the filtration of viscous liquids such as edible oils, and

this is one of its main applications. It is also used to remove fine solids like algae and

bacteria from beer and filtering out yeasts and other fine solids from wine. Other

applications include the filtering of various chemicals, pharmaceutical solutions, solid

particulates from waste water streams, water for municipal systems and swimming pools,

and the clarification of foodstuffs (e.g. sugar solutions and syrups) [53].

High temperature applications

Perlite insulation is used in a variety of high temperature applications such as in the steel

and foundry industries (like hot topping, ladle topping and risering), in insulating

formulations at a service temperature of up to 1000°C, and in the manufacture of

refractory bricks or castables.

Page 63: Elastic And Mechanical Properties Of Expanded Perlite And ...

39

Both raw perlite and expanded perlite can be used in ladle topping applications, but the

raw form is more common. As the perlite ore is added to the ladle, it reacts with the slag

and allows easy removal of the slag layer. The expanded perlite can also act as an efficient

insulating layer to maintain the temperature of molten metal during the steel making

process.

In the hot topping of castings and risers, up to 20% perlite by weight is mixed with

exothermic powders. This controls heat loss from the casting, and thereby prevents

shrinkage cavities. Pre-formed boards for hot tops and hollow cylinders for risers also

perform the same function as hot topping and risering powders and compounds.

Perlite is also used as a cushioning agent in foundry cores and moulding sand mixtures to

compensate for the expansion of silica as it goes through phase changes at temperatures

above 540 °C. It helps to minimise casting defects like buckles, veining, fissuring, and

penetration, and to improve the venting of gases by increasing the permeability of core

sand [53].

An additional use for perlite is in the manufacture of refractories by inducing porosity,

reducing density and thermal conductivity where the average temperature does not exceed

1100 °C. At higher temperatures, perlite refractories are used as backup insulating layers

for higher duty refractories [87].

Zeolite synthesis

Zeolites are a well-defined class of crystalline aluminosilicate minerals. They have three-

dimensional structures arising from a framework of tetrahedral units of [SiO4] - and

[AlO4] arranged in a polyhedral structure linked by all corners [88]. Zeolites have been

successfully used in the chemical industry, agriculture, aquaculture, and environmental

protection because of their remarkable physical and chemical properties which include

Page 64: Elastic And Mechanical Properties Of Expanded Perlite And ...

40

molecular sieving, adsorption, catalytic and ion exchange functions. This has resulted in

a great increase in demand for these products. There are more than 45 natural types of

zeolite; however, only clinoptilolite, mordenite and chabazite and possibly phillipsite are

currently used as commercial products. The use of zeolites in most industrial applications

requires certain specifications. Thus, synthetic zeolites can be good solutions to provide

the strict specifications imposed on adsorption and catalytic processes. Furthermore, it

would be possible to maintain a constant composition of the products during zeolite

synthesis, while natural zeolites usually show considerable chemical and mineralogical

variation [89]. Consequently, zeolite synthesis from low-cost silica-alumina sources has

been the aim of many studies. Synthetic zeolites can be produced from a variety of natural

silica-alumina materials including acidic volcanic glasses, such as natural and expanded

perlite and pumice (e.g. [90, 91]). Experimental investigations into the formation of

zeolites from perlite have shown that this low-cost silica-alumina material is a suitable

material for zeolite synthesis. It has been shown that perlite can be successfully used in

the synthesis of the zeolites ZK-19, W, G, F, Na-Pc, HS, ZSM-5, A, V, Pc, Na-P1, P1,

K-G, Y, Analcime, sodalite octahydrate and calcium zeolite [50, 88, 89, 92-95].

Other industrial applications

Advantage can be taken of the mild abrasiveness of crude perlite in the production of

soaps, cleansers, and polishes. Crude perlite is also used for the stone-washing of fabrics

in the textile industry. Traditionally, pumice was used to age fabrics and to soften the

finish. Perlite, however, has a softer impact on the fabric and produces less wear on

industrial laundry equipment than pumice, which can be excessively abrasive to the fabric

[53].

Page 65: Elastic And Mechanical Properties Of Expanded Perlite And ...

41

Perlite uses also include its application in oil-gas, water and geothermal well cementing

and grouting, where the perlite prevents lost circulation of drilling lubricants. Moreover,

perlite is used as a porous support for catalysts and reagents in various chemical reactions

[19, 76].

2.2.6 Concluding Remarks

In Section 2.2, the physical and chemical properties of perlite, the process of its expansion

and its application in construction, horticulture and industry were discussed. However,

one of the main properties of this material, its mechanical properties, has been missed.

Although this material has found uses in various applications, to the author’s knowledge,

barely any research has been conducted into the mechanical properties of perlite. Hence,

one of the main gaps identified in the literature is perlite’s mechanical properties, such as

strength, Young’s modulus and Poisson’s ratio in its expanded and unexpanded forms.

Page 66: Elastic And Mechanical Properties Of Expanded Perlite And ...

42

Foams

2.3.1 Introduction to foams and syntactic foams

A foam is an assembly of cells (i.e. small compartments) packed in three dimensions to

fill space. Foams can be solid or liquid. In liquid foams, the cell walls are interconnected

liquid films and there is gas inside the cells. Solid foams can be considered as a composite

made of a solid and a fluid (i.e. mostly a gas but it can also be a liquid as is the case in

live tissues) [96]. Gibson and Ashby [97] defined solid foams as cellular materials with a

relative density, which is the density of the foam divided by the density of the solid from

which the cells are made, of less than 0.3. For higher relative densities, there would be a

transition from the cellular structure to solids containing isolated pores. If a solid from

which the foam is made is contained in the cell edges only, and hence the fluid phase is

interconnected, the foams are called open-celled. However, if the solid is not only

contained in the cell edges but also in the cell walls, so that each cell is sealed off from

its neighbours [97], the foams are called closed-celled. Of course, some foams contain a

combination of closed and open cells. Examples of a typical open cell foam and a typical

closed cell foam are illustrated in Figure 2.8. Most foams do not have regular packing of

identical cells, and are comprised of cells with different sizes and shapes. The cell

dimensions can vary, from less than one micrometre in microcellular materials up to

millimetres. The thickness of the cell walls and edges can also vary from 1 to 10 µm as

is the case in some polyurethane foams [98]. In foams, the thickness of the cell edge

(struts) is usually larger than that of the cell walls. This is due to the surface tension

drawing more solid material to the cell edges during the foaming process [97]. The shape

Page 67: Elastic And Mechanical Properties Of Expanded Perlite And ...

43

of cells can also be more or less random due to differences in the cell dimensions, wall

and edge thickness as well as their chemical composition [99-101]. However, a common

pattern may be observed, for example in cork, the cells are usually prismatic and arranged

in parallel columns on staggered bases [96].

(a)

(b)

Figure 2.8. (a) Closed cell Polyurethane foam [102], (b) Open cell Polyurethane foam [3].

Foam-like materials are very common in nature. Examples are cork, sponge, wood and

coral. While natural cellular materials have been used for centuries, such as cork, which

has been used as bungs for wine bottles since Roman times, mankind has made a wide

Page 68: Elastic And Mechanical Properties Of Expanded Perlite And ...

44

variety of cellular materials using polymers, metals, ceramics, and even composites via

different foam processing techniques [97].

One of the main types of man-made cellular materials is polymeric foams. They are made

by first heating a polymer into the liquid form, then gas bubbles are introduced. These are

allowed to grow and when they have reached the required size the material is ‘solidified’

by cross-linking or cooling [97]. The gaseous phase is introduced by either mechanical

stirring or by using blowing agents. There are two type of blowing agents: chemical

blowing agents and physical blowing agents. Physical blowing agents are inert gases (e.g.

nitrogen and carbon dioxide) forced into solution in the hot polymer at high pressure

which expand into bubbles due to reducing pressure. Alternatively, low melting point

liquids (e.g. methylene chloride) are mixed with a polymer and vaporised into bubbles on

heating. Chemical blowing agents are additives which give-off gases due to either

chemical reactions or thermal decomposition [97, 103]. Each of these processes can yield

open or closed cell foams. It is the rheology and surface tension of the fluids in the melt

which govern the final structure of the foam [97]. Polymeric foams can be categorised

into two main groups, the thermoplastic and thermosetting foams. The thermoplastics can

usually be reprocessed and recycled; however, thermosets are not recyclable as they are

usually heavily cross-linked. Within these categories, the polymeric foams can be further

classified as rigid, semi-rigid, semi-flexible or flexible. Furthermore, a solid polymeric

foam can either be composed of closed (e.g. polystyrene foam used for coffee cups) or

open cells (i.e. the polyurethane foam in seat cushions). Each polymeric foam possesses

unique physical, mechanical and thermal properties, which are attributed to the polymer

matrix, the cellular structure (i.e. cell density, expansion ratio, cell size distribution, cell

geometry and cell integrity) and the gas composition (i.e. the molecular weight of the

Page 69: Elastic And Mechanical Properties Of Expanded Perlite And ...

45

constituting components) [104, 105]. These foamed polymers have excellent energy

absorption characteristics and are widely used for shock mitigation in vehicles of all

types, in packaging, in cushioning and as an insulation material. They are also widely

used as cores in sandwich panels, where two relatively thin but stiff face sheets are

attached on either side of a thicker light-weight core [3]. In combination with stiff and

strong face sheets, they produce a construction with higher stiffness to weight and

strength to weight ratios than a structure made of a thick single phase material [106].

It has been shown that a foamed resin with higher specific strength and modulus can be

obtained if the resinous matrix is reinforced with different types of fillers, such as solid

particles, hollow particles, short fibres, etc. [107-110]. However, producing low density

core materials has increased interest in the use of low density fillers such as hollow glass

microspheres [111]. Syntactic foams are a special class of closed cell foams made by

embedding hollow particles in a polymer matrix. They have porosity in the form of closed

cells, and are especially useful in applications where a combination of low density, high

compressive strength and high damage tolerance is required [5]. Moreover, the porosity

enclosed by the thin and stiff shells of hollow particles provides syntactic foams with low

moisture absorption, low thermal and electrical conductivity, as well as high dimensional

stability at elevated temperatures [112]. Most of these structural foams are rigid, although

more ductile and even elastomeric syntactic foams can be manufactured using suitable

hollow particles and matrix materials. There are a wide variety of hollow microspheres

for use in syntactic foams. The diameters of the hollow microspheres ranges from 1 to

350 µm, though they are on average about 50 to 100 µm. Hollow microspheres are made

from glass, ceramic or any polymeric material [2, 113]. Most polymers can be used in the

manufacture of syntactic foams. Producing thermoplastic syntactic foams is generally

Page 70: Elastic And Mechanical Properties Of Expanded Perlite And ...

46

more difficult as these materials are usually melt processed. Syntactic foams are mostly

processed from thermosetting polymers, many of which are available in a castable liquid

'prepolymer' form. The most prevalent thermosetting polymer matrices used in the

manufacture of syntactic foams are the unsaturated polyester and epoxy resins, although

the use of polyurethane, silicones, polyimides, phenol-formaldehyde and urea-

formaldehyde have been reported [113-117]. In the following section, different methods

used to manufacture syntactic foam are explained.

2.3.2 Manufacturing syntactic foams

The manufacturing process used to produce a syntactic foam is important in terms of the

final product’s properties, ease of production and manufacturing cost. Various

manufacturing methods are available for processing syntactic foams. These vary from the

simple blending of components to the coating of particles by resin [113]. The particle and

binder concentration are found to be crucial for ease of manufacturing. In addition, the

choice of process parameters (temperature, mixing time and addition sequence) is

considered influential in the manufacturing process [118].

The most common method of manufacturing syntactic foams involves the impregnation

of the desired quantity of filler particles (e.g. hollow glass microspheres) in a

thermoplastic polymer solution (polymer + solvent) or in liquid thermosetting

prepolymers (which might contain solvent/viscosity control) [117]. This method ensures

the uniform coating of individual particles by resin and hence, a uniform distribution of

resin among particles can be achieved. As the filler particles (e.g. hollow-glass, organic

or carbon microspheres) cannot withstand high pressure, it is practically not suitable to

use extrude or injection-moulding of mixtures (slurry). Instead, the mixture is usually

Page 71: Elastic And Mechanical Properties Of Expanded Perlite And ...

47

‘cast’ to minimise particle damage. Solvent is evaporated before final curing, though the

entire removal of solvent and minimisation of solvent entrapment is a particular

challenge. There are several impregnation techniques in common use. In one method,

particles of known mass are introduced to a resin solution. Subsequently, the solvent is

removed to get a dough, which is transferred into the mould and allowed to be cured [119-

123]. In another method, the mould is filled with the desired quantity of particles and

sometimes vibrated to facilitate particles packing [113, 124]. Subsequently, a

premeasured amount of the resin solution is poured over the particles. The solution

penetrates the porous regions between particles due to gravity and capillary forces, and

sometimes with the aid of a vacuum. Another method is to first fill the mould completely

with particles to measure particle quantity where different particle sizes provide different

quantities [125]. Subsequently, the measured quantity of particles is mixed with the resin

solution and the prepared mixture is poured evenly into the mould. To maintain the

particles in a well dispersed state, compression is applied on top of the mould lid or with

the aid of a vacuum. This method has the advantage of restricting the spheres from

floating to the surface during the foam production. In this method, as in the previously

explained methods, removal of the solvent before final curing is required. These methods

have many disadvantages. Perhaps the most serious problem with excess solvent is

drawing off resin from the microspheres during the solvent’s evaporation. As the resin

and solvent solution is heavier than the microsphere, it tends to sink when the mixing

action stops. This also promotes separation of the mixture into phases of equal density by

increasing the tendency of microspheres to float to the top of the mixture, and the resin

solution to sink to the bottom; the mixture must be mixed continuously to the point that

it becomes stable [126]. Other problems include the potential environmental hazards and

Page 72: Elastic And Mechanical Properties Of Expanded Perlite And ...

48

health effects of volatile solvents, the formation of structural non-uniformities as a result

of the solvent’s removal by heat, the difficulty in producing batch-to-batch homogenous

syntactic foams owing to difficulty in the complete removal of excess solvent from the

mixture before moulding, and the additional cost and energy associated with removal of

the solvent and transporting the mixture to the moulding or curing equipment [127, 128].

The processing of syntactic foams from liquid thermosetting resins without using solvent

has also been reported. This method has been found to be suitable for liquid polyesters

and silicones [113, 129]. In this method, solid particles are directly mixed with a heat

curable thermosetting resin, followed by casting into a desired shape and curing to a foam

[129-131]. The difficulty in this method is non-uniform distributions of particles in the

matrix. To mitigate this problem, the mixture is heated to allow thermosetting resin to

flow and to wet the particles and, hence, a more uniform distribution of particles

throughout the matrix is obtained. In this method, it is important to keep the viscosity as

low as possible to prevent the creation of air bubbles in the final product. Due to the

increase in viscosity, the volume fraction of particles is limited (i.e. the higher the particle

concentration, the higher the viscosity) and hence low density foams (ca. 0.3 g/cm3) can

hardly be achieved. On the one hand, as the particle volume fraction increases, their

tendency to float to the surface of epoxy decreases, which is a common problem in a

simple mixture of epoxy and particles. It has been found that a volume fraction of between

35% to 65%, and preferably about 50%, is suitable to minimise the tendency of the

mixture to separate into discrete phases [129]. Yet the volume fraction of hollow spheres

is adjusted based on the composition, size, and shape of the particles.

When the resin is available in a powdered form, a solid mixture of hollow spheres and

powdered thermoplastic (or powdered thermoset) is prepared by gentle shaking and

Page 73: Elastic And Mechanical Properties Of Expanded Perlite And ...

49

agitating actions [113, 132, 133]. In some cases, a suspending agent is used to facilitate

mixing and is evaporated before moulding [113]. A weighted quantity of the mixture is

then charged into a mould and heat pressed at a temperature range within which the resin

powder melts and curing occurs. The volume fraction of particles is constant for each

sample with the purpose of completely filling the mould with the closest particle packing

possible. By holding the particle volume fraction constant, the density can be dependent

on a single independent variable (either the resin or the voids volume fraction).

Thermoplastic syntactic foams can be manufactured by conventional melt processing

techniques, such as extrusion or injection moulding [117, 134, 135], though many larger

particles may be damaged by the high stresses and hydrostatic pressures associated with

melt processing. Depending on the type of resin, sometimes post-curing after the

moulding operation is required.

Another manufacturing method for syntactic foams consists of resin coating, vacuum

filtration and polymer precipitation [136, 137]. In the coating step, particles are

introduced into a resin solution (e.g. epoxy diluted with acetone) and left there for a

measured amount of time to ensure good adsorption of resin on the surface of the

particles. The mixture is subsequently vacuum filtered and rinsed with liquids while on

the filter in order to precipitate epoxy on particle surfaces and to remove the solvent

simultaneously by leaching. The result is a wet-sand like material which forms into finely

divided coated particles after drying. These particles have a free-flowing behaviour and

are used as a moulding powder. The resin film to particle ratio in the dry powder is

determined by factors such as solution concentration (i.e. dilution ratio), particle/resin

solution ratio, time of contact between the resin solution and the particles, and

temperature. A predetermined quantity of dry coated spheres, calculated to produce a

Page 74: Elastic And Mechanical Properties Of Expanded Perlite And ...

50

close-packed syntactic foam, is loaded into the mould cavity, pressed to a desired volume,

and cured. This method enables the production of different sample densities by changing

the resin concentration while the particles are closely packed in a mould [2].

Syntactic foams have also been produced using spray-up equipment. In this method, resin

(or resin solution) and particles are sprayed using separate adjustable streams, meet and

mix at some point while still in the air and accumulate as a syntactic foam on any desired

surface. In this method, the density of the foam is controlled by changing the flow rate of

each component [113, 138].

Last, but not least, is manufacturing syntactic foam based on the microsphere buoyancy

disclosed in AU Patent No. 2003205443 [139]. This technique involves mixing

microspheres with diluted epoxy resin. The mixture is subsequently left stationary for the

self-packing of microspheres by buoyancy resulting in two phases, i.e. the top phase is

composed of packed microspheres and diluted binder; and the bottom phase contains only

diluted binder. The bottom phase is then drained from the bottom of the mould and the

remaining top phase is allowed to be cured. This method has the advantage of ease of

mixing of the microspheres with the epoxy.

2.3.3 Particulate composites containing naturally occurring fillers

Though these hollow particles provide the syntactic foams with a high strength to weight

ratio, it is of interest to find a material which provides similar properties but which is

significantly cheaper and has a lower carbon footprint. One type of promising material is

naturally occurring volcanic glasses.

There are several studies which have investigated the mechanical behaviour of

composites containing naturally occurring inorganic porous aggregates in a resinous

Page 75: Elastic And Mechanical Properties Of Expanded Perlite And ...

51

matrix. One naturally occurring volcanic aggregate is pumice, which is mainly comprised

of silica (~ 60 wt%) and alumina (~ 16%) [140]. Pumice is formed during volcanic

eruptions when molten rock is ejected into the air. The rapid depressurisation due to the

violent expulsion of magma into the atmosphere causes the release of dissolved gases and

the formation of magma froth. The simultaneous cooling and depressurisation freezes the

bubbles in the matrix and turns the frothy-like magma into a porous ceramic [141, 142].

Fleischer and Zupan [142] used these aggregates in two sizes (4mm and 8mm) in a matrix

of epoxy resin to manufacture a rigid cellular core material. Compressive tests were

conducted and the mechanical response of the pumice–epoxy structure as a function of

relative density was evaluated. However, due to the higher density of pumice (i.e. 1.056

- 1.075 g/cm3), the pumice/epoxy composites were quite dense. Sahin et al. [143, 144]

manufactured PPS (Polyphenylene sulphide) composites reinforced with pumice powder

(<45 µm) at different mass concentrations (1, 5, and 10 wt%) and obtained considerable

improvement in the thermal and mechanical properties of PPS reinforced with 1wt%

pumice powder. It showed increases of 5.6%, 9.8% and 22.5% in the glass transition

temperature, tensile strength and relative degree of crystallinity, respectively. Ramesan

et al. [145] prepared a composite with various loadings of pumice powder (~ 60 µm)

added to a polyvinyl alcohol (PVA) - polyvinyl pyrrolidone (PVP) blend, which is a

biodegradable, water-soluble polymer, and for composites containing 10 wt% pumice

powder they obtained higher AC electrical conductivity, dielectric constant and dielectric

loss, thermal decomposition temperature, and of particular relevance, a higher tensile

strength (~ 29.4 %) than the pure PVA/PVP blend. Alvarado et al. [146, 147] reinforced

PHBV, which is a naturally occurring biopolymer, with 12.8 wt% pumice particles

(~ 147μm) and manufactured PHBV-pumice composites which were stiffer (12.35%)

Page 76: Elastic And Mechanical Properties Of Expanded Perlite And ...

52

and more brittle than PHBV, experiencing almost no plastic deformation, in order to

reduce the cost and density of the current PHBV-based materials for structural and non-

structural applications. Examples of other studies focused on pumice-reinforced

composites can be found in [148-151].

Another mineral aggregate of natural occurrence is vermiculite. Vermiculite resembles

mica in appearance and mainly consists of silica (37 - 42 wt%), magnesium oxide (14 -

12 wt%), alumina (10 - 13 wt%), ferric oxide (5 - 17 wt%), water (8 - 18 wt%) and a

lesser amount of ferrous oxide (1 - 3 wt%) [152]. When subjected to high temperatures

(>900° C), the water situated in the interlayer space is converted to steam which causes

expansion of vermiculite particles to twenty or thirty times their original volume, giving

so-called exfoliated or expanded vermiculite. Consequently, a highly porous material that

is highly thermally insulating is formed which finds use in various industrial applications

[153, 154]. Jun et al. [155] used the fire retardant properties of expanded vermiculite, in

combination with the low thermal conductivity of phenolic resin, in order to develop

flame retardant insulating composites with a density range 0.12 - 0.22 g/cm3 and thermal

conductivity of about 0.045 - 0.069 W/(mK). Verbeek and Du Plessis [156] used

expanded vermiculite particles in combination with phenol formaldehyde resin and

phosphogypsum to manufacture composites with a density as low as 0.986 g/cm3 and

flexural strength as high as 3.04 MPa for building applications. Yu et al. [157] treated

expanded vermiculite powders with benzyldimethyl-octadecyl-ammonium (ODBA) and

mixed it with phenolic resin (98 wt%) to form a composite with better thermal properties

than that of pristine phenolic resin. The thermal decomposition temperature of the

phenolic/treated expanded vermiculite composite reached 482.6 °C in air, which was 48.7

°C higher than that of pristine phenolic resin. Zheng et al. [158] modified natural

Page 77: Elastic And Mechanical Properties Of Expanded Perlite And ...

53

powdered vermiculite (<54 µm) by cation exchange with quaternary phosphonium salts

and blended it with polyethylene terephthalate (PET), which is a semicrystalline

thermoplastic polymer, to form (PET)/phosphonium vermiculite composites with better

mechanical properties than pure PET. The (PET)/phosphonium vermiculite composites

containing 3% modified vermiculite showed higher tensile strength (17.4%) and stiffness

(35%) than pure PET. Patro et al. [159] dispersed natural powdered vermiculite (<25

µm) in either isocyanate or polyether based polyols to synthesise rigid

polyurethane/vermiculite composites and achieved better mechanical and thermal

properties than those of a rigid polyurethane foam without vermiculite powder. When 2.3

wt% vermiculite was dispersed in isocyanate, the polyurethane/vermiculite

nanocomposites showed increases of 40% and 34% in their compressive strength and

modulus, respectively, and a 10% decrease in their thermal conductivity. Qian et al. [160]

modified natural powdered vermiculite (<4µm ) by cation exchange with long-chain

quaternary alkylammonium salts and then dispersed it in polyether based polyols to

synthesise a polyurethane/vermiculite nanocomposite with better mechanical and thermal

properties than pure polyurethane foams having 30% rigidity. The composites containing

5.3 wt % vermiculite showed a >270% increase in tensile modulus, >60% increase in

tensile strength, and a 30% reduction in N2 permeability than pure polyurethane foams.

There are several other studies in the literature focused on vermiculite-reinforced polymer

composites with the same general results [161-166].

Last, but not least, another example of a naturally occurring mineral is volcanic ash.

Volcanic ash is made of tiny fragments of jagged rock, minerals and volcanic glass which

is created during volcanic eruptions and deposited at the surface. These deposits are

abundant, readily accessible, and are rich in silica and alumina. Volcanic ash is recognised

Page 78: Elastic And Mechanical Properties Of Expanded Perlite And ...

54

as a mesoporous material having significant porosity, an appropriate pore structure and

high surface area, which increases the effect of surface adhesion when they are used as

fillers in composites [167]. Bora et al. [168, 169] treated volcanic ash particles with 3-

aminopropyltriethoxysilane (3-APTS) at various concentrations (1, 3 and 5 vol.%) and

used these particles at two concentrations (i.e. 10% and 15%) to reinforce PPS

composites. Treated volcanic ash/PPS composites which showed considerable

improvement in their tensile and flexural properties with 3 vol.% coupling silane agent in

comparison to untreated volcanic ash/PPS composites. Trinidad et al. [170] manufactured

two polymeric matrix composites by mixing volcanic ash with epoxy resin and polyester

resin, individually, and achieved slightly better mechanical properties for volcanic

ash/polyester than volcanic ash/epoxy composites. The volcanic ash/polyester composites

showed 9%, 2.3%, 5.6% and 6.5% higher compressive strength, hardness, modulus of

elasticity, and modulus of rupture, respectively, than those of the volcanic ash/epoxy

composites. Fidan [171] reported manufacturing polyvinyl chloride (PVC) composites

reinforced with volcanic ash at various concentrations (5 - 25 wt%) which resulted in

better thermal stability for PVC composites within the temperature range 25 - 600 °C.

Bora [172] fabricated PVC/volcanic ash composites by reinforcing PVC with volcanic

ash at various mass contents (5 - 25 wt%) which led to an increase in the tensile and

flexural modulus but a decrease in the tensile and flexural strength. All the properties

showed significant reduction as a result of the temperature increase from -10°C to 50 °C.

Avcu et al. [173] outlined the manufacture of volcanic ash filled polyphenylene sulphide

(PPS) composites containing different volcanic ash concentrations (2.5 - 20 wt%). They

reported improvement in the thermal, mechanical and residual mechanical properties of

the PPS composite by an increase in the volcanic ash content, although erosion resistance

Page 79: Elastic And Mechanical Properties Of Expanded Perlite And ...

55

was decreased markedly. Cernohous [174] outlined the melt processing of polypropylene

(PP) with a blowing agent and volcanic ash particles at different mass concentrations (40

- 60 wt%) and produced thermoplastic foamed composites with a density range 0.65 -

0.78 g/cm3 and a flexural modulus in the range 3350 - 4120 MPa. Examples of other

studies which have focused on composite materials made from a polymeric matrix

reinforced with volcanic ash can be found in [175-180]. There are, however, few studies

investigating the use of perlite particles in a resinous matrix. Lukosiute et al. [181]

produced two composites, one was made of an epoxy resin matrix filled with unexpanded

perlite particles at different volume fractions and another one made of an epoxy resin

matrix reinforced with different volume fractions of plasticised PVC particles and

unexpanded perlite particles. The unexpanded perlite particles used in this study had a

mean size of 80 µm, specific weight of 2.37 g/cm3, density of 1.60 g/cm3 and a porosity

of 32%. Perlite/epoxy composites have shown higher tensile strength, elastic modulus

and bending strength but lower impact strength than those containing dispersed PVC

particles. In addition, it was found that an increase in the perlite particle volume fraction

from 7% to 41% increased the stiffness of the perlite/PVC/epoxy composites due to the

higher stiffness of the perlite particles (i.e. 27.5 GPa [181]) compared with the epoxy

matrix (i.e. 2.6 GPa). Sherman and Cameron [182] produced cores for building boards

containing a high proportion of expanded perlite particles (85% by weight) in different

matrices of acrylic resin, epoxy resin and urea-formaldehyde. They found that epoxy resin

provided the core with a higher modulus of rupture and density than the other two resins.

However, the results were not comprehensive as their conclusions were only based on the

two above-mentioned properties. Of all of the available options, perlite is considered the

most promising filler.

Page 80: Elastic And Mechanical Properties Of Expanded Perlite And ...

56

There are several studies which have investigated the use of EP particles in composites

for structural and non-structural applications. Lu et al. [183] manufactured a form-stable

composite by direct impregnation of expanded perlite in paraffin, resulting in good

thermal energy storage, thermal stability and thermal reliability for composites containing

60 wt% paraffin. Later, Lu et al. [184] manufactured a phase change material by

depositing graphene oxide films on the surface of the expanded perlite/paraffin

composite. The heat storage/release performance test results showed that the heat

storage/release rate of the expanded perlite/paraffin composite with 0.5 wt% graphene

oxide was twice as fast as that of the expanded perlite/paraffin composites because of the

enhanced thermal conductivity. Karaipekli et al. [185] also manufactured a form-stable

expanded perlite/paraffin composite material for latent heat thermal energy storage and

found that adding 5 wt% expanded graphite could increase the thermal conductivity of

the form-stable composite by about 46%. Shastri and Kim [186] used a pre-mould

process, which was initially used by Kim [187] in manufacturing glass

microsphere/epoxy syntactic foams and disclosed in AU Patent No. 2003205443 [139],

for consolidation of expanded perlite particles with starch binder as a building material.

Using this method, they manufactured perlite/starch composites in the density range 0.1

- 0.4 g/cm3 and characterised the compressive strength and modulus to be in the range 0.1

- 1.5 MPa and 10 - 85 MPa, respectively, for this range of foam density. Arifuzzaman and

Kim [188] also produced composite foams for building applications using a similar

method of manufacture, based on the buoyancy principle by dispersing expanded perlite

particles in a matrix of sodium silicate. They manufactured these foam composites in the

density range 0.2 - 0.5 g/cm3 and characterised the compressive strength and modulus to

be about 2.8 MPa and 175 MPa, respectively, for the highest foam density containing

Page 81: Elastic And Mechanical Properties Of Expanded Perlite And ...

57

0.35 g/ml sodium silicate. Furthermore, they investigated the effect of expanded perlite

particle size in three ranges, 1 - 2mm, 2 - 3mm and 3 - 4mm, on the compressive strength

of the perlite/sodium silicate foams and found that the particle size did not significantly

affect the compressive strength.

The light weight and porous structure of expanded perlite has also received attentions in

manufacturing syntactic metal foams. Taherishargh et al. [11, 189, 190] produced low

density syntactic foams as an effective energy absorber by the counter-gravity infiltration

of expanded perlite particles with molten aluminium. They investigated the effect of

expanded perlite particle size, particle shape and foam density, on the resulting properties

of foams containing 62 - 65 vol% expanded perlite. Smaller expanded perlite particles

resulted in a metallic foam with a refined cell wall microstructure, smaller cell size, fewer

defects and an equiaxed dendritic morphology. In addition, the homogeneity of the cell

geometry was found to increase with a decrease in particle size, resulting in more uniform

plastic deformation. As a result, foams with smaller EP particles showed a smoother and

steeper stress–strain curve and a higher plateau stress and energy absorption capacity.

Regardless of particle size, the mechanical properties (e.g. compressive strength, modulus

of elasticity, energy absorption efficiency and plateau end strain) were found to increase

with an increase in the foam density. This has also been reported in the literature for a

variety of metallic foams [191-197]. The shape of expanded perlite particles was also

found to affect the structural and mechanical properties of expanded perlite/aluminium

syntactic foam [190]. The irregular shape of raw EP particles was turned into nearly

spherical particles using a rotary tumbling machine. Use of rounded EP particles resulted

in an increase in the modulus of elasticity, 1% offset yield stress, plateau stress and energy

absorption by 38%, 24%, 14% and 19%, respectively. Image processing and

Page 82: Elastic And Mechanical Properties Of Expanded Perlite And ...

58

micro-computed tomography (µCT) revealed strut misalignment, missing cell walls, and

surface roughness inside the pores for foams containing the raw particles, while a

homogenous distribution of nearly spherical pores was observed for foams with rounded

particles. Hence, the superior mechanical properties of the foams with rounded EP

particles were ascribed to their homogenous distribution of pores and fewer structural

defects.

In summary, EP perlite is considered to be a promising material in manufacturing polymer

matrix syntactic foams with regards to its properties (i.e. chemical and physical) and

applications, as explained in the first part of this chapter (see Section 2.2). Perlite, in its

expanded and unexpanded forms, has been used in the manufacture of different

composites; however very few studies have been conducted on the combination of EP

particles with a resinous matrix, especially epoxy. Hence, there is a need for further

investigation of the properties of foams made from perlite particles in a resinous matrix.

Page 83: Elastic And Mechanical Properties Of Expanded Perlite And ...

59

Motivation and Problem Statement

Despite the fact that hollow microsphere/epoxy syntactic foams are known to offer

desirable properties for structural applications (i.e. high strength to weight ratio and high

stiffness to weight ratio), the high cost of the synthetically made filler particles makes

their production expensive. Considering the low density, abundance and low cost of

expanded perlite particles, it would be beneficial to develop new light-weight foam cores

for sandwich structures to replace conventional hollow microsphere/epoxy syntactic

foams at a significantly lower price. In addition, considering the wide use of expanded

perlite particles in different sectors of construction and industry, it would be worthwhile

to investigate the elastic properties of these particles, which have received little attention

in previous research.

Page 84: Elastic And Mechanical Properties Of Expanded Perlite And ...

60

Research Objectives and Research Significance

The main objectives of the current study are as follows:

i. To investigate the mechanical properties of packed beds of expanded perlite

particles.

ii. To introduce suitable models capable of predicting the elastic properties of

packed beds of expanded perlite particles.

iii. To introduce and characterise an economical, light-weight syntactic foam.

iv. To investigate the effects of different parameters such as the filler particle sizes

and foam density on the structure, microstructure and mechanical properties of

the new syntactic foams.

v. To investigate damage mechanisms that occur during mechanical tests and

understand the material’s behaviour under applied loads.

The significance of the current research study are as follows:

i. Producing a syntactic foam having properties comparable with other foams

available in the market but significantly cheaper.

ii. Producing a syntactic foam which can be moulded to accommodate different

shapes for different structural applications.

Page 85: Elastic And Mechanical Properties Of Expanded Perlite And ...

61

3 Chapter Three: Methodology

Introduction

This chapter begins with an introduction to the materials used in this study, followed

by discussion of how these materials were used to prepare packed EP particle beds,

sintered solid perlite, resin and EP/epoxy foam samples. Subsequently, all of the

different tests used to characterise the sample properties (e.g. mechanical and elastic

wave speeds) are explained in detail. Furthermore, the apparatus and procedure used

for the microstructural and damage analysis are explained. Lastly, the theory of

dynamic moduli measurement using the elastic wave speed is described.

Page 86: Elastic And Mechanical Properties Of Expanded Perlite And ...

62

Material

3.2.1 Expanded perlite particles

Expanded perlite particles (EP), shown in Figure 2.1 (b), were supplied by Industrial

Processors Limited (INPRO) and were sieved into the size ranges 1 - 2 mm, 2 - 2.8 mm

and 2.8 - 4mm. The chemical composition of the perlite provided by the supplier is

presented in Table 3.1.

Table 3.1. Chemical composition [198]

Constituent Percentage present

Silica 74%

Aluminium Oxide 14%

Ferric Oxide 1%

Calcium Oxide 1.3%

Magnesium Oxide 0.3%

Sodium Oxide 3.0%

Potassium Oxide 4.0%

Titanium Oxide 0.1%

Heavy Metals Trace

Sulphate Trace

Page 87: Elastic And Mechanical Properties Of Expanded Perlite And ...

63

3.2.2 Epoxy resin

The resin system used for binding perlite particles consisted of West System 105 Epoxy

Resin with 206 Slow Hardener, at a ratio of 5.36:1 by weight. To enable the

manufacturing method, ease the distribution of perlite particles in the binder and control

the binder content in the foam, the epoxy was diluted with acetone (Septone, ASA20).

Acetone was considered to be a suitable solvent as it has a less detrimental effect on the

mechanical properties of cured epoxy in comparison with other solvents such as toluene,

tetrahydrofuran, N-dimethylformamide (DMF) and ethanol [199, 200]. Fourier transform

infrared spectroscopy (FTIR) also showed that the use of acetone does not seem to change

the chemical structure of the liquid epoxy [199, 201]. The dilution with acetone at this

ratio reduced the mix viscosity, which was measured using a RST-Brookfield rheometer.

In addition, dilution with acetone increased the curing time of the epoxy, which was

investigated by monitoring hardness change using a Type D durometer, as shown in

Figure 3.1.

Figure 3.1. Durometer (Type D) for hardness test.

Page 88: Elastic And Mechanical Properties Of Expanded Perlite And ...

64

Sample preparation

3.3.1 Preparation of expanded perlite (EP) particle samples

3.3.1.1 Samples for structural characterisation of EP particles

The tapped density of EP particle beds of known mass were measured using a tapping

device with a graduated measuring cylinder of 100 ml, shown in Figure 3.2. After every

20 taps, the cylinder was rotated to minimise any possible separation of the mass during

tapping down. Five hundred taps were conducted for each density measurement, and the

average of five measurements was considered.

Particle density was also measured using the wax-immersion method (ASTM C914-95).

It should be noted that due to the EP particles’ structure having both open and closed

pores, measurements with an air pycnometer resulted in the wrong density. As expected,

the value given by the air pycnometer lay between the particle density and the true density

of the material. As the wax does not penetrate into the perlite pores, the wax-immersion

method is considered to give the most accurate measurement of the perlite particle

density.

Scanning electron microscopy was used to study the internal structure of perlite particles

by cross–sectioning EP particles carefully. For this purpose, a number of particles were

embedded in epoxy resin poured in a small mould of 30 mm diameter and 30 mm height.

Samples were ground and lapped to a 1m diamond finish. The ground samples were

transferred into an ultrasonic cleanser to remove perlite dust created during the cutting

and polishing. All samples were coated with gold before microscopy to avoid charging.

Page 89: Elastic And Mechanical Properties Of Expanded Perlite And ...

65

Figure 3.2. Tapping device for measuring tapped density of EP particles.

3.3.1.2 Samples for confined quasi-static compressive tests (oedometer tests)

A cylindrical compaction mould made of steel was manufactured to produce samples 35

mm in height and 30 mm in diameter for measuring the properties of packed EP particle

beds. Figures 3.3 (a) and (b) show the constituent parts of the mould and the assembled

form of the mould, respectively. Samples for each particle size were prepared for the

density range 0.10 - 0.40 g/cm3 and three samples were prepared for each density. To this

end, EP particles of known mass were transferred into the mould and compressive loading

was transferred to the EP particles through a plunger moving vertically into the mould.

The compaction process was carried out with a constant displacement rate of 1.0 mm/min

in a computer-controlled 5kN Shimadzu testing machine. The target compaction densities

were achieved by controlling the mass of EP particles within a constant volume. To ensure

homogeneity, specimens were compacted in three layers. To minimise friction between

Page 90: Elastic And Mechanical Properties Of Expanded Perlite And ...

66

the piston and the wall of the mould, lubricant oil was sprayed on the wall of the mould

and the plunger.

(a) (b)

Figure 3.3. Prepared mould (a) The different parts of the mould (b) Assembled mould.

3.3.1.3 Samples for elastic wave tests on packed beds of EP particles

An aluminium cylindrical compaction mould was manufactured to produce samples of 45

mm ± 0.015 mm in height and 80 mm in diameter for measuring the elastic properties of

packed beds of EP particles. The target compaction densities were achieved by controlling

the EP particles’ mass within a constant volume. To ensure homogeneity and minimise

variations in density, the specimens were compacted in several layers (up to five). The

compaction process was carried out in a 5000kN computer-controlled Shimadzu testing

machine using a constant displacement rate of 1.0 mm/min. Lubricant oil was used to

minimise friction between the piston and the walls of the mould. Samples were prepared

with a density range 0.10 - 0.35 g/cm3 and three samples were prepared for each density.

Page 91: Elastic And Mechanical Properties Of Expanded Perlite And ...

67

3.3.2 Preparation of solid perlite samples

Solid perlite was prepared by grinding 27 g of perlite particles with a mortar and pestle

and sieving to <250 µm. The powder was transferred into a mould and uniaxially pressed

into a pellet at 146 MPa. The pellets, resting on a bed of alumina powder, were heated at

10 °C/min to the sintering temperature of 1100 °C which was held constant for 12 hours.

Subsequently the samples were cooled at 10 °C/min to 450 °C and held for an hour.

Cooling continued at the same rate to 300 °C, at which temperature the furnace was turned

off and the sample was left to cool slowly. The sintered pellet was cut and polished into

samples 30 mm in diameter and 16 mm high, as shown in Figure 3.4. The densities of the

prepared sintered pellets were measured at room temperature using the Archimedes

method.

Figure 3.4. Solid perlite sample (sintered perlite powders).

3.3.3 Preparation of resin samples

To measure the compressive, tensile and flexural properties of epoxy resin, two groups

of samples were prepared, one group for the undiluted epoxy resin and another group for

epoxy resin diluted with acetone according to the ratio of 1 to 7 for epoxy + hardener to

Page 92: Elastic And Mechanical Properties Of Expanded Perlite And ...

68

acetone, respectively. Samples were prepared by pouring the epoxy resin into two moulds

made of Acetal, as shown in Figure 3.5.

(a) (b)

Figure 3.5. Mould for manufacturing epoxy resin samples for (a) compression and flexural tests

and (b) tensile tests.

The moulds, shown in Figure 3.5, were designed to manufacture samples according to

ASTM D-695, ASTM D-638 and ASTM D-790 standards for measuring the compressive,

tensile and flexural properties of epoxy resin, respectively, as shown in Table 3.2.

Table 3.2. ASTM D-695, ASTM D-638 and ASTM D-790 for measuring the compressive,

tensile and flexural properties of epoxy resin.

ASTM

D-695

Page 93: Elastic And Mechanical Properties Of Expanded Perlite And ...

69

T (Thickness) = 3.2±0.4 mm, W (Width of narrow section) = 12.7 mm,

LO (Length Overall) = 79.4 mm, WO (Width Overall) = 19 mm,

G (Gage length) = 38.1 mm, R (Radius of fillet) = 38.1 mm

ASTM

D-638

T (Thickness) = 3.2±0.4 mm, W (Width of narrow section) = 13 mm,

L (Length of narrow section) = 57 mm, WO (Width Overall) = 19 mm,

LO (Length Overall) = 165 mm, G (Gage length) = 50 mm, D (Distance between

grips) = 115mm, R (Radius of fillet) = 76 mm

ASTM

D-790

T (Thickness) = 3.2±0.4 mm, W (Width of narrow section) = 12.7 mm,

LO (Length overall) = 79.4 mm

Page 94: Elastic And Mechanical Properties Of Expanded Perlite And ...

70

3.3.4 Preparation of EP/epoxy foam samples

The basic principle for manufacturing EP/epoxy foam composites was based on the

buoyancy method explained in AU Patent No. 2003205443 [139] for manufacturing glass

microsphere/epoxy syntactic foams. The perlite particles provided by Industrial

Processors Limited (INPRO) contained some unexpanded and semi-expanded particles;

thus a device was designed to separate the expanded particles, based on the winnowing

technique. This device, shown in Figure 3.6, was made of two computer fans, which

provided enough air pressure with respect to the density of EP particles, a sieve of 500

µm and a piece of PVC pipe, which was attached to the sieve. The air blown by the fans

passed through the sieve mesh and lifted the less dense expanded perlite particles out of

the PVC pipe, leaving the denser unexpanded particles behind. Separated EP particles,

the epoxy + hardener and acetone were mixed at the ratio of 1:2:14 respectively, in a

screw cap plastic container through gentle agitating and tumbling actions. This particular

mixing ratio was adopted due to the buoyancy method and to producing light-weight

foams containing a thin film of epoxy around the particles. The mixing container was

then left stationary until particles floated to the surface resulting in two phases, i.e. the

top phase of packed perlite particles and diluted binder, and the bottom phase containing

only diluted binder, as shown in Figure 3.7. The top phase was separated by straining and

transferred into a cylindrical steel mould which is shown in Figure 3.3. The same mould

was used in preparing packed beds of EP particles (explained in Section 3.3.1.2) and

EP/epoxy foam samples in order for the dimensions to be comparable and to correctly

measure the crushing strength of the EP particles during sample manufacturing (to be

explained later in Section 4.6.2). The mould was designed to produce samples 35 mm

high and 30 mm in diameter, according to ASTMC365/C365M – 11a.

Page 95: Elastic And Mechanical Properties Of Expanded Perlite And ...

71

Figure 3.6. A hand-made device for separating expanded particles from unexpanded and semi-

expanded particles.

Three types of EP/epoxy foam samples were manufactured for the density range 0.15 -

0.45 g/cm3 by changing the mass inside the mould while keeping the volume constant.

To achieve a constant volume, a universal testing machine was used to compress the

mixtures of EP particles and the binder to the target height of 35mm. Figure 3.8 illustrates

the foam samples having the same density but using three different particle size ranges.

Foams containing one particle size range were considered to be of one type. Foams of

type 1, type 2 and type 3 were defined as foams made with particles in the size ranges of

1 - 2 mm, 2 - 2.8 mm and 2.8 - 4 mm, respectively. Five to eight samples were

manufactured for each density and type of foam. The densities of the prepared foams were

measured at room temperature after the curing time elapsed.

Page 96: Elastic And Mechanical Properties Of Expanded Perlite And ...

72

Figure 3.7. Formation of two phases by buoyancy. The top phase is packed perlite particles and

diluted binder and the bottom phase contained only diluted binder.

Figure 3.8. Samples prepared using three particle size ranges; from the left, 1 - 2 mm, 2 - 2.8

mm and 2.8 – 4 mm.

Page 97: Elastic And Mechanical Properties Of Expanded Perlite And ...

73

Experimental Setup and Tests

3.4.1 Mechanical testing on the packed beds of EP Particles

To measure the properties of the EP particles, oedeometric tests (confined compressive

tests) were conducted on packed EP particle beds. As explained in Section 3.3.1.2,

compaction was conducted in several layers, where each layer had virtually the same

density and almost the same compaction load. Hence, the compaction load was in the

same range for the whole of each sample. During the tests, experimental data were

recorded by built-in aquisition software (Trapezium 2) in the form of force-displacement

curves for each sample. These were subsequently converted to compaction stress-density

curves based on the cross-sectional area of the plunger and the final height of the sample,

measured using callipers after removing the compacted sample from the mould.

In addition, the compacted samples containing EP particles in the range 2 – 2.8 mm were

used to measure the confined elastic modulus of the particles by cyclic loading-unloading

of the pre-compacted particles to 10%, 15% and 20% of the maximum compaction force.

Three to four tests were conducted for each density.

3.4.2 Mechanical testing of EP/epoxy foams

Uni-axial (unconfined) compression tests were conducted on a computer-controlled 5 kN

Shimadzu testing machine. Samples were compressed at a crosshead speed of 0.5

mm/min at room temprature, which ranged from 23 to 25°C. During the tests,

experimental data were recorded by the built-in aquisition software (Trapezium 2) in the

form of force-displacement curves for each sample and these were then converted to

engineering stress-strain curves, based on the intial cross-sectional area and height of the

Page 98: Elastic And Mechanical Properties Of Expanded Perlite And ...

74

test sample. To investigate the porosity dependent rigidity, the elastic gradient of the

material was measured based on ISO 13314. In this standard, the elastic gradient is

determined by loading and unloading between 20% and 70% of the maximum stress (σ20

and σ70). The unloading gradient, or so-called effective elastic modulus, was measured to

alleviate two potential problems affecting the accuracy of the initial loading modulus.

Firstly, there might be some localised plasticity in the sample where cells yield at very

low stress levels. This may result in contributions from both elastic and plastic

deformation to the resulting modulus. Secondly, it is very likely that the sample’s surfaces

touching the loading plates are not perfectly parallel. However, applying load at low

strains will flatten the surfaces until complete contact is established (settling).

Consequently, when the load is released, the unloading gradient is purely elastic and not

affected by localised plastic deformation and the settling effect [11]. To this end, one

initial sample must be tested to obtain the maximum compressive stress (max). Then the

next four samples were loaded to 70% of the maximum stress (σ70) and unloaded to 20%

of the maximum stress (σ20).

Furthermore, a series of confined compression tests was conducted on cured EP/epoxy

foams. For this purpose, samples were tightly enclosed in the cylindrical steel mould

shown in Figure 3.3, with an inner diameter equal to the diameter of the sample. This

mould was the same mould used for manufacturing the EP/epoxy foams. The diameter of

the samples after curing were slightly bigger (30.1 ± 0.045 mm) than before curing (30

mm) which helped in tightly enclosing the foams.

Page 99: Elastic And Mechanical Properties Of Expanded Perlite And ...

75

3.4.3 Elastic wave tests on EP particles

To measure the properties of EP particle compacts, a compaction rig was designed to

include transducers and particles. It consisted of a pair of aluminium rams located inside

the top and bottom of the cylindrical aluminium die enclosing the compacted EP particles.

The rams were designed to have the same height as the transducer and a groove held the

transducer tightly with the ram surface flush with the active element of the transducer. A

schematic representation of the set-up is shown in Figure 3.9.

Figure 3.9. Schematic representation of the experimental set-up for measuring wave velocity in

packed beds of EP particles.

Elastic wave velocity measurements were carried out at a constant contact pressure equal

to the average compaction stress at each density (see Section 3.4.1) using a computer-

controlled Shimadzu testing machine. Two pairs of piezoelectric transducers (S-wave and

P-wave) with a nominal frequency of 1MHz were used to measure the compression, Cp,

and shear wave, Cs, velocities in the axial direction throughout the sample. Each pair of

transducers was placed in the rams and a thin layer of silicone grease was used to improve

Page 100: Elastic And Mechanical Properties Of Expanded Perlite And ...

76

the coupling between the specimen and the transducers. The experimental setup employed

here is similar to the system used by Arroyo et al. for testing clayey rocks [202]. It

included a programmable function generator (Thurlby Thandar® TTi – TG4001) to

generate and transmit elastic waves as well as a digital oscilloscope (Tektronix® TDS

1001C-EDU) to acquire both input and output signals. Signal stacking was applied to

both input and output waves to minimise random noise. Sine pulses with apparent

frequency ranging from 5 to 40 kHz (100 V peak-to-peak) were used as the input signal.

Elastic wave velocities were estimated as the ratio of the travel length to the travel time.

The recorded input and output signals were used to estimate the travel time by considering

the time delay between the start of the input signal and the start of the output signal

(details explained in [203]). The travel length is the height of the samples and was

determined using callipers.

Examples of input and output signals for the low and high density EP particle compacts

are given in Figure 3.10, where the arrival times for the P-waves (tP) and S-waves (tS) are

indicated by vertical dashed lines. The arrival time tP was clearly identified as the

‘common first break time’ in the output signals, irrespective of the sample density. Two

sets of input frequencies were used to determine the arrival time tS depending on the

density of the EP particle compacts. Input frequencies up to 40 kHz were used in dense

samples, whereas good quality output signals were obtained in low density samples (<

150 g/cm3) using frequencies up to 15 - 20 kHz. A step pulse of 10 Hz was also applied

which, in combination with the sine pulses, helped to check the arrival time of the S-

wave. The determination of the arrival time tS required extra care due to the influence of

low-energy P-waves, which travel faster than S-waves, on the output signals. This

phenomenon, known as the near-field effect [202, 204], is strongly frequency-dependent

Page 101: Elastic And Mechanical Properties Of Expanded Perlite And ...

77

and may lead to misleading estimates of tS if the arrival time of the P-wave is not

available. From inspection of Figure 3.10 (b) and (d), it is clear that the arrival time of the

S-wave corresponds to the first important break time, whereas the small disturbance

(‘bump’) observed earlier in the output signals (clearer in the low density samples) is

consistent with the arrival time tP previously estimated in Figure 3.10 (a) and (c).

(a) (b)

Page 102: Elastic And Mechanical Properties Of Expanded Perlite And ...

78

(c) (d)

Figure 3.10. Examples of (a) P-wave and (b) S-wave signals in low density (0.12 g/cm3) EP

particle compacts and (c) P-wave and (d) S-waves in high density (0.3 g/cm3) EP particle

compacts.

Page 103: Elastic And Mechanical Properties Of Expanded Perlite And ...

79

3.4.4 Elastic wave tests on EP/epoxy foams and solid perlite

A frame was designed to measure elastic wave velocities in EP/epoxy foams as well as

the solid perlite samples. The frame consisted of a mobile top beam guided using a

frictionless roller bearing, a stationary bottom beam and a pair of aluminium rams within

which the sample was placed. Rams were designed to hold the transducers tightly while

keeping the active element of the transducers at the same level as their surface. A

schematic representation of the set-up is shown in Figure 3.11.

Figure 3.11. Schematic representation of the ultrasonic experimental set-up for measuring wave

velocity in solid perlite and EP/epoxy samples.

To improve the coupling between the transducer and the sample, a constant pressure was

applied by adding a 5 kg dead weight to the mobile top beam; this will in turn enhance

the quality of the output signals. Two pairs of piezoelectric transducers with a nominal

frequency of 1MHz were used to measure the longitudinal, CL, and shear wave, Cs,

velocities in the axial direction throughout the sample. Each pair of transducers was

placed in the rams and acoustically coupled by a thin layer of silicone grease. Similar to

Page 104: Elastic And Mechanical Properties Of Expanded Perlite And ...

80

the experimental setup used for measuring the elastic properties of EP particles (see

Section 3.4.3), the experimental setup included a programmable function generator

(Thurlby Thandar® TTi – TG4001), a digital oscilloscope (Tektronix® TDS 1001C-

EDU) and an ultrasonic pre-amplifier OLYMPUS® (Panametrics 5656C, with a gain of

40dB). Signal stacking was applied to both the input and output waves to minimise

random noise. Sine pulses with apparent frequency ranging from 10 - 50 kHz (100 V

peak-to-peak) were used as the input signal. Elastic wave velocities were estimated in the

same way explained in Section 3.4.3.

Page 105: Elastic And Mechanical Properties Of Expanded Perlite And ...

81

Microstructural analysis

Scanning electron microscopy (SEM) was used to study the internal structure of perlite

particles and perlite-based foam samples before undergoing compression tests. The

microstructure of expanded perlite particles was studied by carefully cross-sectioning the

particles. For this purpose, a number of particles were embedded in epoxy resin poured

in a small mould of 30 mm diameter and 30 mm height. After 24 hours curing, the sample

was demoulded and the surface of the sample was carefully ground using 240, 320, 600,

800 and 1200 grit silicon carbide papers. A smooth, mirror like surface was subsequently

achieved by polishing with 1µm diamond suspension in distilled water sprayed onto a

polishing cloth. Perlite particles and epoxy resin are both non-conducting materials; hence

the samples were coated with gold before microscopy. The internal structure of the

perlite-based foams was studied by cutting a thickness of 3mm from the sample using a

low speed silicon carbide saw. The thin slices were carefully ground by 800 and 1200 grit

silicon carbide paper. To remove perlite dust created during cutting and polishing the

sample, the thin slices were put into an ultrasonic cleaner for 10 minutes. They were then

dried using a blow dryer and left in an oven at 35 °C for 2 hours. Lastly, the thin slices

were coated with gold and were then ready for microstructural study by SEM.

Page 106: Elastic And Mechanical Properties Of Expanded Perlite And ...

82

Damage observation

Failure mechanisms of the newly developed foams were investigated using both

macroscopic and microscopic observations. During the course of the mechanical testing

experiments, a high resolution camera (Nikon D800E) was used to record an image of the

compressed sample every 10s. This helped to closely follow the damage sequence and

the failure process in the foam samples. Scanning electron microscopy was used after

finishing the test to examine the fractured samples and to determine the failure modes. As

the samples were mostly made up of perlite particles having a brittle nature, the fracture

of the foam under compressive loading generated a large amount of broken pieces of

particles (e.g. perlite dust). These loose particles can cause charging during electron

microscopic observations. Thus, the fractured surface of the foam sample was exposed to

a gentle blow of air before coating with gold. As the fractured samples had uneven

surfaces with a porous structure, coating was also conducted from different directions and

angles in several steps to avoid charging effects in the microscope.

Page 107: Elastic And Mechanical Properties Of Expanded Perlite And ...

83

The theory of dynamic moduli measurement

Elastic stress waves propagate in materials by inducing infinitesimal elastic deformation.

Hence, wave propagation equations which are based on elasticity theory can be used to

measure the elastic moduli of a material if the elastic wave velocities and densities are

measured independently [205, 206]. Two wave types propagate in extended elastic solids:

(i) compression waves in which the displacements are in the direction of the wave

propagation, and (ii) shear waves in which the displacements are perpendicular to the

direction of the wave propagation. In an isotropic linear elastic body, in which the

propagating wave does not interact with the boundary of the medium, the compression

wave or P-wave (CP) and shear wave or S-wave (CS) velocities are given by:

SC (3.1)

2

MCP

(3.2)

where M is the constrained modulus, ρ is the density of the material, λ and µ are the Lamé

constants. The Lamé constants are related to the stiffness tensor, 𝐶𝑖𝑗𝑘𝑙, by:

kljkiljlikklijijkl )]+ (+ [=C (3.3)

where 𝛿 is the Kronecker delta function defined as:

𝛿𝑖𝑗 = {0 𝑓𝑜𝑟 𝑖 ≠ 𝑗1 𝑓𝑜𝑟 𝑖 = 𝑗

, 𝑖, 𝑗 = 1, 2, 3.

Young’s modulus (E) and Poisson’s ratio (𝜈) can be expressed as a function of the density,

compression wave velocity and shear wave velocity of the material as follows:

Page 108: Elastic And Mechanical Properties Of Expanded Perlite And ...

84

22

22

2 43

SP

SP

SCC

CCCE

(3.4)

)(2

222

22

SP

SP

CC

CC

(3.5)

When the characteristic dimensions of a body are not large compared with the

wavelength, these equations no longer hold, and the velocities become dependent upon

the frequency [207]. In the case of an unconstrained cylindrical bar (i.e. uniaxial

conditions) with a radius much smaller than the wave length, the compression wave has

a different velocity called the longitudinal velocity (or rod velocity) CL, given by [207]:

ECL (3.6)

However, the stiffness that controls shear waves in cylindrical bars is the same as the one

in an infinite medium (i.e. Eq. (3.1)) [204]. Consequently, Poisson’s ratio is expressed as:

1

2

1

S

L

C

C (3.7)

In the current study, the above equations are used to characterise the elastic properties of

packed EP particle beds and sintered perlite by assuming that the pores are air-filled voids

which are randomly distributed and oriented throughout the medium. Hence, the

commonly adopted assumption that the porous solid body exhibits isotropic elastic

behaviour in a statistical sense will be adhered to.

With regards to the sample size and wave length, Eqs. (3.4) and (3.5) were used to

characterise the elastic properties of packed beds of EP particles, whereas Eqs. (3.6) and

(3.7) were used to characterise the elastic properties of sintered perlite samples.

Moreover, Eqs. (3.6) and (3.7) were used to characterise the elastic properties of

EP/epoxy foams in the longitudinal direction (i.e. z-axis) by assuming that the particles

Page 109: Elastic And Mechanical Properties Of Expanded Perlite And ...

85

were randomly distributed and oriented throughout the foam and, thus, the EP/epoxy

foams macroscopically exhibit isotropic elastic behaviour.

Page 110: Elastic And Mechanical Properties Of Expanded Perlite And ...

86

4 Chapter Four: Results

Introduction

In this chapter, the properties of solid perlite (pore free), packed EP particle beds, epoxy

resin, and EP/epoxy foams are presented. The elastic properties of low porosity solid

perlite and packed EP particle beds were characterized by means of elastic wave

propagation along the the cylinder axis of the specimens. By adopting an isotropic model,

the Young’s modulus and Poisson’s ratio were used to characterise the elastic response

of the medium. Moreover, elasticity theory was used to estimate the properties of pore

free solid perlite. Using these properties for solid perlite, porosity-elastic moduli relations

were investigated using four analytical models. The models were compared with the

experimental data from this study and their ability to estimate the elastic properties at

different porosity levels was evaluated.

Mechanical properties of cured epoxy resin (before and after dilution with acetone)

measured using quasi-static mechanical tests (according to ASTM standards) and elastic

wave tests are also provided and discussed. In addition, the results from hardness and

viscosity measurements for epoxy resin are presented and the effect of dilution on the

curing time reported.

Following the above characterisation of the resin, the compressive properties of EP/epoxy

foams characterised by quasi-static mechanical tests and the effects of density and particle

size are reported and discussed. Post compression microscopic and macroscopic

observations taken during the test are presented in order to understand the damage

Page 111: Elastic And Mechanical Properties Of Expanded Perlite And ...

87

mechanisms that occurred during the test and give additional insight into the material

behaviour under compressive loading. Moreover, based on the results of mechanical tests,

estimates for the contribution of the particle elastic modulus to the resulting foam

effective elastic modulus are shown.

The elastic properties of Ep/epoxy foams were then characterised using elastic wave tests

along the cylinder axis of the specimens in terms of Young’s modulus and the Poisson’s

ratio are presented. Subsequently, a comparison is made between the properties obtained

by mechanical and elastic wave tests. Furthermore, based on the results of the elastic

wave tests, the contribution of the constituents’ properties to the resulting foam properties

is discussed.

Page 112: Elastic And Mechanical Properties Of Expanded Perlite And ...

88

Elastic properties of sintered solid perlite

Using microscope images (shown in Figure 4.1) taken from the polished surface of the

samples and an image processing technique (MATLAB, MathWorks, MA), the porosity

of the sintered perlite samples was determined to be about 4%.

Figure 4.1. Microscope images taken from the polished surface of sintered perlite. The

polishing marks are designated on the picture in order not to be confused with porosity.

It has been shown that for porosity less than 5%, there is a linear relationship between the

elastic moduli and the porosity [208-211]. Within this range (0 - 5%), elasticity theory

successfully predicts the zero-porosity properties from the elastic moduli of porous solids,

by the following relations [212]:

Page 113: Elastic And Mechanical Properties Of Expanded Perlite And ...

89

)1(0 PEE E (4.1a)

)1(0 PGG c (4.1b)

)1(0 P (4.1c)

)1(0 PCC LLL (4.1d)

)1(0 PCC SSS (4.1e)

where )1129(18/1 0 E

3/5c

)18/(118/119/5 00 c

1)21()1()1()2(22/1 000020 EL

3/1S

and E, G, ν correspond to Young’s modulus, shear modulus, and Poisson’s ratio,

respectively, and subscript 0 refers to the zero porosity values of the properties. The

resulting zero porosity values of the density ρ0, longitudinal wave velocity CL0, shear

wave velocity CS0, Poisson’s ratio ν0, shear modulus G0 and Young’s modulus E0 of

sintered perlite are shown in row 2 of Table 4.1. For comparison, the corresponding values

from solid obsidian, measured by Manghnani et al. [213], are also presented. It can be

seen that the measured wave velocity for perlite with 4% porosity is closer to obsidian

than those for pore-free perlite. This might be due to the obsidian structure which is

typically vesicle-poor [214] (i.e. <2.5% porosity [215]), though the percentage of the

porosity was not reported by Manghnani et al. [213]. The comparison with obsidian

Page 114: Elastic And Mechanical Properties Of Expanded Perlite And ...

90

provides a benchmark for confidence in the measured values for the sintered perlite,

especially Poisson’s ratio.

Table 4.1. Elastic properties of solid component of EP particles. For comparison, the

corresponding data for obsidian by Manghnani et al. [213] are presented.

𝜌

(g/cm3)

𝐶𝐿

(m/s)

𝐶𝑆

(m/s)

𝐺

(GPa)

𝐸

(GPa)

𝜈

4% porosity Perlite 2.312 5616.44 3655.1 30.9 72.93 0.181

0% porosity Perlite 2.411 5836.75 3704.5 33.1 78.33 0.183

Obsidian [213] 2.357 5775 3608 30.7 75.16 0.180

Page 115: Elastic And Mechanical Properties Of Expanded Perlite And ...

91

Properties of packed beds of EP particles

4.3.1 Structural characterisation of EP particles

The bulk and particle densities of EP particles were measured according to the method given

in Section 3.3.1.1 and an average of five measurements for each particle size range is

presented in Table 4.2.

Table 4.2. Bulk density and particle density measurements of EP particles

particle size 1 - 2 mm 2 - 2.8 mm 2.8 - 4 mm

Bulk Density (g/cm3) 0.0703 0.0857 0.0939

Standard Deviation(g/cm3) 0.0003 0.0033 0.0007

Particle Density (g/cm3) 0.181 0.183 0.186

Standard Deviation (g/cm3) 0.016 0.010 0.004

Scanning electron microscopy (SEM) was used to characterise the microstructure of EP

particles. SEM images taken from the outer surface of an EP particle at each size range are

shown in Figure 4.2. As can be seen, the outer surface of the perlite particles is covered with

both closed and open pores and has a froth-like structure.

Page 116: Elastic And Mechanical Properties Of Expanded Perlite And ...

92

(a) (b)

(c)

Figure 4.2. SEM images showing the external structure of an EP particle: (a) in the 1 - 2 mm size

range; (b) in the 2 - 2.8 mm size range; (c) in the 2.8 - 4 mm size range.

Page 117: Elastic And Mechanical Properties Of Expanded Perlite And ...

93

(a) (b)

(c) (d)

Figure 4.3. SEM images showing the internal structure of an EP particle in the: (a) 1 - 2 mm size

range; (b) 2 - 2.8 mm size range; (c) 2.8 - 4 mm size range; (d) 2 - 2.8 mm size range, illustrating

natural reinforcement inside EP particle cells.

Examination of the different microscope images shows that the outer surface of the EP

particles in the size range 1 - 2 mm contains more open pores than the other two size ranges.

Figures 4.3 (a) - (c) reveal that the interior of expanded perlites in all three size ranges has a

three-dimensional cellular structure mainly composed of closed cells. Examination of the

Page 118: Elastic And Mechanical Properties Of Expanded Perlite And ...

94

micrographs also shows that EP particles in the size range 2.8 - 4mm are made up of

polyhedral cells, mostly possessing pentagonal dodecahedron and tetrakaidecahedron

geometries. However, the cells composing the structures of EP particles in the size range 1 -

2 mm and 2 - 2.8 mm do not appear to have a preferred geometrical shape; some are

polyhedral and some are almost spherical. Micrographs taken from EP particles in the size

ranges 2 - 2.8 mm and particularly 1 - 2 mm reveal the presence of inner cell reinforcement;

fibre-like reinforcements spanning the walls of a cell (see Figure 4.3 (d)). This phenomenon

was not observed in EP particles in the size range 2.8 - 4 mm. The governing conditions and

formation mechanism of the fibrous cell reinforcement are not known at this stage.

Examination of the micrographs taken from EP particles of each size range, cell size, wall

thickness and edge cell thickness7 were measured and are presented in Table 4.3. The

measured values show larger thicknesses for the edge cells than for the walls. This can be

ascribed to surface tension effects during the perlite expansion process which draws solid

into the cell edges and leaves a thin wall framed by a thicker edge [11, 216]. It can be seen

that the measured values are highest for particles in the size range 1 - 2 mm. This could

explain the higher crushing resistance of this particle size range, illustrated later in Section

4.6.2.1.

7 Edge thickness: as defined by Gibson and Ashby [17]

Page 119: Elastic And Mechanical Properties Of Expanded Perlite And ...

95

Table 4.3. Cell dimensions of EP particles of 1 - 2 mm, 2 - 2.8mm and 2.8 - 4 mm size range. Each

value is an average of 150 measurements.

Cell Size Wall Thickness Edge Cell Thickness

1 - 2 mm 84.497 µm 0.532 µm 3.067 µm

Standard Deviation 13.437 µm 0.100 µm 0.725 µm

2 - 2.8 mm 61.107 µm 0.512µm 1.493 µm

Standard Deviation 9.397 µm 0.092 µm 0.438 µm

2.8 - 4 mm 53.494 µm 0.516 µm 2.425 µm

Standard Deviation 8.047 µm 0.092 µm 0.472 µm

Fine et al. [217] suggested a geometric relationship given by Eq. (4.2) for calculating the wall

thickness of a glass hollow microsphere from its relative density (true density of a

microsphere divided by the density of glass).

3

1

11

g

truew

r

t

(4.2)

with 𝑡𝑤= wall thickness of the microspheres, r = average radius of the microspheres, ρtrue =

true density of the microspheres, and ρg = density of the glass. Using the results presented in

Table 4.3 and measured particle densities presented in Table 4.2, Eq. (4.2) may be modified

to adapt to the perlite morphology. Two Eqs. (4.3) and (4.4) are suggested for calculating

the ratio of the wall thickness (tw) and cell size (C) to the EP particle diameter (dp) in terms

of their relative density:

Page 120: Elastic And Mechanical Properties Of Expanded Perlite And ...

96

x

UP

EP

p

w

d

t

5.15.1 (4.3)

y

UP

EP

pd

C

5.15.1 (4.4)

where 𝜌𝐸𝑃 is the EP particle density and 𝜌𝑈𝑃 is the density of fully dense perlite. The

exponents x and y for the three particle size ranges obtained by fitting the experimental results

(Table 4.3) to Eq. (4.3) and Eq. (4.4) are presented in Table 4.4.

Table 4.4. Exponents of Eq. (4.3) and Eq. (4.4) for the three particle size ranges.

x y

1 - 2 mm 1.25 1.075

2 - 2.8 mm 1.20 1.15

2.8 - 4 mm 1.20 1.17

4.3.2 Measurement of elastic moduli of packed EP particle beds using quasi-static

mechanical tests

The elastic moduli of EP particles as a function of density were measured by conducting

several loading-unloading compressive tests on packed beds of EP particles (explained in

detail in Section 3.4.1). Considering the correlation coefficient (R2) of unloading gradients

and the deviation between them, it was possible to determine the local yield point for that

compacted sample. Below the determined yield point was considered to be the elastic region

Page 121: Elastic And Mechanical Properties Of Expanded Perlite And ...

97

where the confined elastic modulus could be measured. The results as a function of compact

density are given in Figure 4.4.

Figure 4.4. Constrained modulus of packed EP particle beds as a function of compact density.

Since the particles used in this test were confined, the calculated unloading gradient was not

identical to the elastic modulus of the packed bed of particles but a function of both their

elastic modulus (EP) and Poisson’s ratio (𝜈). Considering 휀𝑦 = 휀𝑧 = 0, the confinement

introduced by the wall of the cylinder, and 𝜎𝑦 = 𝜎𝑧, the unloading gradient (𝜎𝑥 /휀𝑥) or

constrained elastic modulus E* is defined as [218]:

)21()1(

)1(*

PEE (4.5)

There are various methods to determine Poisson’s ratio (𝜈) in the geotechnical literature. A

common one [219] is given by Eq. (4.6):

Page 122: Elastic And Mechanical Properties Of Expanded Perlite And ...

98

k

k

1 ;

245tan2

k (4.6)

where 𝜑 denotes the angle of internal friction. However, it was found that the Poisson’s ratio

(𝜈) calculated from Eq. (4.6) overestimates the real value of this coefficient. Therefore, it is

not applicable where the methods of theoretical elasticity are applied [220]. Sawicki and

Świdziński [220] proposed a method for determining the elastic moduli of particulate

materials using confined compressive tests, with additional measurement of the lateral

stresses. Although their method provides more exact solutions for the elastic moduli, it

requires the conduct of biaxial compressive tests. Performing such tests was not possible in

this study due to limitations of the testing machine. As a result, the study was extended to

measure the elastic moduli of packed beds of EP particles using elastic waves, as explained

in the next section.

4.3.3 Measurement of elastic moduli of packed EP particle beds using elastic waves

The compression and shear wave velocities of packed beds of EP particles at different

densities are shown in Figure 4.5. As can be seen, both the compression and shear wave

velocities reach a plateau for densities higher than 0.2 g/cm3. At lower densities, particles

have more space available to move or rotate. At higher densities, the relative movement of

particles is restrained as the particles reach an optimal structural arrangement (i.e. the particle

fabric). Further compression reduces pore volumes between the EP particles by fracturing

particles which produces both smaller cellular EP particles and fine platy debris such as that

shown in Figure 4.6. Because of the closed cellular EP structure, microscopic analysis

Page 123: Elastic And Mechanical Properties Of Expanded Perlite And ...

99

conducted on EP particles in a previous study [15], indicated that debris does not form inside

EP particles i.e. the debris is restricted to the inter-particle spaces.

(a)

(b)

Figure 4.5. The (a) Compression wave and (b) Shear wave velocities versus compact density.

Page 124: Elastic And Mechanical Properties Of Expanded Perlite And ...

100

Figure 4.6. Formation of platy and fine particles as a result of the brittle crushing of cell walls.

The use of static density in Eqs. (3.1), (3.2), (3.4) and (3.6) can be justified only if the debris

are mechanically linked to the particles [221]. If the debris particles are free moving, they

should not be taken into account in the evaluation of the dynamic compact density.

Consequently, further investigation of the debris was undertaken. The debris concentration

was investigated by separating the EP particles from the debris and measuring their relative

masses at each compact density. Mechanical separation was ineffective without causing more

particle fracture and debris. Therefore, the EP particle compacts were immersed in water for

24 hours to disperse. The dispersion resulted in two phases: a top phase comprising EP

particles floating on the surface and sediment. The sediment was considered to be 100%

debris due to its higher density (note the solid density of perlite in Table 4.1). Beakers

containing the EP particles and sediment were placed in an oven at 120 °C for 24 h. The

debris size was measured and found to be less than 212 µm which is used here to discriminate

between particles and debris. To be sure that all the debris had separated from the EP

particles, the samples were sieved and the fraction bellow 212 µm was added to the debris.

Figure 4.7 (a) illustrates the weight percentage of debris as the EP particle compact density

Page 125: Elastic And Mechanical Properties Of Expanded Perlite And ...

101

increases. This is also demonstrated by the particle size distribution measured after

compaction which is shown in Figure 4.7 (b).

(a)

(b)

Figure 4.7. (a) Percentage of debris at each EP compact density. (b) Cumulative particle size

distribution for entire compacts (EP particles and debris).

Page 126: Elastic And Mechanical Properties Of Expanded Perlite And ...

102

It can be seen that as the compact density increases due to the breakage of particles, the

percentage of larger particles decreases and the amount of debris increases. This post

compaction investigation revealed that the mass of EP particles that resisted crushing at

densities higher than 0.2 g/cm3 remained relatively constant. This was very strongly

correlated with the plateau region of the compression wave and shear wave velocity versus

density curves, suggesting that only the EP particles contribute to sound wave propagation

and the debris plays little role. This was further validated by investigating the evolution of

the inter-particle porosity as compact density increases. To this end, the density of different

EP particles size ranges was measured using the wax immersion method. In this method, wax

was melted in a 5ml measuring cylinder and kept at a constant temperature for each

measurement. EP particles of known mass and size were immersed in the melted wax and

their particle density was measured through the change in the wax’s volume. The results are

shown in Figure 4.8 (a). Using the results shown in Figure 4.7 (b) and Figure 4.8 (a), the

inter-particle porosity (excluding debris) within the total sample volume was calculated and

is presented in Figure 4.8 (b). It can be seen that the inter-particle porosity does not change

significantly as the compact density increases. This can be explained by the breaking of the

original EP particles into smaller EP particles as the compact density increases. The smaller

particles created new inter-particle porosity the increase of which was offset by the

simultaneous production of debris. Considering the mass fraction of the debris at each

compact density (see Figure 4.7 (a)) and the results shown in Figure 4.8 (a), the volume

percentage of debris with respect to the total sample volume as well as with respect to the

inter-particle porosity volume was calculated (Figure 4.8c). As can be seen, the debris does

Page 127: Elastic And Mechanical Properties Of Expanded Perlite And ...

103

not occupy a significant portion of the total sample volume and the inter-particle porosity

volume due to the very high density of the debris compared with the density of EP particles.

Therefore, the raw compact densities were modified by subtracting the debris mass from the

total mass of the EP particles and new densities were calculated. The modified results are

plotted along with the experimental results in Figure 4.5.

It should be noted that both of these results are useful. In practical situations, it is not usually

feasible to separate the debris from the EP particles. However, the raw or experimental results

are readily measured and reflect the effective state of the granular body. Hence, these results

are of most practical interest. Otherwise, if the properties of pure EP particles are desired for

more in-depth studies, the modified results are appropriate.

Young’s moduli of packed beds of EP particles, corresponding to the wave velocities in

Figure 4.5, calculated based on both raw compact density and modified density using Eq.

(3.4) is shown in Figure 4.9 (a). Figure 4.9 (a) also illustrates the normalised value of Young’s

modulus (E/E0) versus the compact porosity. Both curves, calculated from the experimental

and modified wave velocity results show that the Young’s modulus of packed beds of EP

particles increases as the density of the compact increases. In addition, these two curves

provide an upper and lower bound for Young’s modulus of EP particle compacts. In the

authors’ view, the lower bound is closer to the true value considering the minor mechanical

role of the debris.

Figure 4.9 (b) shows Poisson’s ratio of EP particle compacts as a function of compact density

as well as the normalised Poisson’s ratio (𝜈/𝜈0) as a function of porosity. As the calculation

of Poisson’s ratio is independent of density (Eq. (3.5)), the values obtained from the

Page 128: Elastic And Mechanical Properties Of Expanded Perlite And ...

104

experimental and modified wave velocities resulted in the same Poisson’s ratios. Poisson’s

ratio does not show as large a variation with density as Young’s modulus does within this

range of compact density (or porosity). It increases up to density 0.25 g/cm3 followed by a

decreasing trend however the changes are small (<10% over the whole range). There is no

general consensus among researchers about the influence of porosity on Poisson’s ratio.

While some consider Poisson’s ratio as a constant [222], others claim it to be a function of

porosity [223-225]. Such debates may arise from the fact that the Poisson’s ratio is usually

derived as a function of two other elastic moduli (i.e. Young’s and shear moduli) whose

relative dependence on porosity may intensify or supress the dependence of Poisson’s ratio

on porosity. The evolution of Poisson’s ratio in the low density compacts seems to be affected

by the pore character rather than the solid phase material properties. The compacted bed of

EP particles has a double-porosity structure; intra-particle porosity within the individual

particles and inter-particle porosity between the perlite particles. Thus, there are two

phenomena which contribute to Poisson’s ratio of the samples: Poisson’s effect of unit cells

within individual particles and Poisson’s effect of the whole particle. This suggests that in

the low density samples (ρ < 0.15 g/cm3), the bulk elastic response may be dominated by the

Poisson effect of cells as well as the compliance of the inter-particle contact region. The

initial contact areas are relatively small and well separated, hence localised lateral

deformation (Poisson effect) has a greater probability of intruding into the porous region

between particles rather than straining the adjacent particle. Therefore, the overall lateral

deformation is reduced, resulting in suppression of the Poisson’s ratio. Further increase in

the density reduced the porosity between particles and increased the contact size along with

Page 129: Elastic And Mechanical Properties Of Expanded Perlite And ...

105

increasingly more solid phase material being involved in load transfer throughout the bulk.

Consequently, the lateral deformation in both unit cells and individual particles contributes

to the observed Poisson’s ratio.

(a)

(b) (c)

Figure 4.8. (a) Particle density versus particle size. This graph also includes debris density versus

debris size; (b) Inter-particle porosity (excluding debris) versus compact density; (c) Inter-particle

space filled by debris versus compact density.

Page 130: Elastic And Mechanical Properties Of Expanded Perlite And ...

106

(a)

(b)

Figure 4.9. (a) Young’s modulus (E) of packed beds of EP particles versus compact density. (b)

Poisson’s ratio of packed beds of EP particles versus compact density. In both graphs, the upper and

right hand scales allow the normalised moduli versus porosity to be read from the same graphs.

Page 131: Elastic And Mechanical Properties Of Expanded Perlite And ...

107

Mathematical models for prediction for Elastic properties of porous

bodies

A large number of empirical and semi-empirical models have been developed to explain the

porosity dependence of material properties. The majority of these models are successful in

predicting elastic properties within a porosity range of less than 38% (e.g. [222, 226, 227])

and some assume the vanishing of elastic properties at porosity of 50% (i.e. as a result of an

implicit assumption of symmetry between pore and solid phase [228-231]). However, this is

not valid as many granular and porous materials have measurable elastic properties for

porosities greater than 50% (e.g. bodies made of hollow spherical particles [232]).

Consequently, models have been developed which cover the other end of the porosity

spectrum (>50%). In this study, four of these models were used to predict the elastic

properties of EP particle compacts of different densities. It is noteworthy that when choosing

a model, three criteria were considered: (i) interpretation of the physical parameters of the

model; (ii) whether standard relations of linear elasticity hold between all moduli; and (iii)

the ability of the model to satisfy the boundary conditions. The boundary conditions include

the prediction of the solid phase properties at zero porosity and the presence of a critical

porosity, or percolation limit, at which particles no longer form a continuous network. Hence

stiffness (i.e. E, G, K) goes to zero and the Poisson’s ratio approaches an asymptotic value.

This value has been found to be independent of the Poisson’s ratio of the solid phase but is

dependent on the geometry of the solid phase at the critical porosity [233]. Experimental and

numerical investigations have shown that, regardless of the solid phase Poisson’s ratio, the

Poisson’s ratio of a porous body asymptotically approaches a fixed value which has been

Page 132: Elastic And Mechanical Properties Of Expanded Perlite And ...

108

identified differently to be 0.2 [234-236] or 0.25 [223, 237]. Dunn [225] analytically

investigated four different pore shapes as a function of the aspect ratio of a spheroid and

found pore shape is another factor affecting the asymptotic behaviour of Poisson’s ratio. It

was found that the asymptotic value for spherical and needle shaped pores is 0.2 while for

disk-shaped pores and penny shaped cracks it is 0. With respect to experimental and

analytical findings, the critical Poisson’s ratio is considered to lie in the range 25.00 cr

at critical porosity, crP . For a granular material, crP falls within the range of gravity induced

packing states Ptap < Pcr < Pgreen, where Ptap and Pgreen are the porosities of the ‘tapped’ and

‘as-poured’ packing states [238]. In the evaluation of the models, the value of Pcr was

considered to lie within these bounds in order to preserve its physical interpretation.

It should be noted that particle crushing and the production of debris is not taken into account

in any of the models’ assumptions. Hence, it would be more appropriate to apply these

models to the modified values. However, to investigate the effect of particle debris on the

physical features of the models, the models were applied to moduli determined using both

the raw compact densities (hereafter the experimental moduli) and moduli calculated using

the modified densities, hereafter the modified moduli. In the following, each model and its

physical features are explained, and the applicability of these models to the EP particle

compaction data is discussed.

Page 133: Elastic And Mechanical Properties Of Expanded Perlite And ...

109

4.4.1 Phani Models

Phani and Niyogi [239] derived a semi-empirical relation based on a simple model of

applying a uniform state of stress on a porous body of constant cross-sectional area and

constant length. This model has been found to agree well with the data from many

polycrystalline brittle solids over a wide range of porosity [239-242] and is expressed by:

EnaPEE )1(0 (4.7)

where E and E0 are Young’s modulus at volume fraction porosity P and zero, respectively,

and a and En are packing geometry and pore structure dependent parameters, respectively.

This model satisfies the boundary conditions: E = E0 at P = 0, and E= 0 at P = Pcr = 1/a.

Rice [243] theoretically derived the values of Pcr to be 0.785, 0.964 and 1.0 for the cubic

staking of cylindrical, spherical and hexagonal pores, respectively. Knudsen [244] studied

the contact area as a function of bulk density and calculated the value of Pcr for rhombohedral,

orthorhombic and cubic packing of spherical particles to be 0.26, 0.397 and 0.476,

respectively. From the theoretically derived values of Pcr and the existing relationship

between Pcr and parameter a (i.e. Pcr = 1/a), the value of a lies in the range 85.31 a . For

random packing with isolated spherical pores, the constants a and En were found to be about

1 and 2, respectively [239]. However, the increase in En corresponds to the transition of

pores from being spherical to being more interconnected.

Phani and Sanyal [245] derived a relation between the shear modulus and Young’s modulus

of an isotropic porous solid based on the Mori–Tanaka mean-field approach, in which the

Page 134: Elastic And Mechanical Properties Of Expanded Perlite And ...

110

elastic properties are obtained by subjecting the inclusion (i.e. pores) to an effective stress or

strain field. This relation is given by:

0

0

0

0

0

0

3

21)1(

3

2G

E

E

E

EG

n

(4.8)

where G is the shear modulus, E is predicted from Eq. (4.7) and 0n is a constant for a given

data set of a porous material related to pore morphology. The value of 0n can be evaluated

using the experimentally measured values of E and G at a single porosity value. This model

satisfies the boundary conditions: G = G0 when E = E0 at P = 0, and G = 0 when E = 0 at P

= Pcr.

Predicted values of Young’s moduli from Eq. (4.7) along with the ones for shear moduli from

Eq. (4.8) were used to evaluate Poisson’s ratio by,

12

G

E (4.9)

as a function of porosity. The required data for pore-free solid phase (i.e. E0, G0 and ν0) were

obtained from Table 4.1. Non-linear regression coefficients to fit the model to the

experimental and modified moduli are presented in Table 4.5. The regression value for nE in

both cases (i.e. based on the experimental and modified moduli) was indicative of the

compacts’ pore structure being non-spherical and partially interconnected. The regression

value of the constant a (i.e. Pcr = 1/a) gives a Pcr of 1.0. For packed beds of EP particles, the

range Ptap < Pcr < Pgreen was experimentally determined to be 0.96 < Pcr < 0.98. Thus, the

estimated value of Pcr by Eq. (4.7) is slightly higher than the expected range.

Page 135: Elastic And Mechanical Properties Of Expanded Perlite And ...

111

Table 4.5. Physical parameters of mathematical models applied to the experimental and modified

moduli. The physical parameters of the modified moduli are given in parenthesis.

Model Name Physical Parameters

Phani a = 1 (1); nE = 2.8 (3.095); n0 = 1.104 (1.053); Pcr = 1 (1)

Nielsen β = 0.24 (0.12); Pcr = 0.965 (0.979)

Rice1 bʹE = 0.036 (0.02); bʹG = 0.0327 (0.02); Pcr = 0.976 (0.978)

Wang b = 3 (3.63); c = 3.1(3.08); d = 0.95 (1); Pcr = 1 (1)

Gibson-Ashby C1 = 0.1 (0.068); C1ʹ = 0.1 (0.072); C2 =0.037 (0.026);

C2ʹ = 0.039 (0.0276); Pcr = 1 (1)

1. The subscript E in bʹE refers to Young’s modulus and subscript G in bʹG refers to shear modulus

Application of the Phani model to both the experimental and modified moduli is illustrated

in Figures 4.10 (a) - (d). In both cases, Young’s modulus values predicted by Eq. (4.7) show

a considerable deviation (see Figures 4.10 (a) and (c)). Although the predicted Poisson’s

ratios (Figures 4.10 (b) and (d)) do not follow the same trend as the experimentally measured

values, they show slightly better agreement with the ones obtained based on the modified

densities. It should be noted that Poisson’s ratio is a small quantity dependent on the

differences of other elastic moduli and is hence very sensitive to errors in them [212]. The

considerable deviation from the experimental moduli of values predicted by the Phani model

might be ascribed to the effect of the shape factor nE as a single parameter to reflect the

Page 136: Elastic And Mechanical Properties Of Expanded Perlite And ...

112

change in porosity as compaction proceeds. This effect is discussed in more detail in Section

4.4.2.

(a) (b)

(c) (d)

Figure 4.10. (a) Normalised Young’s modulus versus porosity (based on experimental moduli); (b)

Normalised Poisson’s ratio versus porosity (based on experimental moduli); (c) Normalised

Young’s modulus versus porosity (based on modified moduli); (d) Normalised Poisson’s ratio

versus porosity (based on modified moduli).

Page 137: Elastic And Mechanical Properties Of Expanded Perlite And ...

113

4.4.2 Nielson Model

Nielsen [246] proposed a model for predicting the elastic moduli of porous materials based

on the composite sphere assemblage (CSA) approach. CSA considers a porous body to be

constituted of congruent composite elements consisting of a spherical pore embedded in a

concentric spherical shell of matrix material. It is assumed that the composite spheres are

available in an infinite range of sizes and they are distributed in a way where smaller

composite spheres fill all the interstices between larger spheres. This model was originally

introduced by Hashin and assumed the applied stress on the assembled body is uniformly

distributed (hydrostatic) around each inclusion so that strains can be calculated to obtain the

elastic moduli for the body. Nielsen set forth to predict the elastic moduli for porous materials

by assuming two types of composite elements: one as described above and the other one

made of a spherical matrix embedded in a concentric spherical shell of pore. These two types

of CSAs defined two bounds of continuous solid phase (isolated pores) and continuous pore

phase (isolated solid phase) where transition between them (different mixtures of them)

results in porous material with a wide range of porosity of any geometry. The Nielsen model

for Young’s modulus and Poisson’s ratio are expressed as:

)1315)(1())(75(2

))(75)(1(2

000

00

PPPP

PPPEE

crcr

cr (4.10)

)1315)(1())(75(2

)35)(1())(75(2

000

0000

PPPP

PPPP

crcr

crcr (4.11)

where 𝛽 is a shape factor characterising the ability of the material to transfer stress and

providing information on the shape of the pores and their inter-connectivity. The shape factor

Page 138: Elastic And Mechanical Properties Of Expanded Perlite And ...

114

𝛽 is in the range 0< β <1. The lower bound, 𝛽 = 0, is when the pore phase totally surrounds

the solid phase (i.e. particle) while the upper bound, 𝛽 = 1, corresponds to when the solid

phase completely surrounds the pore phase (i.e. isolated pores). Hence, the lower value of 𝛽

correlates with highly interconnected pores and smaller contact areas between particles,

while higher values of 𝛽 indicates that pores are becoming increasingly more isolated.

The Nielsen model satisfies the boundary conditions at zero porosity and correctly predicts

both of the solid phase elastic moduli (i.e. E = E0, 𝜈 = 𝜈0). It also predicts the presence of

the critical porosity at which Young’s modulus goes to zero and Poisson’s ratio is given by

1315

35

0

0

Cr . Given that Poison’s ratio for an isotropic elastic material lies in the

physically realistic range 0 ≤ 𝜈0 ≤ 0.5, this model predicts critical Poisson’s ratio to lie in

the range 0.09 ≤ 𝜈𝑐𝑟 ≤ 0.23, which does not cover the expected range 0 ≤ 𝜈𝑐𝑟 ≤ 0.25.

This limited range may have a significant effect on the quality of predicted values for

Poisson’s ratio.

Non-linear regression analysis was conducted to fit the Nielson model to the experimental

and modified moduli and the results are shown in Table 4.5. The predicted values of Pcr for

both the raw compact and modified densities were within the specified range 0.96 < Pcr <

0.98 and the low value of 𝛽 is representative of a high-volume fraction of pores compared to

the solid phase. Figures 4.10 (a) and (b) show that the agreement between the Neilson model

and the experimental moduli is not very good. In the case of Poisson’s ratio, the discrepancy

can be attributed to the low value of νcr (i.e. 0.09 ≤ 𝜈𝑐𝑟 ≤ 0.23) and the shape factor β.

Similar to nE in the Phani model, the shape factor β in the Nielson model is used as a single

Page 139: Elastic And Mechanical Properties Of Expanded Perlite And ...

115

value to accommodate microstructural change during the compaction of compacts of

different densities and with a wide range of pore shapes and geometries. This problem might

be alleviated by solving Eqs. (4.7) and (4.10) point by point. The results of this calculation

for both of the shape factors (nE and β) as a function of porosity are presented in Figure 4.11

(a). As can be seen, both shape factors change as a function of porosity and the results show

the opposite of what was expected. At lower porosity 𝛽 approaches zero while at successively

higher porosities, where the structure is open and interconnected, it approaches 0.5 for the

experimental moduli and exceeded the upper bound of unity for the modified moduli. Similar

to 𝛽, the evolution of the shape factor nE with porosity does not follow the expected behaviour

defined by Phani and Niyogi. To accommodate these differences, the shape factors were

expressed as a function of porosity P, given by:

134

2321 ....)( n

n PPPPPPS (4.12)

where S is a shape factor, either nE or β and 𝜂𝑛 (n = 1, 2, 3, ...) are empirical constants.

Applying the modified form of the shape factor in Eq. (4.12) to the experimental and

modified moduli up to n = 4, both Nielson’s and Phani’s models were replotted and are

presented in Figures 4.11 (b) and (c). It can be seen that both Nielson’s and Phani’s models

for Young’s modulus show close agreement with the experimental moduli when the shape

factor was defined as a function of porosity. In the case of Poisson’s ratio, the modified form

of the shape factor improved the results of the Phani model but not those from the Neilson

model (see Figure 4.11 (c)). Therefore, as the Nielson model does not exhibit good agreement

with both the Young’s modulus and Poisson’s ratio, it is not considered to be a good model

for the prediction of the elastic properties of EP particle compacts.

Page 140: Elastic And Mechanical Properties Of Expanded Perlite And ...

116

(a) (b)

(c)

Figure 4.11. (a) Shape factor versus compact porosity for the Phani and Neilson models applied to

both the experimental moduli and modified moduli; (b) Modified Phani and Nielson models for

Young’s Moduli as a function of porosity; (c) Modified Phani and Nielson models for Poisson’s

ratio as a function of porosity.

Page 141: Elastic And Mechanical Properties Of Expanded Perlite And ...

117

4.4.3 Minimum solid area (MSA) models

The minimum solid area (MSA) model assumes that the macroscopic elastic response is

related to the load-bearing area of the solid phase material. The key assumption in MSA

models is that normalised elastic moduli scale according to the minimum area of solid phase

per unit area in a plane perpendicular to the direction of the applied load, i.e. MSAE

E

0

. In

this context, a common approach is to assume an idealised microstructure, including the

regular stacking of solid particles in a void matrix, and the regular stacking of pores (spheres,

cylinders, etc.) in a solid matrix. The initial and major development of such models was

Knudsen’s model [244]. Knudsen investigated the ideal microstructure of three particle

stackings (i.e. simple cubic, orthorhombic and rhombohedral) and found that the stacking of

particles, and hence the pores, is significant. In addition, plotting of the resultant models as

a function of porosity on semi-log plots for low to intermediate porosity was approximated

by:

bPeEE 0 (4.13)

where b is related to particle stacking and hence is a function of pore shape, geometry and

alignment with respect to the stress axis. This equation has been criticised [227] for not

satisfying the boundary condition that E = 0 at P = 1. However, such criticism is only valid

when critical porosity occurs, at about unity. Anderson [247] analytically proved the

existence of such exponential dependence on porosity from strain analysis of isolated

ellipsoidal pores. He stated that Eq. (4.13) becomes invalid at very high pore fractions where

constant b does not reflect pore interactions occurring as the concentration of pores increases.

Page 142: Elastic And Mechanical Properties Of Expanded Perlite And ...

118

Rice [243] discussed that this relationship (Eq. (4.13)) provides a good approximation for

effective properties up to 1/3-1/2 of 𝑃/𝑃𝑐𝑟 and holds true for all elastic moduli. In attempts

to estimate the minimum solid area for higher porosity, Rice and Wang proposed two

equations. Rice [248] suggested that the role of pores and solid phase material can be

exchanged (i.e. stacking of pores rather than particles) leading to the theoretical equation:

)1()/1(

0

'crPPb

eMM

(4.14)

where M is the Young’s, shear or bulk moduli and 𝑏′ is related to the pore stacking. Rice

discussed that 𝑏′ decreases as b increases. For instance, for spherical pores in cubic stacking,

the value of b is 2.7 while the value of 𝑏′ is 0.5. For other pore shapes and stackings, the

values of b and 𝑏′ can be obtained from Figure 2 in [248]. This model predicts zero stiffness

as the porosity goes to unity, but it does not satisfy the zero porosity boundary condition

since (1 − exp (−𝑏′(1 − 𝑃/𝑃𝑐𝑟)) does not go to unity as P goes to zero. Rice explained this

by expressing that Eq. (4.14) is a continuation of Eq. (4.13) for higher porosity values and

no extrapolation between them is required.

Wang [222] used the ideal model of spherical particles in a simple cubic stacking developed

by Knudsen [244] and modified it for real microstructures. The modifications account for

misalignment of the uniaxial applied stress, which induces shear and hinge effects at the neck

(the small contact area between two particles). Wang derived a complex relation between

porosity and Young’s modulus, and proposed an approximate solution to his exact solution

which could satisfactorily cover a wide range of porosity, as given by:

...)](exp[ 32 dPcPbPEE S (4.15)

Page 143: Elastic And Mechanical Properties Of Expanded Perlite And ...

119

where b, c and d are empirical constants. Brown et al. [249] analytically derived the value of

b for different idealised pore geometries and orientations and proposed a similar equation to

Wang’s model, but in the context of strength, as given by:

j

i

i

ii

i

ii

i

ii

i

i Pbj

PbPbPb )(!

1)(

!3

1)(

!2

1)/ln( 32

0 (4.16)

where ib is related to the pore geometry and orientation of the ith kind of pore and iP is the

contribution made by pores of the ith kind to the total porosity. Therefore, the constants of

Eq. (4.15) can be expressed in terms of the constants of Eq. (4.16). Wang has shown that the

proposed equation with a quadratic exponent can predict the Young’s modulus of a porous

material up to a porosity of 38% [222] and additional higher order terms can be included for

higher porosities.

The Wang model with a cubic exponent was used for the prediction of elastic moduli of EP

particle compacts. Non-linear regression analysis was conducted to fit the Wang and Rice

models to the experimental and modified moduli. The results are presented in Table 4.5. The

critical porosity value predicted by the Wang model is slightly higher than the specified range

0.96 < Pcr < 0.98, while the value predicted by the Rice model (for both shear and Young’s

moduli) is within the range. As shown in Figures 4.10 (a) and (c), the Rice model shows very

close agreement with the experimental moduli, however, it did not show such agreement with

the modified moduli. Similarly, the Wang model shows better agreement with the

experimental moduli than with the modified Young’s modulus values (see Figures 4.10 (a)

and (c)). The regression parameters in Wang’s model for the experimental Young’s moduli

in terms of the constants in Eq. (4.16), correspond to; b - cylindrical pores aligned

Page 144: Elastic And Mechanical Properties Of Expanded Perlite And ...

120

perpendicular to the loading direction; c - oblate pores (with a ratio of about 0.8) in a cubic

stacking; and d - a combination of about 70% cubical pores in <100> orientation with 30%

cylindrical pores parallel to the loading direction [250-252], respectively. The interpretation

of the regression parameters in Wang’s model for the modified Young’s moduli is similar to

those for the experimental moduli, except for constant b (3.63) which corresponds to a

combination of about 77% cubical pores in <110> direction with 23% cubical pores in <111>

direction. This approximation seems reasonable with respect to the cell shape in EP particles

(Figure 4.3) and the fact that pores can have different directions in a real packing of particles.

The predicted Young’s and shear moduli based on both experimental and modified moduli

were found to lie within the simple cubic and rhombohedral packing of pores. For

comparison, a similar range of values can be acquired from Figure 2 in [248]. Rice discussed

how porous bodies are better modelled by combining two or more idealised pore structures,

as opposed to just one. Consequently, the values of 𝑏′ (Table 4.5) can be ascribed to different

combinations of these two pore packing geometries (e.g. simple cubic and rhombohedral)

and their interaction as porosity evolves. Poisson’s ratio values were calculated using the

predicted Young’s and shear moduli values from Eq. (4.14) and then combining them using

Eq. (4.9). Poisson’s ratio values predicted by the Rice model based on raw compact density

show close agreement with the experimentally measured Poisson’s ratio values. However,

the deviation of Young’s and shear moduli values predicted by the Rice model based on

modified moduli resulted in poor predictive results for Poisson’s ratio (see Figures 4.10 (b)

and (d)). As a whole, the Rice model was shown to be successful in predicting the elastic

properties of EP particle compacts only when the raw compact density is considered

Page 145: Elastic And Mechanical Properties Of Expanded Perlite And ...

121

4.4.4 Gibson and Ashby Model

Gibson and Ashby derived a semi-imperial model for the relationship between elastic moduli

and porosity by employing dimensional arguments (using standard beam theory) for a

cellular solid. For simplicity, they assumed cell struts and walls of uniform dimensions, while

in reality these will usually be tapered toward, and thinner in, the centre (i.e. due to the effect

of surface tension during the foaming process). The correction for this tapering was

implemented by arranging the cubic cells in a staggered stacking, so that their members meet

at their midpoint. This model has been shown by many authors (e.g. [253-258]) to

successfully predict the elastic moduli of closed cell foams, which are given by the following

expression:

)1)(1(1 '

1

22

1

0

PCPCE

E

(4.17)

)1)(1(1 '

2

22

2

0

PCPCE

G

(4.18)

Where 𝐶 and 𝐶′ are constants of proportionality, Ø is the volume fraction of solid in the cell-

struts ,which can be obtained from the relative area of the struts and faces on a plane section

as explained in [259], and the remaining fraction (1− Ø) is the solid contained in the cell

faces. The quadratic term describes the contribution of the cell struts bending to the modulus

and is the same for open cell foams. The linear term corresponds to the cell walls’ lateral

stretching. Gibson and Ashby suggested 𝐶1 and 𝐶1′ should be about unity and 𝐶2 and 𝐶2

should be about 3/8 for porous materials with 𝜈0 = 0.33, where these satisfy the boundary

conditions at P = 0 and P = Pcr = 1. However, the values of 𝐶1 and 𝐶1′ will change based on

Page 146: Elastic And Mechanical Properties Of Expanded Perlite And ...

122

the volume fraction of solid in the cell struts, the variable geometries of the foam and

uncertainty in the value of the solid Young’s modulus, E0 [97]. As a result, the values of these

constants depend on the type of foam and vary from one foam to another.

The Gibson–Ashby relation provides a good approximation for the cellular structure of the

EP particles shown in Figure 4.3. For a given porosity and solid fraction of cell struts (Ø =

0.78), the regression analysis yielded the values 𝐶1 and 𝐶1′ in Eq. (4.17) for both the

experimental and modified Young’s moduli (see Table 4.5). Application of the Gibson and

Ashby model to the experimental and modified Young’s moduli is presented in Figures 4.10

(a) and (c). This model shows good agreement with the experimental Young’s modulus

values. In the case of the modified Young’s moduli, the Gibson and Ashby model

demonstrates fair agreement with the moduli for porosity higher than 0.87, while for lower

porosity a considerable deviation is observed.

Gibson and Ashby suggested that if 𝐶1 and 𝐶1′ are about unity and

8

3

E

G, which holds true

for polycrystalline metallic materials [260], Poisson’s ratio of the foam is about 0.33. Since

constants 𝐶1 and 𝐶1′ are not unity (see Table 4.5), this relation does not hold and Poisson’s

ratio should be calculated using two elastic moduli. For this purpose, the Young’s moduli

from Eq. (4.17) along with the shear moduli predicted from Eq. (4.18) were used to calculate

Poisson ratios by Eq. (4.9). For a given porosity and solid fraction of cell struts (Ø = 0.78),

the regression analysis yielded the values of 𝐶2 and 𝐶2′ in Eq. (4.18) for both the experimental

and modified shear moduli (see Table 4.5). It is noteworthy that although 𝐶1 and 𝐶1′ are not

unity, the ratio of the shear to Young’s moduli is almost 3/8, which gives implicit satisfaction

of the boundary condition (explained above). The results for Poisson’s ratio are presented in

Page 147: Elastic And Mechanical Properties Of Expanded Perlite And ...

123

Figures 4.10 (b) and (d). As can be seen, the Poisson’s ratio values predicted by the Gibson

and Ashby model show good agreement with both the experimental and modified Poisson’s

ratios. Overall, the Gibson and Ashby model gives a satisfactory means of predicting the

elastic properties of EP particle compacts based on the experimental and modified moduli.

Page 148: Elastic And Mechanical Properties Of Expanded Perlite And ...

124

Properties of epoxy resin

The mechanical properties of the cured epoxy resin, with and without dilution with acetone,

were measured according to the specified ASTM standards and are presented in Table 4.6.

Table 4.6. Mechanical properties of cured epoxy resin with and without dilution by acetone.

Properties Epoxy/Hardener Epoxy/Hardener/Acetone

Hardness (Shore D) ASTM D-2240 82 57

Compression yield ASTM D-695 103 MPa 3.40 MPa

Compressive modulus ASTM D-695 3170 MPa 1002 MPa

Tensile strength ASTM D638 45.40 MPa 7.50 MPa

Tensile elongation ASTM D-638 7.3 % 3.20%

Tensile Modulus ASTM D-638 2910 MPa 1834.7 (2022.4) 1 MPa

Flexural strength ASTM D-790 184 MPa 35.70 MPa

Flexural Modulus ASTM D-790 7365 MPa 1370 MPa

Poisson’s ratio 0.39 (0.37) 0.41 (0.38)2

Density 1.18 1.14 (g/cm3)

1 & 2. The measured properties using elastic wave tests are given in parenthesis.

Dilution with acetone changed the curing time of the epoxy which was investigated by

monitoring hardness change using a durometer (type D). The results presented in Figure 4.12

are indicative of a minimum curing time of 45 days for diluted epoxy as no change in

hardness was observed after this period. The dilution with acetone at this ratio reduced the

Page 149: Elastic And Mechanical Properties Of Expanded Perlite And ...

125

mix viscosity of epoxy resin+hardner from 1.80 Pas (with no acetone) to 0.0067 Pas (with

90 wt% acetone); viscosity was measured, shown in Figure 4.13, using RST-Brookfield

rheometer and the data were subsequently analysed using the Bingham model [261]. The

point for 100% epoxy + hardener is not shown in the graph due to scale limitations.

Figure 4.12. Hardness versus time curve used to determine the curing period of the diluted epoxy.

Figure 4.13. Changes in the viscosity of epoxy + hardener diluted with acetone versus acetone

content.

Page 150: Elastic And Mechanical Properties Of Expanded Perlite And ...

126

Properties of EP/epoxy foams

4.6.1 Microstructural characterisation of EP/epoxy foams

Scanning electron microscopy was used to examine the microstructure of foam samples

before undergoing compression tests. Figure 4.14 shows micrographs taken from cross-

sections of the three foam types with the same density. It can be seen that the epoxy has filled

the space between the particles and did not penetrate into the perlite particles, which explains

the low density of the foams. The SEM images also show good bonding between the perlite

particles and the epoxy binder as few gaps were seen at their interfaces. However, this may

not have a significant effect on the resulting properties (mechanical, physical, etc.). This can

be due to the porous structure as well as the size and volume fraction of the perlite particles.

The porous structure of perlite inhibits a continuous layer of adhesion between the particle

and matrix, compared with if there is a solid continuous surface. Regarding the particle size,

it has been reported [262] that particles of smaller size have a higher surface area to volume

ratio leading to better interfacial interaction with the matrix. The maximum of such

interactions for spherical and near-spherical particles was achieved at the size range of

nanometer and submicrometer particles [120, 263-265]. It was found that the particle

contents in the range 25 - 30 vol% provided the maximum interfacial interaction between the

filler (particle) and the matrix. At higher particle content, the overlap of interfacial adhesion

between adjacent particles reduced the overall filler-matrix interfacial interaction [263]. In

the present study, however, the size of the particles ranged from 1 to 2, 2 to 2.8 and 2.8 to 4

mm and the volume fraction ranged from 46 to 58% (to be explained in Section 5.4).

Page 151: Elastic And Mechanical Properties Of Expanded Perlite And ...

127

Therefore, adhesion between the particles and the epoxy cannot be considered to have had a

notable effect on the resulting mechanical properties. This will be investigated further in

Section 4.6.3.

(a) (b)

(c)

Figure 4.14. Micrographs taken from the cross-sections of EP/epoxy foams with a density of 0.15

g/cm3 made with EP particles in the size ranges: a) 1 - 2mm; b) 2 - 2.8mm; and c) 2.8 - 4mm.

In addition, an examination of the micrographs shows fewer gaps (marked on Figure 4.14)

between the epoxy binder and the EP particles in the foams of type 1 and 2 than foams of

Page 152: Elastic And Mechanical Properties Of Expanded Perlite And ...

128

type 3. Hence, it is expected that cracks initiate and develop faster under compressive loading

in the foams of type 3 than in the other two types of foam. This will be investigated in detail

in Section 4.7.

4.6.2 Volume fraction of the epoxy in EP/epoxy foams

In the following, the volume fraction of the epoxy binder in EP/epoxy foams is quantified by

employing an estimation method as well as the measured experimental values. However, the

comparison of the results and a discussion of the ability of the estimation method to quantify

this parameter are held over until Section 5.4.

4.6.2.1 Estimating the volume fraction of epoxy in EP/epoxy foams from the foam

density

The volume fraction of epoxy Ve in the foams was calculated theoretically from the rule of

mixtures given in Eq. (4.19):

)/()( PePceV (4.19)

where ρc is the composite density, ρp is the particle density and ρe is the epoxy density. It

should be noted that the particle densities presented in Table 4.2 cannot be used in Eq. (4.19)

since these densities only refer to undeformed particles. During the sample manufacturing

process, the compaction pressure effectively changed the density and density distribution of

the EP particles (see Figure 4.8 (a)). Therefore, a modified version of Eq. (4.19) was used to

estimate the volume fraction of epoxy in the manufactured foams:

PePfe DDDV / (4.20)

Page 153: Elastic And Mechanical Properties Of Expanded Perlite And ...

129

where Df and Dp are the foam density and particle density as a function of compaction

pressure (σcomp), respectively:

)( compf fD ; )( compp fD

To obtain the particle density as a function of compaction pressure, a group of samples, which

only contained perlite particles was prepared. To this end, EP particles of known mass were

compacted in the same mould used for manufacturing the EP/epoxy foam samples. As the

compaction proceeded, the change in height, and hence density, were calculated as a function

of compaction pressure. Results illustrating the particle density versus compaction pressure

are shown in Figure 4.15 (a).

Using the experimental results and the assumption that there was little or no ingress of epoxy

into the EP particles, the foam density Df and particle density Dp as a function of compaction

pressure (𝜎𝑐) can be expressed by the following equations:

32

2

1 CCCD CCf

(4.21)

32

2

1 CCPD (4.22)

The coefficients Ci and αi for the foam density Df and particle density Dp were determined

using least squares fitting. The coefficients for the foam density and particle density

containing EP particles of three particle size ranges are given in Table 4.7.

Page 154: Elastic And Mechanical Properties Of Expanded Perlite And ...

130

(a)

(b)

Figure 4.15. (a) Foam density as a function of applied pressure. (b) Particle density as a function of

compaction pressure.

Page 155: Elastic And Mechanical Properties Of Expanded Perlite And ...

131

Table 4.7. Coefficients in Equations (4.21) and (4.22) for foam density Df and particle density Dp .

Particle

size range 𝐶1 𝐶2 𝐶3 𝑅2 𝛼1 𝛼2 𝛼3 𝑅2

1-2mm -0.0114 0.1268 0.1549 0.9876 -0.0078 0.1188 0.125 0.9908

2-2.8 mm -0.0129 0.1313 0.1456 0.9759 -0.0174 0.1574 0.0874 0.9807

2.8-4 mm -0.0146 0.1405 0.1329 0.9792 -0.0066 0.1065 0.1016 0.9908

Using Eqs. (4.21) and (4.22), the particle density and the corresponding foam density at each

pressure level were calculated, substituted in Eq. (4.20) and subsequently the volume

percentage of epoxy (Ve) was calculated. Figure 4.16 illustrates the volume percentage of

epoxy as a function of foam density for the three types of foam.

Figure 4.16. Volume fraction of epoxy versus density of the foam.

Page 156: Elastic And Mechanical Properties Of Expanded Perlite And ...

132

The results show that foams of type 3 have the highest volume percentage of epoxy. In

contrast, foams of type 1 show the lowest volume percentage of epoxy versus density. The

volume percentage of epoxy increases with the foam density in the range of 0.15 - 0.4 g/cm3

for foams of type 3 but reaches a maximum earlier at 0.38 g/cm3 for type 1 foams. At higher

densities, though, a gradual decrease in the volume fraction of epoxy is observed, which will

be discussed in Section 5.4.

4.6.2.2 Calculating the volume fraction of epoxy in EP/epoxy foams

The volume fraction of epoxy in each foam sample was calculated from the difference

between the mass of the foam and the mass of the EP particles in that foam. Next, the mass

of the epoxy was divided by the density of the cured epoxy binder (i.e. 1.14 g/cm3) to get the

volume of the epoxy. Dividing the volume of the epoxy by the total sample volume gives the

volume fraction of the epoxy binder in that foam sample. Following this procedure, the

volume fraction of the epoxy binder in the three types of EP/epoxy foams were calculated

and are presented in Figure 4.17. Figure 4.17 shows that the volume percentage of epoxy

within the EP/epoxy foams was between 1.9% and 6% and was mostly in the range 2% - 5%.

Though it seems the volume fraction is almost constant, a decreasing trend in the type 1 and

2 EP/epoxy foams and a polynomial trend in the type 3 foams can be noticed. Overall, the

results presented in Figure 4.17 show that epoxy occupies a very small portion of the

EP/epoxy foams which explains their low density.

Page 157: Elastic And Mechanical Properties Of Expanded Perlite And ...

133

Figure 4.17. Volume fraction of epoxy binder in EP/epoxy foams of type 2.

4.6.3 Compressive response of EP/epoxy foams

Compressive stress-strain curves for the three types of EP/epoxy foams in the density range

0.15 - 0.48 g/cm3 are illustrated in Figure 4.18. The stress-strain curves show a similar trend,

with the maximum stress occurring in the strain range of 0.03 - 0.07, followed by strain-

softening. The stress-strain curves were used to evaluate the maximum stress, effective

elastic modulus, and modulus of toughness for different foam densities and these are

presented in Figure 4.19. Maximum stress is the peak value in the stress-strain diagram;

effective elastic modulus is the unloading gradient (details explained in Section 3.4.2); and

the modulus of toughness is the area under the stress-strain diagram which indicates the

strain-energy density absorbed by the material before it fractures [266]. The compressive

stress shows a linearly increasing trend with density (Figure 4.19 (a)), while the effective

elastic modulus (Figure 4.19 (b)) and the modulus of toughness (Figure 4.19 (c)) show (𝜌)1

2

and 𝜌2 trends, respectively. In addition, the results show that the compressive stress and the

Page 158: Elastic And Mechanical Properties Of Expanded Perlite And ...

134

effective elastic modulus of EP/epoxy foams are independent of the particle size range.

However, a small deviation was observed for the modulus of toughness (Figure 4.19 (c)).

Similar behaviour was observed in Perlite/sodium silicate [188] (explained in Section 2.3.3).

Foams of type 1 and 2 show a slightly higher energy absorption capacity. This can be seen

by comparing the stress-strain curves for the three types of foams (Figure 4.18). The stress-

strain curves for foams made with EP particles of 2 - 2.8 mm (Figure 4.18 (b)) show a plateau

for a significant range of strain and thus a higher energy absorption capacity. This can be

ascribed to the morphology of EP particles in the range of 1 - 2 mm and 2 - 2.8 mm which

have a larger edge cell thickness (Table 4.3) and contain internal fibre reinforcement (Figure

4.3). Other factors contributing to the increase in the modulus of toughness are better bonding

of smaller particles with epoxy when used in foams (Figure 4.14) and slightly higher epoxy

volume fractions in the type 1 and 2 foams (Figure 4.17).

(a)

Page 159: Elastic And Mechanical Properties Of Expanded Perlite And ...

135

(b)

(c)

Figure 4.18. Typical stress-strain curves for the different EP/epoxy foam densities of: (a)

Type 1; (b) Type 2; (c) Type 3.

Page 160: Elastic And Mechanical Properties Of Expanded Perlite And ...

136

(a)

(b)

Page 161: Elastic And Mechanical Properties Of Expanded Perlite And ...

137

(c)

Figure 4.19. Properties of manufactured foams of type 1 ( ⃣ ), type 2 (◇) and type 3 (∆): (a)

Maximum stress versus foam density; (b) Effective modulus versus foam density; (c) Modulus of

toughness versus foam density.

Considering the high volume fraction of EP particles in the manufactured foams, the EP’s

properties could have a significant contribution in the resulting foam properties. To measure

this contribution as a function of density, several loading-unloading confined compressive

tests were conducted on packed beds of EP particles as explained in detail in Section 4.3.2.

As discussed earlier, however, the moduli obtained from these tests were not equal to the

Young’s modulus; they were a function of both the Young’s modulus and Poisson’s ratio.

Therefore, to find the stiffness contribution of EP particles to the effective stiffness of

EP/epoxy foams, a series of additional confined tests was conducted on the EP/epoxy foams,

as explained in Section 3.4.2. As the experimental results have already proven that the

Page 162: Elastic And Mechanical Properties Of Expanded Perlite And ...

138

effective elastic moduli of the EP/epoxy foams are independent of particle size, the confined

tests were only conducted on the foams of type 2. The measured confined modulus of the EP

particles and the EP/epoxy foams as a function of the particle density are presented in Figure

4.20.

Figure 4.20. Confined modulus of EP particles ( ⃣ ) and confined modulus of the foams (◇) as a

function of the density of the packed bed of EP particles.

As can be seen in Figure 4.20, for foams having the same combined mass of EP particles per

unit volume as the packed bed samples of EP particles, the effective elastic modulus is

generally lower. This is due to friction and interaction between the EP particles. The inclusion

of the epoxy as a matrix reduced the friction between EP particles; however, this reduction

could not be quantified. The measured effective elastic moduli of the EP/epoxy foams are

much closer to the elastic moduli of EP particles than to that of epoxy (1002 MPa). This is

clearly indicative of the significant role of EP particles in the stiffness of the EP/epoxy foams

Page 163: Elastic And Mechanical Properties Of Expanded Perlite And ...

139

and can be ascribed to the high-volume fraction of the EP particles (88 - 94 Vol%) in these

foams (i.e. type 2). Though the stiffness provided by the EP particles in the EP/epoxy foams

cannot be measured using quasi-static mechanical tests, an upper 𝐸𝑃𝑈𝑝𝑝𝑒𝑟 and a lower 𝐸𝑃

𝐿𝑜𝑤𝑒𝑟

bound can be predicted using the Voigt and Reuss models given by Eq. (4.23) and Eq. (4.24).

The Voigt and Reuss models have been shown to set the upper and lower bounds of the elastic

moduli for most particulate micro- and nano-composites [262]. It has been shown that all

experimental and predicted values (using models like Kerner [267], Counto [268], Guth [269]

and Paul [270]) fall between these bounds:

e

eecLower

pV

VEEE

1 (4.23)

ece

eecUpper

pVEE

VEEE

1 (4.24)

where E is the elastic modulus and the subscripts p, e and c stand for particle, epoxy and

composite, respectively. Figure 4.21 shows an upper and a lower bound for the elastic

modulus of EP particles at each foam density. To quantify the contribution of EP particles in

the resulting foam stiffness, the effective elastic modulus of the foams as a function of foam

density is illustrated. Note that the Reuss model assumes a state of uniform stress and the

Voigt model a state of uniform strain. In solid materials, a perfectly elastic 100% dense

system of interacting grains, where each grain is anisotropic and under the Reuss condition,

suffers strain (or displacement) singularities or gaps at the grain boundaries. Likewise, the

Voigt condition would lead to stress singularities in the grain boundaries [271]. With regards

to the results presented in Figure 4.21, two things can be inferred. Firstly, the EP/epoxy foams

show Reuss-like behaviour similar to metals, but are atypical of non-plastic materials. In

Page 164: Elastic And Mechanical Properties Of Expanded Perlite And ...

140

metals, a Reuss-like state can be maintained because deformation which would lead to

displacement singularities can be accommodated through localised plastic deformation. In a

similar manner, in EP/epoxy foams the accommodation is through local EP cell wall collapse

at the boundaries between particles. The local failure associated with the cell wall collapse

releases the stress concentration without affecting the whole structure of the EP particle,

which on average behaves in a Reuss-like fashion. Secondly, the particles make a significant

contribution to the resulting elastic modulus (or stiffness) of the EP/epoxy foams.

Figure 4.21. Predicted elastic moduli of EP particles using the Voigt and Reuss models. To

quantify the particles’ contributions to the stiffness of the foam, the effective elastic modulus of the

EP/epoxy foams as a function of foam density is presented.

Page 165: Elastic And Mechanical Properties Of Expanded Perlite And ...

141

Damage analysis

The macroscopic observations taken during the tests coupled with post-test microscopic

observations were used to understand the deformation mechanisms of the EP/epoxy foams

under compressive loading. Figure 4.22 shows a schematic representation of the damage

sequence in the EP/epoxy foam samples, which is irrespective of their density and type.

However, the strain of occurrence was different for each type of foam, as denoted by the

strain range at which each damage step occurred.

Type 1 ε =5.0-6.8% ε =7.9-10.0 % ε =11.0-13.0 % ε =28.0-40.0 %

Type 2 ε =5.6-7.2% ε =8.0-11.0% ε =12.0-13.0% ε =30.0-46.0%

Type3 ε =3.8-4.2% ε =5.3-6.2% ε = 8.9-10.0% ε =27.0-28.0 %

Figure 4.22. Schematic representation of failure in EP/epoxy foams.

As can be seen, damage started and developed faster in the type 3 foam samples. This can be

explained by two factors: interfacial bonding between the particles and the epoxy binder, and

the volume fraction of epoxy. Microscopic observations (explained in Section 4.6.1) showed

a) b) c) d)

Page 166: Elastic And Mechanical Properties Of Expanded Perlite And ...

142

that there is better interfacial adhesion between the particles and the binder as the EP particles

get smaller in size. Foams of type 2 and 1 have the higher resistance to the formation and

propagation of cracks than foams of type 3. This may be due to the better interfacial adhesion

between the particles and the binder and the higher volume fraction of epoxy in comparison

with the type 3 foams. According to the macroscopic observations, cracks started to develop

in the foam when the stress reaches its maximum, at strains of about 3.8 to 7.0% (Figure 4.22

(a)). Cracks on the wall of the sample developed in the direction of the applied load, while

shear type cracks originated from the upper and lower rims. After reaching the maximum

stress, inelastic deformation occurred as the sides of the sample barrelled outward under

compressive loading. The barrelling effect became more prominent in the samples with larger

particle size. For foams of the same type, the barrelling effect reduced as the density

increased. As the compressive load increased, the cracks on the sample wall joined together

and caused the formation of longitudinal cracks (see Figure 4.22 (b)) due to secondary tensile

stresses. These secondary tensile stresses are normal to the applied load and are generated

due to transverse deformation (Poisson effect) [272]. Longitudinal cracks resulted in

longitudinal splitting along the rim of the sample. At the same time, shear type cracks

propagated through the foam and converged at the central part of the sample where horizontal

cracks were generated due to the pure compression (see Figure 4.22 (c)). This type of shear

failure is common in brittle materials such as ceramics and concrete during compression

testing due to the restraining effect of friction on the upper and lower surfaces. The eventual

joining of the longitudinal splitting and shear cracks in the sample gave rise to the formation

of wedge-like fragments on the sides of the samples (see Figure 4.22 (d)). The development

Page 167: Elastic And Mechanical Properties Of Expanded Perlite And ...

143

and formation of the wedge-like fragments can be better understood by analysing the uniaxial

compressive state of stress, assuming frictionless compression:

000

000

00

ij (4.25)

where σ is the applied stress. The state of stress can be decomposed into hydrostatic (σhyd)

and deviatoric (σdev) stress components:

300

03

0

003

2

300

03

0

003

devhydij (4.26)

The deviatoric stress components can be further divided into two parts, both representing a

state of pure shear [124]:

300

03

0

000

000

03

20

003

2

(4.27)

As can be seen, the hydrostatic stress component produces pure compression while the

deviatoric component gives rise to shear stresses in the material. The effect of the deviatoric

and hydrostatic components of applied stress on the damage of EP/epoxy foams can be

observed in the macroscopic pictures such as Figure 4.23, taken from a type 1 sample

compressed to a strain of 12%. In these pictures, the propagation of shear cracks from the

Page 168: Elastic And Mechanical Properties Of Expanded Perlite And ...

144

upper rim and the formation of longitudinal splitting on the wall of the sample can be

observed. Figure 4.23 (b) displays the remnants of the sample, which is typical of all of the

samples after the tests, showing the significant role of shear and compressive stresses on the

failure modes. The shear stress (the deviatoric component) is responsible for the formation

of shear cracks in the foams while the compressive stress (hydrostatic component) gives rise

to the formation of horizontal cracks. Post-test micrographs taken from the failed samples

provide further evidence for the effects of shear and compressive stresses on the failure

modes. Figure 4.24 shows the SEM micrographs taken from the wedge-like side of the

sample. As can be seen, the particles are uncrushed (Figure 4.24 (a)) and cell walls seem to

be fractured along a plane (Figure 4.23 (b)) which could be due to shearing in the sample.

On the other hand, micrographs from the centre of the sample (Figure 4.24 (c)), where

horizontal cracks dominate, show a significant difference. A large proportion of the cell walls

are crushed which is a typical feature of failure under compression.

(a)

Page 169: Elastic And Mechanical Properties Of Expanded Perlite And ...

145

(b)

Figure 4.23. Macroscopic images showing (a) A sample of type 1 compressed to a strain of 12%

(b) A typical remnant of the samples after the test.

(a)

(b)

Page 170: Elastic And Mechanical Properties Of Expanded Perlite And ...

146

(c)

Figure 4.24. SEM images showing (a) Uncrushed EP particles in the wedge-like fractured side of a

failed sample; (b) Cell walls of EP perlite particles fractured along a plane; (c) Crushed cells in the

central region of a uniformly compressed sample.

Page 171: Elastic And Mechanical Properties Of Expanded Perlite And ...

147

Measurement of the elastic moduli of EP/epoxy foams using elastic

wave speed

Elastic properties of the EP/epoxy foams were characterised by means of elastic wave

propagation (compression and shear) along the cylinder axis of the specimen. In addition, a

new series of quasi-static compressive tests were conducted on the EP/epoxy foams in

parallel, as the EP particles previously used (Sections 4.6.1 - 4.6.3) were from a different

batch. Having all of the samples made from the same particle batch ensured that a more

reliable comparison could be made between the properties obtained from the mechanical and

elastic wave tests. Compressive stress-strain curves for the new EP/epoxy foams with

different densities are illustrated in Figure 4.25.

Figure 4.25. Typical stress-strain curve for different EP/epoxy foam densities.

Page 172: Elastic And Mechanical Properties Of Expanded Perlite And ...

148

Stress-strain curves were used to evaluate Young’s modulus of the foams from the unloading

gradient (explained in detail in Section 3.4.2). The results for Young’s modulus obtained

using the quasi-static tests are illustrated as a function of foam density in Figure 4.27 and the

average values are tabulated in Table 4.8. The Young’s moduli obtained for the new series

of EP/epoxy foams were very close to the ones manufactured from another batch of EP

particles, presented in previous Section 4.6.3. Determination of Poisson’s ratio of the foam

by installation of regular strain gauges was not possible. The strain gauges had a tendency to

debond during the compressive test (un-confined one) due to minimal surface contact

afforded by voids on the surface of the foam. Therefore, the work was extended to

characterise the complete elastic properties of EP/epoxy foams by measuring the velocity of

the passage of elastic waves (i.e. longitudinal and shear waves) in the axial direction for a

wide range of foam densities. The longitudinal and shear wave velocities of EP/epoxy foam

at different foam densities are shown in Figure 4.26. Similar to Figure 4.5 for packed EP

particles, both velocities reached a plateau for densities higher than 0.2 g/cm3. This was

investigated in packed EP particle beds (details explained in Section 4.3.3) and was ascribed

to the relatively constant mass of EP particles which resisted crushing at densities higher than

0.2 g/cm3. In addition, the inter-particle porosity was found to be about 50 Vol%, excluding

debris, and in the range 40 - 50 Vol% including debris for compact densities in the range 0.1

- 0.375 g/cm3. In EP/epoxy foams the inter-particle porosity is filled with 3 - 5 Vol% epoxy,

as shown in Figure 4.28. A very similar behaviour to the wave velocities in the EP particle

compacts (Figure 4.5) is indicative of the dominant role of the EP particles in wave

propagation.

Page 173: Elastic And Mechanical Properties Of Expanded Perlite And ...

149

Figure 4.26. The longitudinal wave and shear wave velocities versus foam density.

Figure 4.27. Young’s modulus of EP/epoxy foams versus foam density determined from the

elastic wave speed (○) and mechanical tests (∆); and Young’s modulus of the EP particles versus

foam density determined from the elastic wave speed (◇).

Page 174: Elastic And Mechanical Properties Of Expanded Perlite And ...

150

Table 4.8. Elastic properties of EP particles, epoxy resin and EP/epoxy foams.

Density

(g/cm3)

EP particles EP/epoxy foams

Elastic wave tests1 Mechanical tests2 Elastic wave tests

𝐸 (MPa) 𝜈 𝐸 (MPa) 𝜈 𝐸 (MPa) 𝜈

0.15 35.6 0.277 54.7 - 115.2 0.172

0.20 130.3 0.287 102.2 - 246.7 0.194

0.25 210.3 0.295 145.2 - 359.7 0.219

0.30 275.6 0.300 175.6 - 454.5 0.245

0.35 326.4 0.301 199.7 - 531.6 0.274

0.40 362.5 0.300 223.7 - 591.6 0.305

0.42 372.8 0.299 253.9 - 610.8 0.318

Density Epoxy resin (Diluted with acetone)

(g/cm3) Mechanical tests Elastic wave tests

1.14 𝐸(MPa) 𝜈 𝐸(MPa) 𝜈

1834.7 0.42 2022.4 0.38

1. Elastic properties of the EP particles are given as a function of the corresponding EP/epoxy foam

density.

2. Elastic properties of the EP/Epoxy foams are given as a function of the EP/epoxy foam density

Page 175: Elastic And Mechanical Properties Of Expanded Perlite And ...

151

Figure 4.28. Volume fraction of the epoxy binder in the second series of EP/epoxy foams versus

foam density.

Isotropic elastic properties of the foams were investigated in terms of two elastic constants E

and ν, given by Eqs. (3.6) and (3.7), and illustrated as a function of foam density in Figures

4.27 and 4.30, respectively. For comparison, the Young’s modulus results obtained from

elastic wave tests are plotted with the ones determined from quasi-static mechanical tests

(Figure 4.27). Both plots follow the same qualitative pattern and show that the Young’s

modulus increases with the foam density. However, the measured Young’s moduli from the

elastic wave speed are more than twice the values obtained by mechanical tests. Williams

and Johnson [273] also reported dynamic moduli (measured using elastic wave tests) were

about twice the static moduli for cancellous cell composites. They ascribed this discrepancy

to a difference in the strain rate. The difference in strain rate can influence the results for a

continuous material by about 10 - 15% [274], which is consistent with the measured values

for epoxy resin (Table 4.6), and for granular material by up to about 40% [275]. Nevertheless,

Page 176: Elastic And Mechanical Properties Of Expanded Perlite And ...

152

the observed discrepancy (≥ 100%) cannot be attributed only to the strain rate difference.

There are three other possible hypotheses to explain such a discrepancy. The first is the

deformation mechanism of the EP particles during unloading. The foams contain a high-

volume fraction of EP particles, consequently not all of the interstices between the particles

are filled with epoxy and some particles are in direct contact with each other. In this case (in

the absence of epoxy between particles), Hardin [276] explained that both elastic and plastic

deformation occurs during the unloading of particulate materials and strain cannot be

separated into discrete elastic and plastic components. This will in turn adversely affect the

Young’s modulus measured using quasi-static mechanical tests.

The second possible explanation for this discrepancy could be the strain level at which

Young’s modulus was measured by the mechanical and elastic wave tests. It was

hypothesised that loading to 70% of the maximum stress (ISO 13314), which was about 3%

strain, could deteriorate the microstructure of EP/epoxy foams by local failure of EP particle

cell walls which leads to lowering the stiffness of the material. On the contrary, the material

displacements produced by the passage of elastic waves are in the order of a few angstrom

[277], and hence Young’s modulus was measured at very low strains and not affected by

microstructural deterioration. To test this hypothesis, one sample with a density 0.26 g/cm3

was loaded to 70% of the maximum load and while the load was held constant, longitudinal

elastic waves were propagated through the sample and Young’s modulus was measured. The

result, which is designated by a filled black circle in Figure 4.27, not only does not show a

lower Young’s modulus value in comparison with previously measured ones but a slightly

higher value than the average ones. This result invalidates the above hypothesis and shows

Page 177: Elastic And Mechanical Properties Of Expanded Perlite And ...

153

that loading to 70% of the maximum load does not adversely affect the Young’s modulus

and should even provide slightly higher moduli by compressing the foam to a higher density.

This is also demonstrated by conducting cyclic compressive tests on a sample with a density

of 0.26 g/cm3. The specimen was loaded to 2.5% of the maximum load, and then unloaded

at the same crosshead speed to 2N load. The loading-unloading continued to 70% of the

maximum load with a load increment of 2.5%, as shown in Figure 4.29 (a). It can be seen

that as the stress increases, the gradient of the unloading path gets steeper which is measured

and presented in Figure 4.29 (b). It clearly shows that loading to 70% of the maximum load

does not detoriorate the internal microsturcture of EP/epoxy foam but makes the foam stiffer.

Nevertheless, these results show that ISO 13314 cannot give a unique Young’s modulus for

EP/epoxy foams, but only the particular elastic modulus of the material at 70% of the

maximum load.

The third hypothesis relates to the deformation of the epoxy ligaments, as well as the

deformation of the perlite cell struts. In mechanical testing where the effective Young’s

modulus (EMech) is determined from the unloading gradient between 70% and 20% of the

maximum load, the unloading displacement is proportional to the compliance MechE

1. As

Young’s modulus measured by the elastic waves (EWave) is bigger than the Young’s modulus

measured by mechanical tests EWave > EMech, the mechanical testing case is more compliant

MechWave EE

11 . If the true compressive modulus is taken as

WaveE , then a larger unloading

displacement (mechanically) is observed than was expected from the pure compressive value,

Page 178: Elastic And Mechanical Properties Of Expanded Perlite And ...

154

which would be proportional to WaveE

1. One simple way for a mechanical system to show a

large elastic displacement is in bending or buckling. The epoxy ligaments are oriented so as

to respond by bending/buckling (n = 1, both ends of a column pinned). However, wave

velocity measurements use small displacements so that direct compressive displacement is

the same as the bending displacement. A similar situation occurs within EP particle struts.

This is supported by the experimental results in Sections 4.3.2 and 4.3.3 where the

constrained modulus of packed beds of EP particles was measured using quasi-static

mechanical and elastic wave tests. That result also showed that for the density range 0.17 -

0.37 g/cm3, the constrained modulus measured using elastic waves was higher than that

measured by the mechanical tests by up to about 25%. Therefore, the combined effects of the

strain rate difference, yielding of EP particles during unloading, and the buckling

deformation in both EP particles and epoxy ligaments may result in a much lower

‘mechanical test’ modulus than the one measured by elastic wave tests.

Notwithstanding, both these properties (i.e. static and dynamic moduli) are useful. For

applications where the material is subjected to compressive load, the stiffness of the material

is characterised by Young’s modulus obtained using quasi-static mechanical tests. On the

other hand, when they are used in a vibrating structure, for example, for sound absorption in

buildings, inside the fuselages of aeroplanes and in machinery enclosures, the true elastic

properties measured by the elastic wave tests are of importance.

Page 179: Elastic And Mechanical Properties Of Expanded Perlite And ...

155

(a)

(b)

Figure 4.29. (a) stress-strain curves and (b) Young’s modulus measured from the gradient of the

unloading path in Figure 4.29 (a) at different stress level in cyclic compressive tests conducted on a

sample with density of 0.26 g/cm3.

To evaluate the relative contribution of EP particles to the stiffness of EP/epoxy foams,

Young’s modulus of the EP particles measured in Section 4.3.3 were plotted with those of

the EP/epoxy foams in Figure 4.27. For foam densities of 0.15 g/cm3 and 0.2 g/cm3, the EP

Page 180: Elastic And Mechanical Properties Of Expanded Perlite And ...

156

particles contribute to the stiffness of the EP/epoxy foams by 12% and 49%, respectively.

For higher foam densities, however, this contribution reached a constant value of about 62%.

There are three main factors contributing to the stiffness of EP/epoxy foams: i) epoxy binder,

ii) EP particles, and iii) the interaction between the EP particles and epoxy binder. The role

of the epoxy binder in the stiffness of EP/epoxy foams seems to be almost constant with

respect to the constant range (3 - 5%) of the epoxy volume fraction in EP/epoxy foams, as

shown in Figure 4.28. However, the contribution of EP particles and the interaction between

the EP particles and epoxy binder appear to be variable. This might be ascribed to the number

of contact points and the contact size (or contact surface area) among the EP particles as well

as the contact size between the EP particles and the epoxy binder. As particles are crushed

into smaller particles during foam densification, the number of contact points per unit volume

as well as the contact size will increase among the particles, and also between the EP particles

and the binder. These result in more mechanical bonds between the constituents and hence

an increase in the stiffness of EP/epoxy foams. In low density foams (< 0.2 g/cm3), a greater

portion of large EP particles exist (see Figure 4.7) and thus a smaller number of contact points

and smaller contact size are present in comparison to foams at higher densities. Beyond 0.2

g/cm3, the intra-particle volume becomes constant. This will be discussed in Section 5.4.

The Poisson’s ratio of EP/epoxy foams, corresponding to the wave speed measurements in

Figure 4.26, were calculated using Eq. (3.7) and are presented in Figure 4.30. It can be seen

that Poisson’s ratio of the foams monotonically and almost linearly increases with respect to

the foam density. This is indicative of the material’s increasing resistance to change in

volume rather than to distort under uniaxial load.

Page 181: Elastic And Mechanical Properties Of Expanded Perlite And ...

157

Figure 4.30. Poisson’s ratio of epoxy resin (­­­­), foam (○) and packed beds of EP particles ( ⃣ )

versus the foam density.

To investigate the effect of the constituent materials on the foam Poisson’s ratio, Poisson’s

ratio of EP particles and the cured diluted epoxy resin are also plotted on the same figure.

Poisson’s ratio of the foams begins lowest, intersects with the Poisson’s ratio of EP particles

at 0.38 g/cm3, and approaches the Poisson’s ratio of the epoxy binder with increase in the

foam density. Conversely, Poisson’s ratio of the packed EP particles is almost constant with

increase in the foam density. However, when the particles and epoxy are combined, the

resultant Poisson’s ratio is far from constant. Without epoxy, rotation can occur about the

points of contacts between the particles – the inclusion of the epoxy elastically constrains

these motions. This situation is similar to the assumptions considered in [278] concerning the

elastic behaviour of granular materials as assemblies of randomly distributed particles. Figure

Page 182: Elastic And Mechanical Properties Of Expanded Perlite And ...

158

5 in [278] shows that Poisson’s ratio increases by decreasing ks/ kn ratio, where ks is the

resistance of a contact between two particles to shear displacement and kn is the resistance

of a contact to compression along a line connecting the centre of the two particles. The

decrease in ks/ kn can be due to: i) increase in kn, ii) decrease in ks, or iii) concurrent increase

in kn and decrease in ks. At this stage, it is suggested that the decrease in ks/ kn is responsible

for the observed increase in Poisson’s ratio with increase in foam density, however it is not

clear which case (i, ii or iii) is active. This will be investigated by numerical simulation in a

future study.

Page 183: Elastic And Mechanical Properties Of Expanded Perlite And ...

159

5 Chapter Five: Discussion

Introduction

In Chapter 4, there was a very focused discussion of each specific result. In this chapter, the

results from the previous chapters will be discussed in a broader context, which includes: a)

Comparison of the manufacturing methods for EP/epoxy foams with other syntactic foams;

b) Comparison of Young’s moduli of packed EP particle beds found experimentally using

the oedometric tests and elastic wave tests, as well as the ones predicted using the Voigt and

Reuss models; c) Analysis of the method for estimation of the volume fraction of epoxy

binder in the EP/epoxy foams; d) Comparison of the compressive properties of EP/epoxy

foams with the foams discussed in the literature; e) Comparison of the damage modes in

EP/epoxy foams under compressive loading with a variety of existing foams.

Page 184: Elastic And Mechanical Properties Of Expanded Perlite And ...

160

Manufacturing method

In this study, EP/epoxy foams were manufactured by the buoyancy method (AU Patent No.

2003205443 [139]), as described in Section 3.3.4. Alternative methods for manufacturing

foams were reviewed in Section 2.3.2. In comparison with those methods, the buoyancy

method used here has several advantages. These include avoiding breakage of the fragile EP

particles during mixing with the epoxy prior to pressing. This is essential for the very low

density foams but is also useful for the high density foams. In addition, this method offers

the homogeneous distribution of EP particles in the matrix, the feasibility of manufacturing

foams with density as low as 0.15 g/cm3 and ease of manufacture.

On the other hand, this method has some drawbacks. In this method, the mixture of EP

particles and diluted epoxy binder creates a two-phase system where just the top phase is

used and the bottom phase is disposed of. Even though 2 - 5 vol% of the epoxy binder is

present in the EP/epoxy foams, a large amount of acetone is required to dilute the epoxy

binder (the mixing ratio is explained in Section 3.3.4). This is a large waste of materials as

the residual bottom phase, including a large amount of acetone, is thrown away. The second

problem is the curing time which is significantly longer (40 days) than the competing

methods mentioned in Section 2.3.2 which require from less than 2 days and up to 7 days.

The third problem is compressing EP particles of different mass within a constant volume to

obtain different densities. The compression causes breakage and crushing of EP particles and

thus reduces the properties which are pertinent to the cellular structure of EP particles, such

as the thermal and acoustic insulating properties.

Page 185: Elastic And Mechanical Properties Of Expanded Perlite And ...

161

Of the alternative methods explained in Section 2.3.2, two methods are considered suitable

for EP/epoxy foams. The first one involves the filling of a mould with a desired quantity of

particles and pouring a pre-measured amount of epoxy over the particles. The other is to fill

a mould completely with particles and then measure their mass. This is followed by mixing

the particles with an epoxy solution and then transferring the mixture to a mould. Both

methods have the advantage of using a small amount of acetone which is contrary to the

buoyancy method. In addition, the first method has the advantages of avoiding particle

fracture and having accurate information regarding the volume fraction of particles and

binder. The second method also has the advantage of a homogeneous distribution of particles

within the foam as well as restricting the particles from floating to the surface during foam

production. These methods, however, have several disadvantages in addition to those

mentioned in Section 2.3.2. The first method has the potential for an inhomogeneous

distribution of particles owing to the separation of phases due to differences in the constituent

densities. Moreover, there is a possibility of binder concentrating in the lower part of the

foam due to gravity, especially as the particles are free to move and do not completely fill

the mould. The second method has a high potential to break particles during mixing with the

epoxy binder before the moulding process, which is especially of concern for manufacturing

of the lowest density foams.

Taking these factors into consideration, the first method, with the constraint of filling the

mould completely, has the greater potential for EP/epoxy manufacture. Here, different foam

densities could be achieved by using different particle sizes, which have different densities

Page 186: Elastic And Mechanical Properties Of Expanded Perlite And ...

162

in Figure 4.8 (a), within a constant volume. In addition, using a low viscosity epoxy is

suggested rather than diluting standard epoxy with acetone.

Page 187: Elastic And Mechanical Properties Of Expanded Perlite And ...

163

Young’s modulus of packed EP particles

Considering the high-volume fraction of EP particles in the EP/epoxy foams, the Young’s

modulus of packed beds of EP particles were required to be determined. For this purpose,

oedometer tests (confined quasi-static compressive tests) were conducted on packed EP

particle beds, as detailed in Section 4.3.2. However, since the particles in this test were

confined, the calculated unloading gradient was equal to the constrained modulus of the

packed particle beds and a function of both their Young’s modulus (EP) and Poisson’s ratio

(𝜈), given by Eq. (4.5). As it was not possible to obtain Poisson’s ratio by oedometer tests,

attempts were made to estimate the Young’s modulus of packed EP particle beds using the

Voigt and Reuss models (i.e. Eqs. (4.23) and (4.24)) which give an upper 𝐸𝑝𝑈𝑝𝑝𝑒𝑟

and a lower

𝐸𝑝𝐿𝑜𝑤𝑒𝑟 bound, shown in Figure 4.18. To characterise the complete elastic properties, the

work was extended to investigate the elastic properties of packed EP particle beds by

measuring the velocity of the passage of elastic waves (i.e. compression and shear) in the

axial direction for a wide range of compaction densities.

Compression and shear wave velocity measurements and Eqs. (3.4) and (3.5) were used to

measure the Young’s modulus and Poisson’s ratio of a packed bed of EP particles at each

compact density (see Figure 4.9). Having determined the Poisson’s ratio of the EP particle

beds, the constrained moduli obtained earlier using the oedometer tests were converted to

Young’s modulus, hereafter referred to as the quasi-static Young’s modulus. However, this

Young’s modulus is an estimate as the Poisson’s ratios obtained by the elastic wave tests

would be slightly different from the ones that could be obtained from quasi-static testing;

Page 188: Elastic And Mechanical Properties Of Expanded Perlite And ...

164

Poisson’s ratio from a quasi-static test is expected to be slightly higher. It has been found that

Young’s modulus is dependent on the strain rate or frequency of the tests [1]. Generally,

Young’s modulus increases with strain rate or frequency of the tests. This will in turn result

in a lower Poisson’s ratio, as shown for the epoxy in Table 4.6, which was attributed to the

difference in the deformation rate of the specimens [2, 3].

For comparison, the estimated Young’s modulus of packed EP particles from the quasi-static

tests were re-plotted with the ones predicted using the Voigt and Reuss models as a function

of the foam density, as shown in Figure 5.1. In addition, Young’s moduli obtained using the

elastic wave tests, in terms of both the experimental compact density and the modified

density, are shown in Figure 5.1. As the density is not involved in the calculation of Young’s

modulus from the mechanical results, the mechanical results cannot be modified to reflect

the correction of density excluding debris.

For the most part, the mechanical test results lie in between the two wave speed determined

moduli which could be considered as upper and lower bounds for constrained moduli of EP

particle compacts. For a compacted density of 0.1 g/cm3, the constrained moduli determined

from mechanical tests is higher than those determined by elastic wave tests. This indicates

that an additional resistance to displacement is present in the mechanical tests, i.e. the

unloading response observed in the previous work was not truly elastic. It has been widely

discussed in the literature that particulate materials yield during unloading, and consequently

the true elastic behaviour is restricted to infinitesimal increments of unloading [4, 5].

Therefore, the measurement of elastic properties, especially in low density (high porosity)

compacts where the size of inter-particle contacts is very small is best performed by elastic

Page 189: Elastic And Mechanical Properties Of Expanded Perlite And ...

165

wave velocity measurements. In addition, Young’s modulus for a density of 0.1 g/cm3

measured by quasi-static tests do not fall within the bounds of the Voigt and Reuss models

which are a simple test of validity. The predicted values of Young’s modulus for most

particulate micro- and nano-composites [6] are found to fall within the identified upper and

lower bound by the Voigt and Reuss models. The predicted values which fall within these

bounds are considered to be valid. Hence, with regards to the elastic wave Young’s moduli

and Voigt and Reuss estimations, the quasi-static Young’s moduli for the compacts with a

density higher than 0.15 g/cm3 are considered to be valid.

Figure 5.1. Young’s moduli of the packed EP particle beds measured using mechanical tests and

estimated using the Voigt and Reuss models.

Within this range (0.15 - 0.37 g/cm3), the Young’s moduli determined by elastic waves are

higher than those determined by the mechanical tests. As these two types of measurement

impose strains that differ by 3 - 5 orders of magnitude and occur at widely differing strain

rates, they explore different aspects of a material’s behaviour [7, 8] which can influence the

Page 190: Elastic And Mechanical Properties Of Expanded Perlite And ...

166

results for a continuous material by about 10 - 15% [7], and for a granular material by up to

40% [9]. Nevertheless, the lower and upper bounds determined by the experimental and

modified Young’s moduli using elastic-wave tests appear to provide a reasonable range in

which the Young’s moduli of packed EP particles can be explored.

Page 191: Elastic And Mechanical Properties Of Expanded Perlite And ...

167

Structural characterisation of EP/epoxy foams

The manufactured EP/epoxy foams had a structure of a packed EP particle bed with the epoxy

resin partially filling the interstices. Therefore, the experimental results on the packed EP

particle beds presented in Section 4.3.3 helped to understand the structure of packed EP

particles and hence the EP/epoxy foams. As mentioned in Section 4.6.3, since the properties

of EP/epoxy foams show almost no dependency on particle size, type 2 foams were chosen

and the rest of the tests, including the elastic wave speed test, were conducted on this type of

foam. Hence, this discussion on the structure of the EP/epoxy foam is based on experiments

conducted on only the type 2 foams, with the assumption that other foam types behave

similarly.

Using the calculated inter-particle porosity shown in Figures 4.8 (b) and (c), the volume

fraction of EP particles in EP/epoxy foams was calculated and is presented in Figure 5.2 (a).

It can be seen that the volume fraction of EP particles increase with the foam density,

reaching a peak at a density of 0.375 g/cm3, and then start to decrease for higher foam

densities. This can be explained by fracturing of more EP particles into smaller particles and

debris, as shown in Figure 4.7, which results in a reduction of the volume of EP particles and

hence their fraction in the EP/epoxy foams at high densities (> 0.375 g/cm 3).

In addition, the percentage of porosity at each foam density was calculated as the sum of the

inter-particle porosity and the intra-particle porosity. The inter-particle porosity at each foam

density was calculated by subtracting the sum of the volume fraction of the EP particles

(Figure 5.2) and the epoxy binder (Figure 4.17) from the total sample volume:

Page 192: Elastic And Mechanical Properties Of Expanded Perlite And ...

168

sample

epoxyParticlesample

V

VVVPorositycleInterparti

)( (5.1)

Intra-particle porosity was calculated by:

i

perlitesolid

EPi fPsizeparticleoneinsidePorosity *1

(5.2)

n

i

iPPorositycleIntraparti1

(5.3)

where ρEP is the density of an EP particle of a certain size presented in Figure 4.8 (a),

𝜌solid perlite is the density of the pore free perlite given in Table 4.1, and fi is the fraction of

that particle size within a compact. The calculated inter-particle porosity from Eq. (5.1) and

the intra-particle porosity from Eq. (5.3) were added together and the total porosity of the

EP/epoxy foams at each density was calculated and is presented in Figure 5.2 (b).

The volume percentage of epoxy in the EP/epoxy foams was estimated using a method

introduced in Section 4.6.2.1, and the results for three types of EP/epoxy foams were

presented in Figure 4.16. In addition, the epoxy volume fractions in three types of EP/epoxy

foams were measured based on experimental data (explained in detail in Section 4.6.2.2) and

are shown in Figure 4.17. Both the results showed that the volume fraction of epoxy is

influenced by the EP particle size. However, while the estimated values show significant

change in epoxy volume fraction with density, the experimentally measured ones show only

a slight change in epoxy volume fraction with density. Moreover, the experimentally

measured values show that the measured epoxy volume fractions for the type 1 foams were

slightly higher than the type 2 and type 3 foams, and that the measured values for the type 2

foams were slightly higher than the type 3 ones. This is reasonable given the fact that smaller

Page 193: Elastic And Mechanical Properties Of Expanded Perlite And ...

169

particles have a bigger surface area per unit volume and thus a larger amount of epoxy is

required to cover those surfaces.

(a)

(b)

Figure 5.2. (a) Volume fraction of EP particles and (b) Porosity within inter-particle space (by

consideration of debris) and total porosity in type 2 EP/epoxy foams.

Page 194: Elastic And Mechanical Properties Of Expanded Perlite And ...

170

The estimated values (Figure 4.16) were reasonable for the average behaviour of type 1 and

2 foams of, however they show a significant deviation with the experimentally measured

values for type 3 foams. This indicates that there is a particle size limitation, and that the

estimation method does not give a reasonable trend for the foams made with EP particles

larger than 2.8 mm.

Considering the epoxy volume fraction, the inter-particle and intra-particle porosity as well

as the investigation of SEM images, it can be concluded that EP/epoxy foams have a structure

which is schematically drawn in Figure 5.3. The porosity inside the EP particles is not drawn

but the figure shows how the inter-particle porosity is distributed. In addition, it shows that

the epoxy resin is drawn to the EP particles by surface tension while keeping the inter-particle

space relatively open. Hence, it can be seen that low density EP/epoxy foams have a similar

interior structure to high density foams, but at a different scale. Moreover, it should be noted

that though the particle volume fraction in different density foams varies in the range 46 -

58%, the resultant foam density is mainly caused by the density of the EP particles. The

higher the foam density, the larger the portion of smaller particles and debris that were shown

to have higher densities (Figure 4.8 (a)). Accordingly, the total foam porosity decreases from

91 Vol% to 79 Vol% while the inter-particle porosity reduces from 50 Vol% to 37 Vol%.

This clearly shows that the intra-particle porosity is constant for foams of different density

( 41 - 42 Vol%). Therefore, while the experimental results on the packed EP particles showed

(Figure 4.8 (b)) that the inter-particle porosity was almost constant excluding debris, the real

compact structures, which included the debris, had a constant intra-particle porosity. Overall,

the porous structure of the EP particles contributed to the porosity of the EP/epoxy foams by

Page 195: Elastic And Mechanical Properties Of Expanded Perlite And ...

171

as much as about 42% and the rest of the foam porosity is related to the packing of EP

particles and the void spaces between particles.

(a) (b)

Figure 5.3. A schematic representation of the internal structure of (a) low density and (b) high

density EP/epoxy foams.

Page 196: Elastic And Mechanical Properties Of Expanded Perlite And ...

172

Compressive behaviour and compressive properties

A schematic representation of the compressive stress-strain curves for elastomeric, elastic-

plastic, and brittle single-phase foams (i.e. foams that are manufactured from one type of

material) is shown in Figure 5.4. They show linear elasticity at low stress levels, followed by

a long stress plateau regime truncated by a regime of densification where the stress rises

steeply. The gradient of this regime (the initial slope of the stress-strain curve) is considered

as Young’s modulus. The linear elastic region is followed by a plateau region which

characterises the energy absorption capacity of a foam through the deformation of cells. The

final regime of densification is associated with the touching of opposite collapsed cell walls

and if compression continues to a higher strain (called densification) the stress-strain curve

should reach the elastic response of the solid material from which the cell walls are made.

(a) (b) (c)

Figure 5.4. Schematic representation of compressive stress-strain curves for a) an elastomeric, b) an

elastic-plastic, c) a brittle foam [11].

Page 197: Elastic And Mechanical Properties Of Expanded Perlite And ...

173

As can be seen, no drop after the peak stress has usually been observed in the stress-strain

curves for elastomeric foams (Figure 5.4 (a)), while a small drop in elastic-plastic foams

(Figure 5.4 (b)) and a significant drop in the brittle foams or foams with little plasticity

(Figure 5.4 (c)) has been observed. It has also been found that increasing the foam density

increases the Young’s modulus and plateau stress, but shortens the strain at which

densification occurs [10, 11].

Similar to the single-phase foams explained above, the stress-strain curves for syntactic

foams start with an approximately linear region corresponding to the elastic behaviour of the

foam until the peak stress is reached. Peak stress designates the point of crack initiation. In

some stress-strain curves, the peak stress is followed by a sharp stress drop while in others

there is only a slight decrease in the stress level. The sharper stress drop suggests that it is

more difficult for cracks to initiate in the matrix which is more prominent as the matrix

volume fraction and the strain rate increases [12]. On the contrary, increasing the volume

fraction of hollow microspheres decreases the stress drop after the peak stress [13]. These

imply that the failure initiation in syntactic foams does not depend on the strength of the

hollow microspheres, but is mainly related to the properties of the polymeric matrix. After

this decrease, a plateau region starts at which the stress becomes nearly constant for further

compression. The plateau region corresponds to the energy absorption in the process of the

collapse of cells comprising a microsphere and its surrounding epoxy resin. Syntactic foams

with higher volume fractions of hollow microspheres show a larger plateau region. When a

significant portion of the microspheres are crushed, further compression results in the

densification of the foam which is visible as an upward trend in the stress-strain curve

Page 198: Elastic And Mechanical Properties Of Expanded Perlite And ...

174

following the plateau region. This point is considered the failure point for syntactic foams

because at this point most of the load bearing microspheres have been crushed. It is worth

noting that these three distinct regions are found to be influenced by the strain rate of the test

and the geometry of the samples. As the strain rate increases, the plateau stress level and

energy dissipation capacity of the foam increases [14, 15]. Gupta [12, 16] investigated the

effects of the specimen aspect ratio on the fracture behaviour and compressive properties. It

was found that the peak compressive strength depends on the mechanical properties of the

hollow microspheres and matrix resin, and is independent of the specimen aspect ratio.

However, the specimen behaviour during compression shows remarkable differences with

respect to the aspect ratio. Specimens with low aspect ratio show no drop after the peak stress

and show a shorter plateau region, while specimens with high aspect ratio either show a small

decrease after peak stress and a larger horizontal plateau region [16] or a large drop and no

plateau regime [12].

A schematic representation of the stress-strain curves in the current study is presented in

Figure 5.5. The damage mechanisms causing these features will be discussed in Section 5.6.

Similar to the foams explained above, there is a linear elastic region ending in a peak stress.

The peak stress is followed by a significant drop in stress i.e. strain softening. Similar

behaviour was noticed in the compressive stress–strain curves of expanded perlite/starch

foams [17], cenosphere/epoxy syntactic foams [12] and pumice/epoxy composites [18].

Following the strain softening, the stress becomes nearly constant for further compression in

the plateau region. Contrary to the single-phase foams and syntactic foams discussed

previously, there is no densification regime in the stress-strain curves of EP/epoxy foams.

Page 199: Elastic And Mechanical Properties Of Expanded Perlite And ...

175

The point at the end of the plateau where the stress started to decrease continuously is the

failure point (designated in Figure 5.5) for EP/epoxy foams. At this point, the major portion

of the structure has collapsed and the remnants are incapable of bearing substantial amounts

of load. The compressive stress-strain curves (see Figure 4.18) also show that an increase in

a foam’s density increases the peak stress, Young’s modulus, the energy absorption capacity

and hence the modulus of toughness of the EP/epoxy foam. This behaviour is similar to the

single phase foams and syntactic foams discussed above.

Figure 5.5. Schematic representation of the compressive stress-strain curve for EP/epoxy foams.

To compare the compressive properties of the newly manufactured perlite-based foams with

currently available foams, Ashby diagrams of the average compressive strength and modulus

of the foams as a function of density are presented in Figure 5.6. The perlite-based foams are

illustrated by small red ellipses which set them apart from the other foams. A line with a

slope of one is drawn in both Figures 5.6 (a) and (b). All materials lying on lines with this

Page 200: Elastic And Mechanical Properties Of Expanded Perlite And ...

176

slope have the same specific strength (𝜎𝐶

𝜌) or specific modulus (

𝐸𝐶

𝜌), respectively. Specific

properties provide a better comparison of different foams having different densities. As

shown in Figure 5.6 (a), the specific compressive stress of the newly developed foams in the

density range 300 - 440 kg/m3 (0.300 - 0.440 g/cm3) is comparable with that of alumina

foam (0.745 g/cm3), aluminium-silicon carbide foam (0.27 g/cm3), closed cell Phenolic foam

(0.035 g/cm3), and closed cell polypropylene (PP) foam (0.030 g/cm3). The specific modulus,

however, is lower than alumina foam and aluminium-silicon carbide foam and higher than

the other two above-mentioned foams. In addition, the specific compressive stress of the

EP/epoxy foams in the density range of 300 - 440 kg/m3 (0.300 - 0.440 g/cm3) is comparable

with EP/starch foam (0.3 g/cm3 and 0.375 g/cm3) [17] and EP/sodium silicate foam

containing 0.20 g/ml sodium silicate (0.3 g/cm3 and 0.4 g/cm3) [19] while both foams have

lower specific compressive modulus than that of EP/epoxy foams. Additionally, Figure 5.6

(b) does shows that the specific compressive modulus of perlite-based foams in the density

range of 155 - 440 kg/m3 (0.155 - 0.440 g/cm3) is comparable with that of rigid closed cell

polyurethane foam (0.24 - 0.6 g/cm3), closed cell polyethylene terephthalate (PET) foam

(0.108 - 0.15 g/cm3), closed cell Phenolic foam (0.12 g/cm3), hollow glass microsphere/epoxy

syntactic foams (0.139 and 0.27 g/cm3), and closed cell polystyrene (PS) foam (0.05 g/cm3).

Therefore, the newly developed perlite-based foams may be adaptable for applications where

similar or better properties can be achieved using this economically advantageous material.

Page 201: Elastic And Mechanical Properties Of Expanded Perlite And ...

177

(a)

(b)

Figure 5.6. Compressive strength (a) and compressive modulus (b) plotted against density for

currently available foams (Ashby et al., 2000) and results obtained for perlite-based foams ( ).

Page 202: Elastic And Mechanical Properties Of Expanded Perlite And ...

178

Damage mechanisms under compressive loading

Post-test microscopy observations, coupled with macroscopic observations taken during the

tests, revealed the presence of three different failure modes (i.e. longitudinal splitting, shear

failure, and compression failure) for all of the EP/epoxy foams regardless of their particle

size and density. However, the strain to activate each mode was different for each foam type.

In addition, these observations helped to understand the damage sequence and its correlation

with the peak stress, strain softening, plateau region and failure point of the stress-strain

curves for EP/epoxy foams, shown in Figure 5.5. The peak stress is the point at which cracks

started to appear in the specimens, due to shear stresses in the form of shear type cracks

originating from the sample corners, by secondary tensile stresses in the form of longitudinal

cracks, and by compressive stresses in the form of horizontal cracks in the sample centre (as

explained in Section 4.7). The strain softening might be due to the combined effects of the

very low density of the EP/epoxy foams, the small EP cell wall thickness (0.512 - 0.532 µm),

the brittle nature of the EP particles, and the brittleness and low volume fraction of the epoxy

matrix. Moreover, the development and propagation of longitudinal cracks in the samples

causes disintegration of the sample structure which was even more prominent in samples

with lower density due to easier lateral deformation under secondary tensile stresses. These

together make the foam incapable of bearing a substantial amount of compressive load as the

strain increases. The plateau region corresponds to where: i) shear cracks propagated and

meet horizontal cracks in the sample centre; ii) longitudinal cracks meet and form

longitudinal splitting; and iii) the remaining part of the sample, which is in the shape of an

Page 203: Elastic And Mechanical Properties Of Expanded Perlite And ...

179

hourglass, underwent uniform compressive deformation. The failure point which led to the

end of the plateau region corresponds to the formation of wedge-like fragments by the

intersection of the longitudinal splitting and shear cracks in the sample and the completion

of the collapse of the EP particles along with the surrounding resin matrix.

It is worth noting that similar modes of failure were observed in cenosphere/epoxy [12] as

well as in hollow glass microspheres/epoxy syntactic foams under compressive load [16, 20].

On the other hand, in perlite/A356 syntactic foams with densities lower than 1.06 g/cm3

similar modes of failure to the EP/epoxy foams were observed [21]. At higher densities,

however, cells deformed layer by layer (e.g. by buckling or bending of the cell walls), in a

uniform manner and no shear or wedge-like failure modes were observed [22]. This was

explained by the brittle behaviour of the cell walls in foams with densities lower than 1.06

g/cm3. In higher density foams, strain hardening occurred in deformed cell walls which

resulted in the successful transfer of stress from the distorted layer of cells to the adjacent

areas and thus uniform growth of damage through the whole sample block. Shear-like failure

has also been commonly observed in metal foams showing ductile behaviour [23]. For

example, in cenosphere/aluminium syntactic foams [24] during compression, deformation

was observed to initiate from the upper and lower corners of the sample and developed along

the plane of shear at an angle of approximately 45° towards the centre of the sample where

compression failure was dominant due to the concentration of the compressive stress in this

region. The sheared zone grew layer by layer during deformation and the maximum

deformation was observed to occur within the sheared zones and in the centre of the samples.

Page 204: Elastic And Mechanical Properties Of Expanded Perlite And ...

180

However, the fully developed wedge-like failure observed in the EP/epoxy foams was absent

[22].

In metal foams with little plasticity, e.g. Zn−22Al alloy foam [25], the deformation

mechanisms mainly consist of brittle crushing and crumbling of cell walls, cell wall tearing

and large shear fracture. In all these foams after an initial linear elastic region, the stress-strain

curve is characterised by a significant drop in load resistance at the onset of plastic

deformation and jagged, oscillatory plateau behaviour associated with the progressive

crushing of brittle cell walls (see schematic representation in Figure 5.4 (c)). This is the

opposite of what has been observed in ductile metal foams which show little or no drop in

load at the beginning of plasticity and smooth plateau behaviour before densification [11].

Polymeric viscoelastic foams have been shown to behave differently [11, 26-29]. These may

be further categorised as elastomeric, elastic-plastic or brittle foams. In these materials, linear

elastic deformation is controlled by cell struts bending and if they contain closed cells, by

cell wall stretching. Under compressive loading, the elastic deformation is followed by the

collapse of the cells by elastic buckling in elastomeric foams (e.g. EPDM foam [30]), plastic

buckling of cell walls in elastic-plastic foams (e.g. rigid polyvinyl chloride foam [31]), and

brittle crushing of the cell walls and formation of shear fractures in brittle foams (e.g. rigid

polyurethane [32] and epoxy-based polymeric foams [33]). Deformation mechanisms

associated with elastomeric and elastic-plastic foams, particularly the elastic buckling

deformation of cell walls and the formation of plastic hinges, was not observed in the

EP/epoxy foams. However, the crushing of cells in the EP/Epoxy foams is similar to what

has been observed in other brittle foams.

Page 205: Elastic And Mechanical Properties Of Expanded Perlite And ...

181

The role of the different structural components in the failure is also of interest. X-ray

microtomography was recently used in conjunction with finite element modelling in order to

study the progressive collapse of glass hollow sphere reinforced epoxy samples under

compressive loads [34]. In that study, it was clearly demonstrated that the glass spheres near

the centre of the sample are the first to collapse and transfer load to the surrounding epoxy

and spheres. Following this, some sphere collapse along 45 degree planes, analogous to our

shear planes, was also evident. It is possible that the perlite particles within the samples in

the current study underwent a similar collapse sequence, commencing in the central region

leading to the shear cracks and splitting, essentially the reverse of the sequence described in

Section 4.7. However, as the core of the sample is not visible in our work, we are unable to

clarify this at present. In contemplating the damage sequence, it is necessary to note some

fundamental differences between the tomographic study and the work reported here. The

volume fraction of glass cenospheres used was quite low (0.3), the cenospheres had quite

thick walls (15 microns) compared with perlite (0.32 microns), and perlite is a multi-cellular

particle. Therefore, perlite particles under load begin to crush far more easily than

cenospheres, however as multi-cellular particles, they can exist in a variety of intermediate

states between uncrushed and crushed i.e. they can crush progressively, cell by cell.

These comparisons serve to highlight that failure by a combination of shear, compression

and longitudinal splitting observed in this work is characteristic of porous structures

consisting of brittle cell walls, although these may be represented to different degrees in

different systems.

Page 206: Elastic And Mechanical Properties Of Expanded Perlite And ...

182

6 Chapter Six: Summary

Conclusions

This study investigated the potential of expanded perlite particles (EP particles) in the

manufacture of light-weight syntactic foams. The structural, microstructural and physical

properties of EP particles were characterised. Two geometrical relationships were suggested

for relating the micro-structure of EP particles to their macroscopic properties (Section 4.3.1).

In section 4.3.2, the elastic modulus of packed EP particles beds were measured using quasi-

static mechanical tests. However, as the test was conducted in a confined situation, the

calculated unloading gradient was not identical to the Young’s modulus of the packed bed of

particles but a function of both their Young’s modulus and Poisson’s ratio. Hence in section

4.3.3, the elastic properties of packed EP particles beds, in terms of two isotropic elastic

moduli (Young’s modulus and Poisson’s ratio), were characterised using elastic wave speed

measurements along the axial direction. Particle distributions within EP particle compacts

after compaction showed that as the compact density increases, the percentage of larger

particles decreases due to breakage into smaller particles accompanied by the production of

debris. Despite this process, it was found that the inter-particle porosity remains relatively

constant with increasing compact density. The mass % of debris rises to quite high levels

owing to the high mass density of the debris particles, however, the volume of debris never

exceeded 20% of the available inter EP-particle space. Due to the formation of debris, the

Page 207: Elastic And Mechanical Properties Of Expanded Perlite And ...

183

experimental compact densities were modified by deducting the debris mass from the total

mass of EP particles and new densities were calculated. Young’s moduli calculated based on

the experimental and modified density values showed a decrease with porosity. However,

Poisson’s ratio was relatively independent of porosity.

The properties of solid perlite were investigated by sintering powdered perlite into low

porosity solid samples and the application of elasticity theories (Section 4.2). The elastic

properties of solid perlite were found to be very close to those for solid obsidian reported by

Manghnani et al. [20]. Using these properties of solid perlite, the porosity-elastic moduli

relations were investigated in Section 4.4. Four analytical models predicting the elastic

moduli of packed beds of EP particles from the properties of the parent material were

investigated. Although these models were not developed to explicitly handle debris, the

Wang, Rice, and Gibson and Ashby models showed reasonable agreement with the

experimental moduli. However, none of the original models predicted the modified moduli

well although the Wang model and the Gibson and Ashby models gave a reasonable average

trend. It was shown that modifying the Phani model shape factor and expressing it as a

function of porosity can provide a satisfactory means of predicting the elastic properties of

EP particle compacts based on both the experimental and the modified densities.

The ultimate purpose of the EP particles was for the formation of syntactic foams. Thus, the

research was extended in Section 4.6 by producing light-weight foams synthesised by

dispersing EP particles in a matrix of epoxy. Three types of foams containing different sized

particles were fabricated for a density range of 0.15 - 0.45 g/cm3. Quasi-static compressive

tests were conducted on the EP/epoxy foams and the effects of particle size and foam density

Page 208: Elastic And Mechanical Properties Of Expanded Perlite And ...

184

variation on the compressive properties were investigated. The compressive stress and

effective elastic modulus were shown to be independent of the particle size. However, a slight

variation was observed in modulus of toughness as foams made with EP particles in the size

range 2 - 2.8 mm showed a slightly higher energy absorption capacity. On the contrary,

EP/epoxy foams showed a strong dependence of the compressive properties on the foam

density. The strength increased linearly, peaking at 1.77 MPa, whereas the effective elastic

modulus and modulus of toughness increased at parabolically increasing and decreasing

rates, respectively. In Section 4.7, post-test SEM observations coupled with photogrammetry

during the tests revealed the presence of three different failure modes for all of the foams,

regardless of their particle size and density. However, the strain to activate each mode was

different for each foam type.

The specific compressive stress of perlite–epoxy foams in the density range of 0.3 - 0.44

g/cm3 was found to be comparable with that of foams such as alumina (0.745 g/cm3),

aluminium–silicon carbide (0.27 g/cm3), closed cell phenolic foam (0.035 g/cm3), and closed

cell PP foam (0.030 g/cm3). The specific compressive modulus, however, was found to be

lower than alumina foam and aluminium–silicon carbide foam but higher than closed cell

phenolic foam and closed cell PP foam. In addition, the specific compressive stress of the

EP/epoxy foams in the density range 0.300 - 0.440 g/cm3 is comparable with EP/starch foam

(0.3 g/cm3 and 0.375 g/cm3) [186] and EP/sodium silicate foam containing 0.20 g/ml sodium

silicate (0.3 g/cm3 and 0.4 g/cm3) [188] while both foams have lower specific compressive

modulus than that of EP/epoxy foams. Nevertheless, the specific compressive modulus of

perlite–epoxy foams in the density range 0.155 - 0.44 g/cm3 was found to be comparable with

Page 209: Elastic And Mechanical Properties Of Expanded Perlite And ...

185

those of a number of rigid closed cell foams made from polyurethane (0.4 g/cm3), PET (0.15

g/cm3), phenolic (0.12 g/cm3) and PS (0.05 g/cm3) materials.

In the additional work presented in Section 4.8, the elastic properties of EP/epoxy foams

were characterised by means of elastic wave propagation (compression and shear) along the

the cylinder axis of the specimens. By adopting an isotropic model, the Young’s modulus

and Poisson’s ratio were used to characterise the elastic response of the medium. The young’s

moduli obtained using quasi-static compressive tests were compared with those obtained

using elastic wave tests. Both followed the same qualitative pattern, however the Young’s

moduli measured using elastic waves were more than twice the values obtained by

mechanical tests. This discrepancy was explained by the combined effects of the strain rate

difference, strain level, yielding of EP particles during unloading, and buckling deformation

in both EP particles and the epoxy ligaments. Poisson’s ratio showed an increasing trend with

the foam density and appeared to be influenced by the increase in contact surface area

between the particles and the matrix as the foam density increased.

In summary, all of the objectives of the current research listed in Section 2.5, have been

addressed:

vi. EP particles were introduced as a cost-efficient porous filler to produce light-weight

syntactic foams.

vii. Structural, microstructural, physical, and mechanical properties of EP particles were

characterised.

viii. Three models suitable for predicting the elastic properties of packed beds of EP

particles were introduced. Two of which (i.e. Wang, and Gibson and Ashby) were

Page 210: Elastic And Mechanical Properties Of Expanded Perlite And ...

186

considered to give an average reasonable trend and the other one (i.e. Phani) with our

modified shape factor was introduced to give the best agreement with the data

generated in this study.

ix. An economical light-weight EP/epoxy foam was manufactured.

x. Microstructural, structural, and mechanical properties of EP/epoxy foams were

characterised using the optical and scanning electron microscopy, quasi-static

compressive tests and elastic wave speed tests.

xi. The effects of the EP particle sizes and foam density on mechanical properties and

behaviour of the foam was investigated.

xii. Damage mechanisms occurring during the tests and behaviour of EP/epoxy foams

under compressive loads were investigated.

Future Research

The mechanical properties (e.g. compressive strength and stiffness) of EP/epoxy foams will

be improved by improving the bonding between the EP particles and the epoxy binder. Since

inorganic fillers such as EP particles generally have a poor affinity with organic resins, they

cannot be chemically bonded and the bonding between the EP particles and the resin is

through the mechanical retention in their irregular pores. To improve the adhesion, the EP

particles could be initially coated with organic compounds and then embedded in the resin

matrix. Organic compounds which have three or more ethylenically unsaturated groups, such

as acrylates or methacrylates of polyhydric alcohols, have been found to be very effective in

bonding inorganic fillers to organic resins [150]. Since not all of the ethylenically unsaturated

Page 211: Elastic And Mechanical Properties Of Expanded Perlite And ...

187

groups in the organic compounds are consumed during the coating step of the EP particles,

the coated EP particles would be very reactive and could be chemically bonded to organic

resins. As a result of concurrent mechanical and chemical bonds between the EP particles

and the epoxy resin, significant improvement in the compressive strength, compressive

stiffness and bending strength of the EP/epoxy foams is expected. The mixing ratio of the

organic compounds to the EP particles varies in the range 20:80 - 80:20 and the optimum

value should be found by experimentation. To this end, the coated EP particles need to be

embedded in the matrix of epoxy. However, a different manufacturing method rather than

the buoyancy method used in the current study, is recommended. Due to problems associated

with the buoyancy method such as an increased curing time, and the breakage and crushing

of EP particles at the compression step, it would be well worth investigating other methods.

One promising method would be the filling of a mould completely with the closest particle

packing possible and then pouring a pre-measured amount of resin solution over the particles.

However, instead of diluting the epoxy with acetone which degrades the mechanical

properties and greatly increases the curing time (see Section 4.5), a low viscosity resin such

as CRACKBOND® SLV-302 can be used. This epoxy resin has a viscosity of 195 cP at 24°C

and a curing time of 7 days. This manufacturing method would have the advantage of a

quicker curing time and a significant reduction in the breakage and crushing of EP particles.

In addition, it would be easy to calculate the volume fraction of the epoxy and EP particles

which was a challenge in the current study. Such EP/epoxy foams should be fabricated with

different particle size ranges (e.g. 0.710 – 1 mm, 1 - 1.18 mm, 1.8 - 1.40 mm and 1.40 - 2.0

mm) to cover a density range of 0.15 - 0.45 g/cm3. The structural, microstructural, mechanical

Page 212: Elastic And Mechanical Properties Of Expanded Perlite And ...

188

and thermal properties of the foams would be investigated, and the optimal mixing ratio of

the organic compounds to EP particles, in terms of the resultant properties, would be

determined. Moreover, as one of the main applications of syntactic foams is to use as the core

material in sandwich plates and shells, additional work to attach a stiff skin to the top and

bottom of the core should be undertaken. The skin itself could be a composite of one of three

types: i) glass fibre/epoxy, ii) carbon fibre/epoxy, and iii) hybrid glass/carbon/epoxy. The

prepared panels would then be subjected to flexural and fatigue tests, and the results would

be analysed and the best sandwich panels determined.

Page 213: Elastic And Mechanical Properties Of Expanded Perlite And ...

189

References

1. Gibson, L.J., Modelling the mechanical behavior of cellular materials. Materials Science and

Engineering: A, 1989. 110: p. 1-36.

2. Lee, S.M., Pocessing and Preparation of Syntactic foams, in Handbook of composite

reinforcements. 1992, VCH Publishers and VCH vch verlagsgesellschaft mbH. p. 257.

3. Gong, L., Kyriakides, S., and Jang, W.Y., Compressive response of open-cell foams. Part I:

Morphology and elastic properties. International Journal of Solids and Structures, 2005.

42(5–6): p. 1355-1379.

4. Gupta, N., Woldesenbet, E., hore, K., and Sankaran, S., Response of Syntactic Foam Core

Sandwich Structured Composites to Three-Point Bending. Journal of Sandwich Structures &

Materials, 2002. 4(3): p. 249-272.

5. Gupta, N., Karthikeyan, C.S., Sankaran, S., and Kishore, Correlation of Processing

Methodology to the Physical and Mechanical Properties of Syntactic Foams With and

Without Fibers. Materials Characterization, 1999. 43(4): p. 271-277.

6. Bardella, L. and Genna, F., Elastic design of syntactic foamed sandwiches obtained by filling

of three-dimensional sandwich-fabric panels. International Journal of Solids and Structures,

2001. 38(2): p. 307-333.

7. Peroni, L., Scapin, M., Avalle, M., Weise, J., and Lehmhus, D., Dynamic mechanical

behavior of syntactic iron foams with glass microspheres. Materials Science and

Engineering: A, 2012. 552: p. 364-375.

8. Shutov, F.A., Syntactic polymer foams, in Chromatography/Foams/Copolymers. 1986,

Springer Berlin Heidelberg: Berlin, Heidelberg. p. 63-123.

9. Ghosh, D., Wiest, A., and Conner, R.D., Uniaxial quasistatic and dynamic compressive

response of foams made from hollow glass microspheres. Journal of the European Ceramic

Society, 2016. 36(3): p. 781-789.

10. Alkan, M. and Doğan, M., Adsorption of Copper(II) onto Perlite. Journal of Colloid and

Interface Science, 2001. 243(2): p. 280-291.

11. Taherishargh, M., Belova, I.V., Murch, G.E., and Fiedler, T., Low-density expanded perlite–

aluminium syntactic foam. Materials Science and Engineering: A, 2014. 604(0): p. 127-134.

12. Ciullo, P.A., Industrial minerals and their uses: a handbook and formulary. 1996, Westwood,

N.J. : Noyes Publications.

13. Broxtermann, S., Taherishargh, M., Belova, I.V., Murch, G.E., and Fiedler, T., On the

compressive behaviour of high porosity expanded Perlite-Metal Syntactic Foam (P-MSF).

Journal of Alloys and Compounds, 2017. 691: p. 690-697.

14. Allameh-Haery, H., Kisi, E., and Fiedler, T., Novel cellular perlite–epoxy foams: Effect of

density on mechanical properties. Journal of Cellular Plastics, 2016.

15. Allameh-Haery, H., Wensrich, C.M., Fiedler, T., and Kisi, E., Novel Cellular perlite-epoxy

foams: effects of particle size. Journal of Cellular Plastics, 2016.

16. Allameh-Haery, H., Kisi, E., Pineda, J., Suwal, L.P., and Fiedler, T., Elastic properties of

green expanded perlite particle compacts. Powder Technology, 2017. 310: p. 329-342.

17. Moavenzadeh, F., ed. Concise Encyclopedia of Building and Construction Materials.

Advances in Materials Science and Engineering. 1990, MIT Press: Cambridge, MA. 698.

18. Bush, A.L., Lightweight Aggregate, in United states mineral resources, D.A. Brobst and

W.P. Pratt, Editors. 1973, United States government printing office: Washington, WA. p.

333-355.

Page 214: Elastic And Mechanical Properties Of Expanded Perlite And ...

190

19. Otis, L.M., Perlite, in Mineral facts and problems, United States Bureau of Mines, Editor.

1960, United states Government Printing Office: Washington, WA. p. 581-587.

20. Breese, R.O.Y. and Barker, J.M. Perlite. in Industrial Minerals and Rocks. 1994. Society for

Mining, Metallurgy and Exploration, Littleton, CO, United States.

21. Bates, R.L., Geology of the Industrial Rocks and Minerals. 1960, New York, NY: Harper &

Brothers.

22. Doğan, M. and Alkan, M., Removal of methyl violet from aqueous solution by perlite. Journal

of Colloid and Interface Science, 2003. 267(1): p. 32-41.

23. Topçu, İ.B. and Işıkdağ, B., Manufacture of high heat conductivity resistant clay bricks

containing perlite. Building and Environment, 2007. 42(10): p. 3540-3546.

24. Sarı, A., Tuzen, M., Cıtak, D., and Soylak, M., Adsorption characteristics of Cu(II) and

Pb(II) onto expanded perlite from aqueous solution. Journal of Hazardous Materials, 2007.

148(1–2): p. 387-394.

25. Alkan, M., Karadaş, M., Doğan, M., and Demirbaş, Ö., Adsorption of CTAB onto perlite

samples from aqueous solutions. Journal of Colloid and Interface Science, 2005. 291(2): p.

309-318.

26. Singh, M. and Garg, M., Perlite-based building materials — a review of current applications.

Construction and Building Materials, 1991. 5(2): p. 75-81.

27. Barker, J.M. and Santini, K., Industrial Minerals and Rocks - Commodities, Markets, and

Uses J.E. Kogel, N.C. Trivedi, J.M. Barker, and S.T. Krukowski, Editors. 2006, Society for

Mining, Metallurgy, and Exploration (SME). p. 685-702.

28. Johnstone, S.J. and Johnstone, M.G., Minerals for the chemical and allied industries 2nd ed.

1961, London, UK: Chapman and Hall

29. Gunning , D.F. and Ltd., M.A., Perlite Market Study for British Columbia. 1994. p. 100.

30. Jackson, F.L., Processing perlite for use in insulation applications. Transactions of the

American Institute of Mining, Metallurgical, and Petroleum Engineers, Society, 1986. 280(pt

A): p. 40-45.

31. Saunders, E., Agglomerating particulate perlite. 1979: US.

32. Kadey, F.L., Jr. , Perlite, in Industrial Minerals and Rocks, S.J. Lefond, Editor. 1983,

American Institute of Mining, Metallurgical, and Petroleum Engineers. p. 997-1015.

33. King, E.G., Todd, S.S., and Kelley, K.K., Perlite: Thermal data and energy required for

expansion. 1948, United States Bureau of Mines: Washington, DC, United States. p. 15.

34. Pettifer, L., Perlite; diversification the key to overall expansion. Industrial Minerals

(London), 1981. 171: p. 55-75.

35. Wang, B., Smith, T.R.U.S., Masters, A.L.U.S., and Danvers, N.J.K., Micronized perlite filler

product, C. Advanced Minerals, Editor. 2009: US.

36. AUSTIN, G.S. and BARKER, J.M., Commercial perlite deposits of New Mexico and North

America. Mack, GH. Austin, GS, & Barker, JM (eds) New Mexico Geological Society,

Guidebook, 1998. 49: p. 271-277.

37. Roulia, M., Chassapis, K., Kapoutsis, J.A., Kamitsos, E.I., and Savvidis, T., Influence of

thermal treatment on the water release and the glassy structure of perlite. Journal of

Materials Science, 2006. 41(18): p. 5870-5881.

38. Doğan, M. and Alkan, M., Some physicochemical properties of perlite as an adsorbent.

Fresenius Environmental Bulletin, 2004. 13(36): p. 251-257.

39. Cheremisinoff, N.P., Handbook of Water and Wastewater Treatment Technologies. Elsevier.

106-120.

40. Afshari, H., Ashraf, S., Ghaffar Ebadi, A., Jalali, S., Abbaspour, H., Sam Daliri, M., Toudar,

S.R., and Study of the effects irrigation water salinity and pH on production and relative

Page 215: Elastic And Mechanical Properties Of Expanded Perlite And ...

191

absorption of some elements nutrient by the tomato plant. American Journal of Applied

Sciences, 2011. 8(8): p. 766-772.

41. Burriesci, N., Arcoraci, C., Antonucci, P., and Polizzotti, G., Physico-chemical

characterization of perlite of various origins. Materials Letters, 1985. 3(3): p. 103-110.

42. Sersale, R., Burriesci, N., Pino, L., and Bart, J.C.J., Iron distribution in some Italian tuffs.

Materials Letters, 1984. 3(1–2): p. 51-57.

43. Koukouzas, N., Rare earth elements in volcanic glass: A case study from Trachilas perlite

deposit, Greece. Chemie Der Erde, 1997. 57: p. 351-362.

44. Yanev, Y. and Yordanov, Y., REE, Th, U, Hf, Sc, Co and Ta in the cesium-bearing perlites,

adularites and zeolitites from the Eastern Rhodopes. Comptes rendus de l'Académie bulgare

des Sciences, 1991. 44(4): p. 71-74.

45. Zielinski, R., Lipman, P., and Millard, H., Minor-element abundances in obsidian, perlite,

and felsite of calc-alkalic rhyolites. American Mineralogist, 1977. 62(5-6): p. 426-437.

46. Stolper, E., Water in silicate glasses: an infrared spectroscopic study. Contributions to

Mineralogy and Petrology, 1982. 81(1): p. 1-17.

47. Saisuttichai, D. and Manning, D.A., Geochemical characteristics and expansion properties

of a highly potassic perlitic rhyolite from Lopburi, Thailand. Resource geology, 2007. 57(3):

p. 301-312.

48. Herskovitch, D. and Lin, I.J., Upgrading of raw perlite by a dry magnetic technique.

Magnetic and Electrical Separation, 1996. 7(3): p. 145-161.

49. Scott, P.W. and Bristow, C.M., eds. Industrial Minerals and Extractive Industry Geology.

2002, The Geological Society Bath, Uk.

50. Barth-Wirsching, U., Höller, H., Klammer, D., and Konrad, B., Synthetic zeolites formed

from expanded perlite: Type, formation conditions and properties. Mineralogy and

Petrology, 1993. 48(2-4): p. 275-294.

51. Chakir, A., Bessiere, J., Kacemi, K.E.L., and Marouf, B.b., A comparative study of the

removal of trivalent chromium from aqueous solutions by bentonite and expanded perlite.

Journal of Hazardous Materials, 2002. 95(1–2): p. 29-46.

52. Fornasiero, R.B., Phytotoxic effects of fluorides. Plant Science, 2001. 161(5): p. 979-985.

53. Kendall, T., No sign of the bubble bursting- Perlite uses & Markets. Industrial Minerals,

2000(393): p. 51-53.

54. Johnstone, S.J. and Johnstone, M.G., Minerals for the chemical and allied industries. 1961:

Chapman and Hall.

55. Sengul, O., Tasdemir, C., and Tasdemir, M.A., Influence of aggregate type on mechanical

behavior of normal- and high-strength concretes. ACI Materials Journal, 2002. 99(6): p. 528-

533.

56. Topçu, İ.B. and Işıkdağ, B., Effect of expanded perlite aggregate on the properties of

lightweight concrete. Journal of Materials Processing Technology, 2008. 204(1–3): p. 34-38.

57. ASTM, Standard Specification for Lightweight Aggregates for Structural Concrete. 2014,

ASTM International: West Conshohocken.

58. American Standards for Testing Materials, Standard specification for lightweight aggregates

for insulating concrete. 2009, ASTM International: West Conshohocken, PA.

59. Pevzner, E. and Gontmakher, V., Method for manufacture of foamed perlite material. 2002:

US.

60. Lura, P., Bentz, D.P., Lange, D.A., Kovler, K., and Bentur, A. Pumice aggregates for internal

water curing. in International RILEM Symposium on Concrete Science and Engineering: A

Tribute to Arnon Bentur. 2004. RILEM Publications SARL.

Page 216: Elastic And Mechanical Properties Of Expanded Perlite And ...

192

61. Yu, L.-H., Ou, H., and Lee, L.-L., Investigation on pozzolanic effect of perlite powder in

concrete. Cement and Concrete Research, 2003. 33(1): p. 73-76.

62. Sengul, O., Azizi, S., Karaosmanoglu, F., and Tasdemir, M.A., Effect of expanded perlite on

the mechanical properties and thermal conductivity of lightweight concrete. Energy and

Buildings, 2011. 43(2): p. 671-676.

63. Singh, M. and Garg, M., Perlite-based building materials—a review of current applications.

Construction and Building Materials, 1991. 5(2): p. 75-81.

64. Yilmazer, S. and Ozdeniz, M.B., The effect of moisture content on sound absorption of

expanded perlite plates. Building and Environment, 2005. 40(3): p. 311-318.

65. Berge, B., The ecology of building materials. 2nd ed. 2009, Oxford, UK: Elsevier’s Science

& Technology.

66. Englert, M.H.U.S., Acoustical tile containing treated perlite, I. Usg Interiors, Editor. 1999:

US.

67. Nelson, C.R.U.S., Expanded perlite annealing process, C. United States Gypsum, Editor.

2008: US.

68. Baig, M.A.U.S., Acoustic ceiling tiles made with paper processing waste, L.L.C. Usg

Interiors, Editor. 2012: US.

69. Denning Paul, S., Insulating product and its manufacture. 1962: US.

70. Keskey, W., Camisa, J.D., and Meath, K.R., Composite board, containing certain latex

binder compositions, C. The Dow Chemical, Editor. 1992: EP.

71. DePorter, C.D., Dawson, S.D., Battaglioli, M.V., and Sandoval, C.P., Perlite-based

insulation board, I. Johns Manville International, Editor. 2000: US.

72. Izard, D.G.U.S. and Englert, M.H.U.S., Method for manufacturing a mineral wool panel, I.

Usg Interiors, Editor. 1993: US.

73. Fernando, J.A.U.S. and Rioux, R., Ultra low weight insulation board, I.L. Unifrax, Editor.

2013: US.

74. Greve, D.R.U.S. and Richards, T.W.U.S., Fire door core, C. Georgia-Pacific, Editor. 1979:

US.

75. Mukherjee, S., Applied Mineralogy: Applications in Industry and Environment

2011, New York, NY: Springer

76. Dickson, T.W., Perlite. The mineral industry of New South Wales. Vol. 31. 1968, Sydney,

Australia: Dept. of Mines, Geological Survey of New South Wales

77. Beikircher, T. and Demharter, M., Heat Transport in Evacuated Perlite Powders for Super-

Insulated Long-Term Storages up to 300 °C. Journal of Heat Transfer, 2013. 135(5): p.

051301-051301.

78. Kropschot, R.H. and Burgess, R.W. Perlite for cryogenic insulation. in Proceedings of the

Cryogenic Engineering Conference, 1962. 1962.

79. Demharter, M., Heat transport in evacuated perlite powder insulations and Its application in

long-term

hot water storages, in Physics. 2011, Technische Universität München.

80. Fulk, M.M., Evacuated powder insulation for low temperatures, in Progress in Cryogenics,

K. Mendelssohn, Editor. 1959, Heywood &Company: London, UK. p. 65-84.

81. Demharter, M., Heat Transport in Evacuated Perlite Powder Insulations and Its Application

in Long-Term Hot Water Storages. Master's Thesis, Technische Universität München,

Faculty of Physics, 2011.

82. Black Igor, A. and Fowle Arthur, A., Insulating assembly, U.S. Patent, Editor. 1964.

83. Dubé, W.P., Sparks, L.L., and Slifka, A.J., Thermal conductivity of evacuated perlite at low

temperatures as a function of load and load history. Cryogenics, 1991. 31(1): p. 3-6.

Page 217: Elastic And Mechanical Properties Of Expanded Perlite And ...

193

84. Elias, H.G., Plastics, general survey. Ullmann's Encyclopedia of Industrial Chemistry, 2000.

85. Dedeloudis, C. and Karalis, T., Milled expanded volvanic glass as lamellar filler, B.I.M. S.A,

Editor. 2012: EP.

86. Wang, B., Roulston, J.S.U.S., Palm, S.K.U.S., and Hayward, C., Perlite products with

controlled particle size distribution, C. Advanced Minerals, Editor. 2002: US.

87. Agioutantis, Z., Agioutantis, Z.G., and Komnitsas, K. Book of proceedings: advances in

mineral resources management and environmental geotechnology. 2006. Heliotopos

Conferences.

88. Kongkachuichay, P. and Lohsoontorn, P., Phase diagram of zeolite synthesized from perlite

and rice husk ash. SCIENCEASIA, 2006. 32: p. 13-16.

89. Christidis, G.E., Paspaliaris, I., and Kontopoulos, A., Zeolitisation of perlite fines:

mineralogical characteristics of the end products and mobilization of chemical elements.

Applied Clay Science, 1999. 15(3–4): p. 305-324.

90. Ottanà, R., Saija, L.M., Burriesci, N., and Giordano, N., Hydrothermal synthesis of zeolites

from pumice in alkaline and saline environment. Zeolites, 1982. 2(4): p. 295-298.

91. Antonucci, P.L., Crisafulli, M.L., Giordano, N., and Burriesci, N., Zeolitization of perlite.

Materials Letters, 1985. 3(7–8): p. 302-307.

92. Wang, P., Shen, B., Shen, D., Peng, T., and Gao, J., Synthesis of ZSM-5 zeolite from expanded

perlite/kaolin and its catalytic performance for FCC naphtha aromatization. Catalysis

Communications, 2007. 8(10): p. 1452-1456.

93. Khodabandeh, S. and Davis, M.E., Alteration of perlite to calcium zeolites. Microporous

Materials, 1997. 9(3–4): p. 161-172.

94. Steenbruggen, G. and Hollman, G.G., The synthesis of zeolites from fly ash and the properties

of the zeolite products. Journal of Geochemical Exploration, 1998. 62(1–3): p. 305-309.

95. Christidis, G.E. and Papantoni, H., Synthesis of FAU type zeolite Y from natural raw

materials: hydrothermal SiO2-Sinter and Perlite glass. The Open Mineralogy Journal, 2008.

2: p. 1-5.

96. Rosa, M.E., An introduction to solid foams. Philosophical Magazine Letters, 2008. 88(9-10):

p. 637-645.

97. Gibson, L.J. and Ashby, M.F., Cellular solids: structure and properties. 1999: Cambridge

university press.

98. Smits, G.F., Effect of Cellsize Reduction on Polyurethane Foam Physical Properties. Journal

of Thermal Insulation and Building Envelopes, 1994. 17(4): p. 309-329.

99. Rosa, M.E., Pereira, H., and Fortes, M., Effects of hot water treatment on the structure and

properties of cork. Wood and Fiber Science, 1990. 22(2): p. 149-164.

100. Pereira, H., Rosa, M.E., and Fortes, M., The cellular structure of cork from Quercus suber L.

IAWA Journal, 1987. 8(3): p. 213-218.

101. Fortes, M.A., Rosa, M.E., and Pereira, H., A cortiça. 2004: IST Press Lisboa.

102. Cateto, C.A., Barreiro, M.F., Ottati, C., Lopretti, M., Rodrigues, A.E., and Belgacem, M.N.,

Lignin-based rigid polyurethane foams with improved biodegradation. Journal of Cellular

Plastics, 2013. 50(1): p. 81-95.

103. Khemani, K.C., Polymeric foams: an overview, in Polymeric Foams: Science and

Technology, K.C. Khemani, Editor. 1997, American Chemical Society: Washington, DC.

104. Martelli, F., in Twin Screw Extruders. 1983, Van Nostrand Reinhold Co. : New York.

105. Lee, S.-T., Park, C.B., and Ramesh, N.S., Polymeric foams: science and technology. 2006:

CRC Press.

106. Fatt, M.S.H. and Chen, L., A viscoelastic damage model for hysteresis in PVC H100 foam

under cyclic loading. Journal of Cellular Plastics, 2015. 51(3): p. 269-287.

Page 218: Elastic And Mechanical Properties Of Expanded Perlite And ...

194

107. Leidner, J. and Woodhams, R.T., The strength of polymeric composites containing spherical

fillers. Journal of Applied Polymer Science, 1974. 18(6): p. 1639-1654.

108. Nicolais, L. and Mashelkar, R.A., The strength of polymeric composites containing spherical

fillers. Journal of Applied Polymer Science, 1976. 20(2): p. 561-563.

109. Spanoudakis, J. and Young, R.J., Crack propagation in a glass particle-filled epoxy resin.

Journal of Materials Science, 1984. 19(2): p. 473-486.

110. Alonso, M.V., Auad, M.L., and Nutt, S., Short-fiber-reinforced epoxy foams. Composites

Part A: Applied Science and Manufacturing, 2006. 37(11): p. 1952-1960.

111. Gupta, N. and Nagorny, R., Tensile properties of glass microballoon-epoxy resin syntactic

foams. Journal of Applied Polymer Science, 2006. 102(2): p. 1254-1261.

112. Gupta, N., Zeltmann, S.E., Shunmugasamy, V.C., and Pinisetty, D., Applications of Polymer

Matrix Syntactic Foams. JOM, 2014. 66(2): p. 245-254.

113. Puterman, M., Narkis, M., and Kenig, S., Syntactic Foams I. Preparation, Structure and

Properties. Journal of Cellular Plastics, 1980. 16(4): p. 223-229.

114. Okuno, K. and Woodhams, R.T., Mechanical Properties and Characterization of Phenolic

Resin Syntactic Foams. Journal of Cellular Plastics, 1974. 10(5): p. 237-244.

115. Luoma, E.J. and Watkins, L., Syntactic Foams, in Modern Plastics Encyclopedia. 1979,

McGraw-Hill: New York. p. 273.

116. Lubin, G., Handbook of fiberglass and advanced plastics composites. 1975: RE Krieger

Publishing Company.

117. Lee, S.M., Handbook of composite reinforcements. 1992: John Wiley & Sons.

118. Suh, K., Klempner, D., and Frisch, K., Handbook of Polymeric Foams and Foam Technology.

Hanser, Munich, 2004: p. 189-232.

119. Landrock, A.H., 6. Solvent Cementing and Adhesive Bonding of Foams, in Handbook of

Plastic Foams. William Andrew Publishing/Noyes.

120. Yung, K.C., Zhu, B.L., Yue, T.M., and Xie, C.S., Preparation and properties of hollow glass

microsphere-filled epoxy-matrix composites. Composites Science and Technology, 2009.

69(2): p. 260-264.

121. Gupta, N., Woldesenbet, E., and Mensah, P., Compression properties of syntactic foams:

effect of cenosphere radius ratio and specimen aspect ratio. Composites Part A: Applied

Science and Manufacturing, 2004. 35(1): p. 103-111.

122. Kim, H.S. and Plubrai, P., Manufacturing and failure mechanisms of syntactic foam under

compression. Composites Part A: Applied Science and Manufacturing, 2004. 35(9): p. 1009-

1015.

123. Gupta, N. and Maharsia, R., Enhancement of Energy Absorption in Syntactic Foams by

Nanoclay Incorporation for Sandwich Core Applications. Applied Composite Materials,

2005. 12(3): p. 247-261.

124. Gupta, N., Kishore, Woldesenbet, E., and Sankaran, S., Studies on compressive failure

features in syntactic foam material. Journal of Materials Science, 2001. 36(18): p. 4485-4491.

125. Samsudin, S., Ariff, Z., Zakaria, Z., and Bakar, A., Development and characterization of

epoxy syntactic foam filled with epoxy hollow spheres. Exp. Pol. Lett, 2011. 7: p. 653-660.

126. Mudge, R.S., Process for making a low density syntactic foam product and the resultant

product. 1988, US Patent.

127. Matthews, A.M., Impact resistant thermoplastic syntactic foam composite and method. 1994,

US Patent.

128. Dodiuk, H. and Goodman, S.H., Handbook of thermoset plastics. 2013: William Andrew.

129. Du Pont, P.S., Freeman, J.E., Ritter, R.E., and Wittmann, A., Fiber-reinforced syntactic foam

composites and method of forming same. 1986, US Patent.

Page 219: Elastic And Mechanical Properties Of Expanded Perlite And ...

195

130. Harrison, E.S., Bridges, D.J., and Melquist, J.L., Thermosetting syntactic foams and their

preparation. 2000, US Patent.

131. Kishore, Shankar, R., and Sankaran, S., Gradient syntactic foams: Tensile strength, modulus

and fractographic features. Materials Science and Engineering: A, 2005. 412(1–2): p. 153-

158.

132. Narkis, M., Puterman, M., and Kenig, S., Syntactic Foams II. Preparation and

Characterization Of Three-Phase Systems. Journal of Cellular Plastics, 1980. 16(6): p. 326-

330.

133. Ashton-Patton, M.M., Hall, M.M., and Shelby, J.E., Formation of low density

polyethylene/hollow glass microspheres composites. Journal of Non-Crystalline Solids,

2006. 352(6–7): p. 615-619.

134. Patankar, S.N., Das, A., and Kranov, Y.A., Interface engineering via compatibilization in

HDPE composite reinforced with sodium borosilicate hollow glass microspheres.

Composites Part A: Applied Science and Manufacturing, 2009. 40(6–7): p. 897-903.

135. Patankar, S.N. and Kranov, Y.A., Hollow glass microsphere HDPE composites for low

energy sustainability. Materials Science and Engineering: A, 2010. 527(6): p. 1361-1366.

136. Cravens, T.E., Syntactic Foams Utilizing Saran Microspheres. Journal of Cellular Plastics,

1973. 9(6): p. 260-267.

137. Narkis, M., Gerchcovich, M., Puterman, M., and Kenig, S., Syntactic Foams III. Three-Phase

Materials Produced from Resin Coated Microballoons. Journal of Cellular Plastics, 1982.

18(4): p. 230-232.

138. Seamark, M.J., Innovative Use of Syntactic Foam in GRP Sandwich Construction for Wind

Generator Nacelles, in Composite Structures 4: Volume 1 Analysis and Design Studies, I.H.

Marshall, Editor. 1987, Springer Netherlands: Dordrecht. p. 333-341.

139. Kim, H.S., Syntactic foam. 2003,

Patents: Australia.

140. Hossain, K.M.A., Development of volcanic pumice based cement and lightweight concrete.

Magazine of Concrete Research, 2004. 56(2): p. 99-109.

141. Taherishargh, M., Belova, I.V., Murch, G.E., and Fiedler, T., Pumice/aluminium syntactic

foam. Materials Science and Engineering: A, 2015. 635: p. 102-108.

142. Fleischer, C.A. and Zupan, M., Mechanical Performance of Pumice-reinforced Epoxy

Composites. Journal of Composite Materials, 2010. 44(23): p. 2679-2696.

143. Sahin, A., Yildiran, Y., Avcu, E., Fidan, S., and Sinmazcelik, T., Mechanical and Thermal

Properties of Pumice Powder Filled PPS Composites. Acta Physica Polonica A, 2014.

125(2): p. 518-520.

144. Sahin, A., Karsli, N.G., and Sinmazcelik, T., Comparison of the mechanical,

thermomechanical, thermal, and morphological properties of pumice and calcium

carbonate-filled poly(phenylene sulfide) composites. Polymer Composites, 2016. 37(11): p.

3160-3166.

145. Ramesan, M.T., George, A., Jayakrishnan, P., and Kalaprasad, G., Role of pumice particles

in the thermal, electrical and mechanical properties of poly(vinyl alcohol)/poly(vinyl

pyrrolidone) composites. Journal of Thermal Analysis and Calorimetry, 2016. 126(2): p. 511-

519.

146. Alvarado, A., Morales, K., Srubar, W., and Billington, S., Effects of natural porous additives

on the tensile mechanical performance and moisture absorption behavior of PHBV-based

composites for construction. Stanford Undergraduate Research Journal, 2011. 10.

147. Srubar, W.V. and Billington, S.L., PHBV/ground bone meal and pumice powder engineered

biobased composite materials for construction. 2013, Google Patents.

Page 220: Elastic And Mechanical Properties Of Expanded Perlite And ...

196

148. Han, B., Sun, Z., Chen, Y., Tian, F., Wang, X., and Lei, Q. Space charge distribution in Low-

density Polyethylene (LDPE)/Pumice composite. in 2009 IEEE 9th International Conference

on the Properties and Applications of Dielectric Materials. 2009.

149. Ramesan, M.T., Jose, C., Jayakrishnan, P., and Anilkumar, T., Multifunctional ternary

composites of poly (vinyl alcohol)/cashew tree gum/pumice particles. Polymer Composites,

2016: p. n/a-n/a.

150. Masuhara, E., Nakabayashi, N., Nagata, K., and Takeyama, M., Composite filler and dental

composition containing the same. 1982, US Patents.

151. Ahmetli, G., Dag, M., Deveci, H., and Kurbanli, R., Recycling studies of marble processing

waste: Composites based on commercial epoxy resin. Journal of Applied Polymer Science,

2012. 125(1): p. 24-30.

152. Rashad, A.M., Vermiculite as a construction material – A short guide for Civil Engineer.

Construction and Building Materials, 2016. 125: p. 53-62.

153. Straaten, H.P.V., Rocks for Crops: Agrominerals of Sub-Saharan Africa. 2002, Nairobi,

Kenya: ICRAF 338.

154. Kogel, J.E., Trivedi, N.C., Barker, J.M., and Krukowski, S.T., Industrial Minerals and Rocks

- Commodities, Markets, and Uses (7th Edition). Society for Mining, Metallurgy, and

Exploration (SME).

155. JUN, S., XIAOLIN, F., YONG, R., BOWEN, C., and XINYU, R., Preparation method of

phenolic resin/ expanded vermiculite composite flame retardant heat preservation material.

2016, Chinese Patents.

156. Verbeek, C.J.R. and du Plessis, B.J.G.W., Density and flexural strength of phosphogypsum–

polymer composites. Construction and Building Materials, 2005. 19(4): p. 265-274.

157. Yu, J., He, J., and Ya, C., Preparation of phenolic resin/organized expanded vermiculite

nanocomposite and its application in brake pad. Journal of Applied Polymer Science, 2011.

119(1): p. 275-281.

158. Zeng, X., Cai, D., Lin, Z., Cai, X., Zhang, X., Tan, S., and Xu, Y., Morphology and thermal

and mechanical properties of phosphonium vermiculite filled poly(ethylene terephthalate)

composites. Journal of Applied Polymer Science, 2012. 126(2): p. 601-607.

159. Patro, T.U., Harikrishnan, G., Misra, A., and Khakhar, D.V., Formation and characterization

of polyurethane—vermiculite clay nanocomposite foams. Polymer Engineering & Science,

2008. 48(9): p. 1778-1784.

160. Qian, Y., Lindsay, C.I., Macosko, C., and Stein, A., Synthesis and Properties of Vermiculite-

Reinforced Polyurethane Nanocomposites. ACS Applied Materials & Interfaces, 2011. 3(9):

p. 3709-3717.

161. Xu, J., Meng, Y.Z., Li, R.K.Y., Xu, Y., and Rajulu, A.V., Preparation and properties of

poly(vinyl alcohol)–vermiculite nanocomposites. Journal of Polymer Science Part B:

Polymer Physics, 2003. 41(7): p. 749-755.

162. Tjong, S.C., Meng, Y.Z., and Hay, A.S., Novel Preparation and Properties of

Polypropylene−Vermiculite Nanocomposites. Chemistry of Materials, 2002. 14(1): p. 44-51.

163. Mittal, V., Epoxy—Vermiculite Nanocomposites as Gas Permeation Barrier. Journal of

Composite Materials, 2008. 42(26): p. 2829-2839.

164. Zhang, Y., Han, W., and Wu, C.-F., Preparation and Properties of Polypropylene/Organo-

Vermiculite Nanocomposites. Journal of Macromolecular Science, Part B, 2009. 48(5): p.

967-978.

165. Harikrishnan, G., Lindsay, C.I., Arunagirinathan, M.A., and Macosko, C.W., Probing

Nanodispersions of Clays for Reactive Foaming. ACS Applied Materials & Interfaces, 2009.

1(9): p. 1913-1918.

Page 221: Elastic And Mechanical Properties Of Expanded Perlite And ...

197

166. Li, X., Lei, B., Lin, Z., Huang, L., Tan, S., and Cai, X., The utilization of organic vermiculite

to reinforce wood–plastic composites with higher flexural and tensile properties. Industrial

Crops and Products, 2013. 51: p. 310-316.

167. Avcu, E., Çoban, O., Bora, M.Ö., Fidan, S., Sınmazçelik, T., and Ersoy, O., Possible use of

volcanic ash as a filler in polyphenylene sulfide composites: Thermal, mechanical, and

erosive wear properties. Polymer Composites, 2014. 35(9): p. 1826-1833.

168. Bora, M.Ö., Çoban, O., Fİdan, S., Kutluk, T., and Sinmazçelİk, T., Surface Modification

Effect of Volcanic Ash Particles Using Silane Coupling Agent on Mechanical Properties of

Polyphenylene Sulfide Composites. Acta Physica Polonica A, 2016. 129(4): p. 495-497.

169. Bora, M.Ö., Çoban, O., Kutluk, T., Fİdan, S., and Sinmazçelİk, T., The influence of heat

treatment process on mechanical properties of surface treated volcanic ash

particles/polyphenylene sulfide composites. Polymer Composites, 2016.

170. Trinidad, S., Luis, J., Vite, J., Franco, A., Mendoza Nuñez, M., and Gutiérrez Torres, C.d.C.

Mechanical behavior of ceramic and polymer composites reinforced with volcanic ashes. in

Key Engineering Materials. 2010. Trans Tech Publ.

171. Fidan, S., Volcanic Ash Reinforcement Concentration Effect on Thermal Properties of

Polyvinyl Chloride Composites. Acta Physica Polonica A, 2015. 127(4): p. 1002-1003.

172. Bora, M.Ö., Evaluation of Volcanic Ash Concentration Effect on Mechanical Properties of

Poly(Vinyl Chloride) Composites. Acta Physica Polonica A, 2015. 127(4): p. 1004-1006.

173. Avcu, E., Çoban, O., Özgür Bora, M., Fidan, S., Sınmazçelik, T., and Ersoy, O., Possible use

of volcanic ash as a filler in polyphenylene sulfide composites: Thermal, mechanical, and

erosive wear properties. Polymer Composites, 2014. 35(9): p. 1826-1833.

174. Cernohous, J.J., Compositions and methods for producing high strength composites. 2012,

Google Patents.

175. Çoban, O., Özgür Bora, M., Kutluk, T., Fidan, S., and Sinmazçelik, T., Heat treatment effect

on thermal and thermomechanical properties of polyphenylene sulfide composites reinforced

with silane-treated volcanic ash particles. Polymer Composites, 2016: p. n/a-n/a.

176. Çoban, O., Bora, M.Ö., Kutluk, T., Fİdan, S., and Sinmazçelİk, T., Effect of Silane as

Coupling Agent on Dynamic Mechanical Properties of Volcanic Ash Filled PPS Composites.

Acta Physica Polonica A, 2016. 129(4): p. 492-494.

177. Vite-Torres, M., Vite, J., Laguna-Camacho, J.R., Castillo, M., and Marquina-Chávez, A.,

Abrasive wear on ceramic materials obtained from solid residuals coming from mines. Wear,

2011. 271(9–10): p. 1231-1236.

178. Vite-Torres, J., Vite-Torres, M., Laguna-Camacho, J.R., Escalante-Martínez, J.E., Gallardo-

Hernández, E.A., and Vera-Cardenas, E.E., Wet abrasive behavior of composite materials

obtained from solid residuals mixed with polymer and ceramic matrix. Ceramics

International, 2014. 40(7, Part A): p. 9345-9353.

179. Baker, C.H. and Smail, V., Polyurethane composite matrix material and composite thereof.

2010, Google Patents.

180. Clay, E.L. and Baker, J.L., Natural sandwich of filled polyurethane foam. 1979, Google

Patents.

181. Lukosiute, I., Levinskas, R., Sapragonas, J., and Kviklys, A., Transition Layer in Composites

with a Plasticized Filler and Its Influence on the Strength Properties. Mechanics of

Composite Materials, 2004. 40(2): p. 151-158.

182. Sherman, N. and Cameron, J.H., Method of manufacturing improved mineral board. 1981.

183. Lu, Z., Xu, B., Zhang, J., Zhu, Y., Sun, G., and Li, Z., Preparation and characterization of

expanded perlite/paraffin composite as form-stable phase change material. Solar Energy,

2014. 108: p. 460-466.

Page 222: Elastic And Mechanical Properties Of Expanded Perlite And ...

198

184. Lu, Z., Hou, D., Xu, B., and Li, Z., Preparation and characterization of an expanded

perlite/paraffin/graphene oxide composite with enhanced thermal conductivity and leakage-

bearing properties. RSC Advances, 2015. 5(130): p. 107514-107521.

185. Karaıpeklı, A., Sarı, A., and Kaygusuz, K., Thermal Characteristics of Paraffin/Expanded

Perlite Composite for Latent Heat Thermal Energy Storage. Energy Sources, Part A:

Recovery, Utilization, and Environmental Effects, 2009. 31(10): p. 814-823.

186. Shastri, D. and Kim, H.S., A new consolidation process for expanded perlite particles.

Construction and Building Materials, 2014. 60: p. 1-7.

187. Kim, H., Syntactic foam. 2005, Google Patents.

188. Arifuzzaman, M. and Kim, H.S., Novel mechanical behaviour of perlite/sodium silicate

composites. Construction and Building Materials, 2015. 93: p. 230-240.

189. Taherishargh, M., Sulong, M.A., Belova, I.V., Murch, G.E., and Fiedler, T., On the particle

size effect in expanded perlite aluminium syntactic foam. Materials & Design (1980-2015),

2015. 66, Part A: p. 294-303.

190. Taherishargh, M., Belova, I.V., Murch, G.E., and Fiedler, T., The effect of particle shape on

mechanical properties of perlite/metal syntactic foam. Journal of Alloys and Compounds,

2017. 693: p. 55-60.

191. Jiang, B., Wang, Z., and Zhao, N., Effect of pore size and relative density on the mechanical

properties of open cell aluminum foams. Scripta Materialia, 2007. 56(2): p. 169-172.

192. Ramamurty, U. and Paul, A., Variability in mechanical properties of a metal foam. Acta

Materialia, 2004. 52(4): p. 869-876.

193. Wen, C.E., Yamada, Y., Shimojima, K., Chino, Y., Hosokawa, H., and Mabuchi, M.,

Compressibility of porous magnesium foam: dependency on porosity and pore size. Materials

Letters, 2004. 58(3–4): p. 357-360.

194. Kolluri, M., Karthikeyan, S., and Ramamurty, U., Effect of Lateral Constraint on the

Mechanical Properties of a Closed-Cell Al Foam: I. Experiments. Metallurgical and

Materials Transactions A, 2007. 38(9): p. 2006-2013.

195. Luong, D.D., Strbik Iii, O.M., Hammond, V.H., Gupta, N., and Cho, K., Development of high

performance lightweight aluminum alloy/SiC hollow sphere syntactic foams and compressive

characterization at quasi-static and high strain rates. Journal of Alloys and Compounds,

2013. 550: p. 412-422.

196. Li, K., Gao, X.L., and Subhash, G., Effects of cell shape and cell wall thickness variations on

the elastic properties of two-dimensional cellular solids. International Journal of Solids and

Structures, 2005. 42(5–6): p. 1777-1795.

197. Sugimura, Y., Meyer, J., He, M.Y., Bart-Smith, H., Grenstedt, J., and Evans, A.G., On the

mechanical performance of closed cell Al alloy foams. Acta Materialia, 1997. 45(12): p.

5245-5259.

198. http://www.australianperlite.com/perlite/. Perlite. 2016 [cited 2016 18/09/2016]; Available

from: http://www.australianperlite.com/perlite/.

199. Lau, K.-t., Lu, M., Chun-ki, L., Cheung, H.-y., Sheng, F.-L., and Li, H.-L., Thermal and

mechanical properties of single-walled carbon nanotube bundle-reinforced epoxy

nanocomposites: the role of solvent for nanotube dispersion. Composites Science and

Technology, 2005. 65(5): p. 719-725.

200. Hong, S.G. and Wu, C.S., DSC and FTIR Analyses of The Curing Behavior of

Epoxy/dicy/solvent Systems on Hermetic Specimens. Journal of Thermal Analysis and

Calorimetry, 2000. 59(3): p. 711-719.

201. Loos, M.R., Coelho, L.A.F., Pezzin, S.H., and Amico, S.C., The effect of acetone addition on

the properties of epoxy. Polímeros, 2008. 18: p. 76-80.

Page 223: Elastic And Mechanical Properties Of Expanded Perlite And ...

199

202. Arroyo, M., Pineda, J., and Romero, E., Shear Wave Measurements Using Bender Elements

in Argillaceous Rocks. 2010.

203. Suwal, L.P. and Kuwano, R., Disk shaped piezo-ceramic transducer for P and S wave

measurement in a laboratory soil specimen. Soils and Foundations, 2013. 53(4): p. 510-524.

204. Santamarina, J.C., Klein, A., and Fam, M.A., Elastic waves in the continuum, in Soils and

waves: Particulate materials behavior, characterization and process monitoring

2001, John Wiley and Sons Ltd.

205. Auld, B.A., Acoustic fields and waves in solids. 1973: Рипол Классик.

206. Krautkrämer, J. and Krautkrämer, H., Ultrasonic testing of materials. third ed. 1983, New

York: Springer-Verlag Berlin Heidelberg New York.

207. Wachtman, J.B., Mechanical and Thermal Properties of Ceramics: Proceedings. 1969: US

Department of Commerce, National Bureau of Standards.

208. Rice, R.W., Microstructure dependence of mechanical behavior of ceramics, in Treatise on

materials science and technology, R.K. MacCrone, Editor. 1977, Academic Press: New

York. p. 199-381.

209. Rossi, R.C., Prediction of the Elastic Moduli of Composites. Journal of the American

Ceramic Society, 1968. 51(8): p. 433-440.

210. Dutta, S.K., Mukhopadhyay, A.K., and Chakraborty, D., Assessment of Strength by Young's

Modulus and Porosity: A Critical Evaluation. Journal of the American Ceramic Society,

1988. 71(11): p. 942-947.

211. Hasselman, D.P.H. and Fulrath, R.M., Effect of Small Fraction of Spherical Porosity on

Elastic Moduli of Glass. Journal of the American Ceramic Society, 1964. 47(1): p. 52-53.

212. Panakkal, J.P., Willems, H., and Arnold, W., Nondestructive evaluation of elastic parameters

of sintered iron powder compacts. Journal of Materials Science, 1990. 25(2): p. 1397-1402.

213. Manghnani, M.H., Schreiber, E., and Soga, N., Use of ultrasonic interferometry technique

for studying elastic properties of rocks. Journal of Geophysical Research, 1968. 73(2): p.

824-826.

214. Castro, J. and Cashman, K.V., Constraints on rheology of obsidian lavas based on

mesoscopic folds. Journal of Structural Geology, 1999. 21(7): p. 807-819.

215. Richnow, J., Eruptional and post-eruptional processes in rhyolite domes. 1999.

216. Kramar, D. and Bindiganavile, V., Impact response of lightweight mortars containing

expanded perlite. Cement and Concrete Composites, 2013. 37(0): p. 205-214.

217. Fine, T., Sautereau, H., and Sauvant-Moynot, V., Innovative processing and mechanical

properties of high temperature syntactic foams based on a thermoplastic/thermoset matrix.

Journal of Materials Science, 2003. 38(12): p. 2709-2716.

218. Wood, D.M., Soil mechanics: a one-dimensional introduction. 2009: Cambridge University

Press.

219. Selvadurai, A.P., Elastic analysis of soil-foundation interaction. 2013: Elsevier.

220. Sawicki, A. and Świdziński, W., Elastic moduli of non-cohesive particulate materials.

Powder Technology, 1998. 96(1): p. 24-32.

221. Mei, J., Liu, Z., Wen, W., and Sheng, P., Effective dynamic mass density of composites.

Physical Review B, 2007. 76(13): p. 134205.

222. Wang, J.C., Young's modulus of porous materials. Journal of Materials Science, 1984. 19(3):

p. 801-808.

223. Ramakrishnan, N. and Arunachalam, V.S., Effective elastic moduli of porous solids. Journal

of Materials Science, 1990. 25(9): p. 3930-3937.

224. Arnold, M., Boccaccini, A.R., and Ondracek, G., Prediction of the Poisson's ratio of porous

materials. Journal of Materials Science, 1996. 31(6): p. 1643-1646.

Page 224: Elastic And Mechanical Properties Of Expanded Perlite And ...

200

225. Dunn, M.L. and Ledbetter, H., Poisson's ratio of porous and microcracked solids: Theory

and application to oxide superconductors. Journal of Materials Research, 1995. 10(11): p.

2715-2722.

226. Spriggs, R.M., Expression for Effect of Porosity on Elastic Modulus of Polycrystalline

Refractory Materials, Particularly Aluminum Oxide. Journal of the American Ceramic

Society, 1961. 44(12): p. 628-629.

227. Hasselman, D.P.H., On the Porosity Dependence of the Elastic Moduli of Polycrystalline

Refractory Materials. Journal of the American Ceramic Society, 1962. 45(9): p. 452-453.

228. Tai Te, W., The effect of inclusion shape on the elastic moduli of a two-phase material.

International Journal of Solids and Structures, 1966. 2(1): p. 1-8.

229. Martin, L.P., Dadon, D., and Rosen, M., Evaluation of Ultrasonically Determined Elasticity-

Porosity Relations in Zinc Oxide. Journal of the American Ceramic Society, 1996. 79(5): p.

1281-1289.

230. Hill, R., A self-consistent mechanics of composite materials. Journal of the Mechanics and

Physics of Solids, 1965. 13(4): p. 213-222.

231. Budiansky, B., On the elastic moduli of some heterogeneous materials. Journal of the

Mechanics and Physics of Solids, 1965. 13(4): p. 223-227.

232. Andersen, O., Waag, U., Schneider, L., Stephani, G., and Kieback, B., Novel Metallic Hollow

Sphere Structures. Advanced Engineering Materials, 2000. 2(4): p. 192-195.

233. Schwartz, L.M., Feng, S., Thorpe, M.F., and Sen, P.N., Behavior of depleted elastic

networks: Comparison of effective-medium and numerical calculations. Physical Review B,

1985. 32(7): p. 4607-4617.

234. Wojciechowski, K.W. and Novikov, V.V., Negative Poisson’s ratio and percolating

structures. Task Quarterly, 2001. 5: p. 5-11.

235. Arns, C.H., Knackstedt, M.A., and Pinczewski, W.V., Accurate Vp:Vs relationship for dry

consolidated sandstones. Geophysical Research Letters, 2002. 29(8): p. 44-1-44-4.

236. Markov, K. and Preziosi, L., Heterogeneous Media: Micromechanics Modeling Methods and

Simulations. Meccanica, 2001. 36(2): p. 239-240.

237. Ashkin, D., Haber, R.A., and Wachtman, J.B., Elastic Properties of Porous Silica Derived

from Colloidal Gels. Journal of the American Ceramic Society, 1990. 73(11): p. 3376-3381.

238. Hentschel, M.L. and Page, N.W., Elastic properties of powders during compaction. Part 3:

Evaluation of models. Journal of Materials Science, 2006. 41(23): p. 7902-7925.

239. Phani, K.K. and Niyogi, S.K., Young's modulus of porous brittle solids. Journal of Materials

Science, 1987. 22(1): p. 257-263.

240. Phani, K.K. and Niyogi, S.K., Elastic modulus-porosity relationship for Si3N4. Journal of

Materials Science Letters, 1987. 6(5): p. 511-515.

241. Zhang, L., Gao, K., Elias, A., Dong, Z., and Chen, W., Porosity dependence of elastic

modulus of porous Cr3C2 ceramics. Ceramics International, 2014. 40(1, Part A): p. 191-198.

242. Dı́az, A. and Hampshire, S., Characterisation of porous silicon nitride materials produced

with starch. Journal of the European Ceramic Society, 2004. 24(2): p. 413-419.

243. Rice, R.W., Porosity of Ceramics: Properties and Applications. 1998: CRC Press.

244. Knudsen, F.P., Dependence of Mechanical Strength of Brittle Polycrystalline Specimens on

Porosity and Grain Size. Journal of the American Ceramic Society, 1959. 42(8): p. 376-387.

245. Phani, K.K. and Sanyal, D., The relations between the shear modulus, the bulk modulus and

Young's modulus for porous isotropic ceramic materials. Materials Science and Engineering:

A, 2008. 490(1–2): p. 305-312.

246. Nielsen, L.F., Elastic properties of two-phase materials. Materials Science and Engineering,

1982. 52(1): p. 39-62.

Page 225: Elastic And Mechanical Properties Of Expanded Perlite And ...

201

247. Andersson, C.A., Derivation of the Exponential Relation for the Effect of Ellipsoidal Porosity

on Elastic Modulus. Journal of the American Ceramic Society, 1996. 79(8): p. 2181-2184.

248. Rice, R.W., Extension of the Exponential Porosity Dependence of Strength and Elastic

Moduli. Journal of the American Ceramic Society, 1976. 59(11-12): p. 536-537.

249. Brown, S.D., Biddulph, R.B., and Wilcox, P.D., A Strength–Porosity Relation Involving

Different Pore Geometry and Orientation. Journal of the American Ceramic Society, 1964.

47(7): p. 320-322.

250. Rice, R.W., Comparison of physical property-porosity behaviour with minimum solid area

models. Journal of Materials Science, 1996. 31(6): p. 1509-1528.

251. Rice, R.W., Comparison of stress concentration versus minimum solid area based

mechanical property-porosity relations. Journal of Materials Science, 1993. 28(8): p. 2187-

2190.

252. Rice, R.W. The porosity dependence of physical properties of materials: a summary review.

in Key Engineering Materials. 1996. Trans Tech Publ.

253. Maji, A.K., Schreyer, H.L., Donald, S., Zuo, Q., and Satpathi, D., Mechanical properties of

polyurethanefoam impact limiters. Journal of Engineering Mechanics, 1995. 121(4): p. 528-

540.

254. Hohe, J.r. and Becker, W., Effective stress-strain relations for two-dimensional cellular

sandwich cores: Homogenization, material models, and properties. Applied Mechanics

Reviews, 2001. 55(1): p. 61-87.

255. Kanny, K., Mahfuz, H., Carlsson, L.A., Thomas, T., and Jeelani, S., Dynamic mechanical

analyses and flexural fatigue of PVC foams. Composite Structures, 2002. 58(2): p. 175-183.

256. Yin, B., Li, Z.-M., Quan, H., Yang, M.-B., Zhou, Q.-M., Tian, C.-R., and Wang, J.-H.,

Morphology and Mechanical Properties of Nylon-1010-filled Rigid Polyurethane Foams.

Journal of Elastomers and Plastics, 2004. 36(4): p. 333-349.

257. Sciamanna, V., Nait-Ali, B., and Gonon, M., Mechanical properties and thermal conductivity

of porous alumina ceramics obtained from particle stabilized foams. Ceramics International,

2015. 41(2): p. 2599-2606.

258. Raj, R.E. and Daniel, B.S.S., Structural and compressive property correlation of closed-cell

aluminum foam. Journal of Alloys and Compounds, 2009. 467(1–2): p. 550-556.

259. Reitz, D.W., Schuetz, M.A., and Glicksman, L.R., A Basic Study of Aging of Foam Insulation.

Journal of Cellular Plastics, 1984. 20(2): p. 104-113.

260. Zwikker, C. and Smoluchowski, R., Physical properties of solid materials. Physics Today,

1955. 8: p. 17.

261. Barnes, H.A., Hutton, J.F., and Walters, K., An introduction to rheology. Vol. 3. 1989:

Elsevier.

262. Fu, S.-Y., Feng, X.-Q., Lauke, B., and Mai, Y.-W., Effects of particle size, particle/matrix

interface adhesion and particle loading on mechanical properties of particulate–polymer

composites. Composites Part B: Engineering, 2008. 39(6): p. 933-961.

263. Todd, M.G. and Shi, F.G., Molecular Basis of the Interphase Dielectric Properties of

Microelectronic and Optoelectronic Packaging Materials. IEEE Transactions on

Components and Packaging Technologies, 2003. 26(3): p. 667-672.

264. Todd, M.G. and Shi, F.G., Characterizing the interphase dielectric constant of polymer

composite materials: Effect of chemical coupling agents. Journal of Applied Physics, 2003.

94(7): p. 4551-4557.

265. Douce, J., Boilot, J.-P., Biteau, J., Scodellaro, L., and Jimenez, A., Effect of filler size and

surface condition of nano-sized silica particles in polysiloxane coatings. Thin Solid Films,

2004. 466(1–2): p. 114-122.

Page 226: Elastic And Mechanical Properties Of Expanded Perlite And ...

202

266. Hibbeler, R.C., Mechanics of materials. 2008, Singapore: Prentice Hall.

267. Kerner, E.H., The Elastic and Thermo-elastic Properties of Composite Media. Proceedings

of the Physical Society. Section B, 1956. 69(8): p. 808.

268. Counto, U.J., The effect of the elastic modulus of the aggregate on the elastic modulus, creep

and creep recovery of concrete. Magazine of Concrete Research, 1964. 16(48): p. 129-138.

269. Guth, E., Theory of Filler Reinforcement. Journal of Applied Physics, 1945. 16(1): p. 20-25.

270. Paul, B., PREDICTION OF ELASTIC CONSTANTS OF MULTIPHASE MATERIALS.

Technical Report No. 3, in Other Information: Orig. Receipt Date: 31-DEC-59. 1959. p.

Medium: X; Size: Pages: 22.

271. Kisi, E.H. and Howard, C.J., Stress and elastic constants, in Applications of neutron powder

diffraction. 2008, Oxford University Press: Oxford. p. 486.

272. Gupta, N. and Maharsia, R., Enhancement of Energy Absorption in Syntactic Foams by

Nanoclay Incorporation for Sandwich Core Applications. Applied Composite Materials,

2005. 12(3-4): p. 247-261.

273. Williams, J.L. and Johnson, W.J.H., Elastic constants of composites formed from PMMA

bone cement and anisotropic bovine tibial cancellous bone. Journal of Biomechanics, 1989.

22(6): p. 673-682.

274. Chavez, H.L., Alonso-Guzmán, E., Graff, M., and Arteaga-Arcos, J., Prediction of the Static

Modulus of Elasticity Using Four non Destructive Testing. Revista de la Construcción.

Journal of Construction, 2014. 13(1).

275. KURAMA, S. and Elif, E., Characterization of Mechanical Properties of Porcelain Tile

Using Ultrasonics. Gazi University Journal of Science, 2012. 25(3): p. 761-768.

276. Hardin, B.O. and Blandford, G., Elasticity of Particulate Materials. Journal of Geotechnical

Engineering, 1989. 115(6): p. 788-805.

277. Matikas, T., Karpur, P., and Shamasundar, S., Measurement of the dynamic elastic moduli of

porous titanium aluminide compacts. Journal of materials science, 1997. 32(4): p. 1099-1103.

278. Hicher, P.Y. and Chang, C.S., Evaluation of two homogenization techniques for modeling the

elastic behavior of granular materials. Journal of Engineering Mechanics, 2005. 131(11): p.

1184-1194.