You are on page 1of 102

UNIVERISTY OF MALTA

Faculty of Engineering

Department of Mechanical Engineering

FINAL YEAR PROJECT

B. ENG. (Hons)

Natural, Hybrid and Lightweight Composite Panels for Maritime Applications

by

Ritienne Sultana

A dissertation submitted in partial fulfilment


of the requirements of the award of

Bachelor of Engineering (Hons.) of the University of Malta


Copyright Notice

1. Copyright in text of this dissertation rests with the Author. Copies (by any process)
either in full, or of extracts may be made only in accordance with regulations held by
the Library of the University of Malta. Details may be obtained from the Librarian.
This page must form part of any such copies made. Further copies (by any process)
made in accordance with such instructions may not be made without the permission
(in writing) of the Author.

2. Ownership of the right over any original intellectual property which may be contained
in or derived from this dissertation is vested in the University of Malta and may not
be made available for use by third parties without the written permission of the
University, which will prescribe the terms and conditions of any such agreement.

ii
Abstract

Sandwich panels consist of two thin, strong and stiff facings that are separated from each
other by a thick, low-density core. Sandwich structures offer a high flexural stiffness
and an excellent strength-to-weight ratio. The advantages offered by sandwich panels
are increasingly being applied in the maritime industry, mainly in ship bulkheads and in
yacht and racing boat hulls.

The boat building industry is dominated by the use of glass-fibre reinforced polymers,
that create environmental problems related to emissions and end-of-life disposal. The
recent increase in environmental awareness has brought about a change in the materials
used. In many cases, natural and bio-based materials offer weight reductions, tailored
functionality and economic benefits, while being an ecological alternative. The aim of
this study is to investigate the suitability of bio-based sandwich panels for maritime
applications.

The bio-based sandwich panels considered in this study have Biotex® flax fibre facings
and Amorim® CoreCork NL20 as core, bonded together by ONE Super Sap® epoxy.
Single and multi-layered sandwich panels were considered, to study the effects that
intermediate layers have on the overall mechanical properties of the sandwich structure.
Sandwich behaviour was studied under different loading configurations by conducting
mechanical tests, adopting standard ASTM testing procedures.

Under flexure, the single-layered sandwich panels were able to reach loads that are
between 15 to 30 % higher than the loads reached by the multi-layered panels. Both
sandwich layups exhibited signs of face wrinkling on the top facing and face yielding
and/or fracture of the bottom facing. In some cases, face fracture led to crack propagation
through the cork thickness. In multi-layered sandwich panels, crack propagation through
the core was delayed by the intermediate flax layer.

The behaviour of the single and multiple sandwich layups was also investigated under
edgewise compression. The multi-layered specimens managed to withstand loads that
are approximately 6 % higher than those applied to the single specimens. It was observed
that the two sandwich configurations experienced different failure modes. The
single-layered sandwiches generally failed by Euler buckling, while the thinner core
layers of the multiple layup promoted failure by shear crimping.

iii
Acknowledgments

I would like to thank all staff members at the Faculty of Engineering, especially Prof.
Inġ. Claire De Marco for her patience, advice and dedicated assistance throughout this
study. Her expertise and technical knowledge of the subject proved very useful during
the course of this research project and I am eternally grateful for her continuous support
and guidance. I would also like to thank Mr. Kevin Farrugia of the Solid Body
Mechanics Lab for all his help with the fabrication and testing of the sandwich panels.

I also wish to express my deep sense of gratitude to my family and my boyfriend for
their continuous help, support and encouragement in both my academic and my personal
life throughout my years at the University of Malta.

iv
Table of Contents

1 Introduction and Objectives .................................................................................... 1

1.1 Sandwich Panels............................................................................................... 1

1.2 Study Overview ................................................................................................ 1

1.3 Objectives ......................................................................................................... 2

2 Literature Review ................................................................................................... 3

2.1 Introduction ...................................................................................................... 3

2.2 Historical Background of Sandwich Structures ............................................... 4

2.3 Current Alternatives ......................................................................................... 4

2.3.1 Face Materials ........................................................................................... 5

2.3.2 Core Materials........................................................................................... 7

2.3.3 Adhesive Materials ................................................................................... 9

2.4 Environmental Issues of Sandwich Panels ..................................................... 11

2.5 Mechanical Testing of Sandwich Panels ........................................................ 14

2.6 Conclusion...................................................................................................... 15

3 Theoretical Analysis ............................................................................................. 16

3.1 Overview ........................................................................................................ 16

3.2 Micromechanics of a Lamina ......................................................................... 16

3.3 Sandwich Beam Theory ................................................................................. 18

3.3.1 Flexural Rigidity of Sandwich Panels .................................................... 18

3.3.2 Stresses in Sandwich Panels ................................................................... 20

3.3.3 Transverse Shear Rigidity ....................................................................... 22

3.4 Long Beam Flexure Test ................................................................................ 23

3.4.1 Deflection due to bending moments ....................................................... 23

3.4.2 Deflection due to shear forces ................................................................ 26

3.4.3 Total Deflection ...................................................................................... 26

v
3.4.4 Calculation of Facing Properties............................................................. 27

3.4.5 Theoretical Failure Modes ...................................................................... 27

3.5 Short Beam Shear Test ................................................................................... 29

3.5.1 Deflection due to bending moments ....................................................... 29

3.5.2 Deflection due to shear forces ................................................................ 30

3.5.3 Total Deflection ...................................................................................... 31

3.5.4 Calculation of the Core Shear Properties ................................................ 31

3.5.5 Theoretical Failure Modes ...................................................................... 31

3.6 Flexural Stiffness, Transverse Shear Rigidity and Core Shear Modulus ....... 32

3.7 Edgewise Compressive Strength Test ............................................................ 33

3.7.1 Calculation of the Facing Compressive Stress ....................................... 33

3.7.2 Theoretical Failure Modes ...................................................................... 33

3.8 Conclusion...................................................................................................... 35

4 Sandwich Panel Construction and Testing ........................................................... 36

4.1 Overview ........................................................................................................ 36

4.2 Sandwich Panel Materials .............................................................................. 36

4.3 Sandwich Panel Design .................................................................................. 38

4.4 Sandwich Panel Fabrication ........................................................................... 39

4.5 Testing ............................................................................................................ 44

4.5.1 Long Beam Flexure Test......................................................................... 44

4.5.2 Short Beam Shear Test ........................................................................... 46

4.5.3 Edgewise Compression Test ................................................................... 47

4.6 Conclusion...................................................................................................... 48

5 Results and Discussion ......................................................................................... 49

5.1 Overview ........................................................................................................ 49

5.2 Results ............................................................................................................ 49

5.2.1 Long Beam Flexure Tests according to ASTM D7249 [30] .................. 52

vi
5.2.2 Short Beam Shear Tests following ASTM C393 [31] ............................ 58

5.2.3 Application of ASTM D7250 [43].......................................................... 64

5.2.4 Edgewise Compression Tests following ASTM C364 [33] ................... 69

5.3 Conclusion...................................................................................................... 72

6 Conclusions and Suggestions for Future Work .................................................... 74

6.1 Conclusions .................................................................................................... 74

6.2 Suggestions for Future Work ......................................................................... 76

7 References ............................................................................................................. 78

8 Appendix – Material Datasheets ........................................................................... 85

8.1 Biotex® 2/2 twill 200 g flax fibre [48] .......................................................... 85

8.2 Amorim® CoreCork NL20 [50] .................................................................... 87

8.3 ONE Super Sap® High Bio-Based Laminating Epoxy [51] .......................... 88

vii
List of Figures

Figure Page

Figure 2.1: Schematic diagram of a sandwich panel 3


Figure 2.2: Categories of natural fibres [14] 6
Figure 2.3: Flax, bio-resin and paper honeycomb sandwich by 13
EcoTechnilin [14]
Figure 2.4: Failure modes of sandwich beams [1] 15
Figure 3.1: Schematic of a unidirectional lamina 16
Figure 3.2: Cross-sectional area of the sandwich panel along 18
section X-X [1]
Figure 3.3: Cross-sectional area of a multiple layer sandwich panel 19
Figure 3.4: Direct and shear stresses acting across the cross-sectional 22
area of the sandwich panel provided that Ec ≪ Ef and t ≪ c
[1]
Figure 3.5: Loading configuration for 4-point bending as given in 23
ASTM D7249 [30]
Figure 3.6: Sections taken along the length of the beam 23
Figure 3.7: Shear and moment diagrams for a beam under 4-point 24
bending
Figure 3.8: Face yielding failure [39] 28
Figure 3.9: Different modes of face wrinkling failure [39] 28
Figure 3.10: Loading configuration for 3-point bending as given in ASTM 29
C393 [31]
Figure 3.11: Shear and moment diagrams for a beam under 3-point 30
bending
Figure 3.12: Core shear failure [39] 32
Figure 3.13: Loading configuration for the edgewise compressive strength 33
test
Figure 3.14a: Euler buckling [39] 34
Figure 3.14b: Shear buckling [39] 34
Figure 3.15a: Face sheet yielding [39] 35

viii
Figure 3.15b: Face wrinkling [39] 35
Figure 4.1a: Single layup 38
Figure 4.1b: Multiple layup 38
Figure 4.2: Preparation of cork and flax by drilling holes and tapering 39
the edges
Figure 4.3: Wetting the flax with resin 40
Figure 4.4: Vacuum bagging layup sequence 40
Figure 4.5: Two panels curing under pressure 41
Figure 4.6: Single and multiple layer test specimens 42
Figure 4.7: Long beam flexure set-up, 4-point standard configuration 44
Figure 4.8: Long beam, 4-point loading deflection 45
Figure 4.9: Long beam, 3-point loading deflection 46
Figure 4.10: Short beam shear test loading configuration 47
Figure 4.11: Edgewise compression test loading configuration 48
Figure 5.1a: Photomicrograph of the top section of the single-layered 50
panel
Figure 5.1b: Photomicrograph of the bottom section of the single-layered 50
panel
Figure 5.2a: Photomicrograph of the top section of the multi-layered 51
panel
Figure 5.2b: Photomicrograph of the bottom section of the multi-layered 51
panel
Figure 5.3: Applied load (kN) vs. mid-span deflection (mm) for the 53
single-layer, long beam flexure tests under 4-point loading
Figure 5.4: Applied load (kN) vs. mid-span deflection (mm) for the 53
multi-layer, long beam flexure tests under 4-point loading
Figure 5.5: Matrix yielding, following 4-point bending 54
Figure 5.6: Comparison of the theoretical and experimental values for 55
the facing modulus, Ef
Figure 5.7: Applied load (kN) vs. mid-span deflection (mm) for the 56
single-layer, long beam flexure specimen, S6 under 3-point
loading

ix
Figure 5.8: Applied load (kN) vs. mid-span deflection (mm) for the 56
multi-layer, long beam flexure specimen, M6 under 3-point
loading
Figure 5.9a: Top view of specimen S6 showing signs of face wrinkling 57
failure
Figure 5.9b: Side view of specimen S6 showing signs of face wrinkling 57
failure
Figure 5.10a: Top view of specimen M6, showing face fracture 58
Figure 5.10b: Side view of specimen M6, showing crack propagation 58
Figure 5.11: Applied load (kN) vs. mid-span deflection (mm) for the 59
single layer, short beam flexure tests under 3-point loading
Figure 5.12: Applied load (kN) vs. mid-span deflection (mm) for the 59
multiple layer, short beam flexure tests under 3-point loading
Figure 5.13: Comparison of the core shear strength, τCS [50] 61
Figure 5.14a: Fibre and resin failure of the single-layered specimen, S6 62
Figure 5.14b: Resin fracture of the multi-layered specimen, M6 62
Figure 5.15: Comparison of the theoretical (α = 0.5) and experimental 66
values calculated for the sandwich flexural stiffness
Figure 5.16: Comparison of the theoretical and experimental transverse 67
shear rigidity
Figure 5.17: Applied load (kN) vs. end shortening (mm) for the single 69
layer, edgewise compression tests
Figure 5.18: Applied load (kN) vs. end shortening (mm) for the multiple 69
layer, edgewise compression tests
Figure 5.19: A single-layered specimen exhibiting Euler buckling 72
Figure 5.20: Another single-layered specimen that failed by Euler 72
buckling
Figure 5.21: A buckled multi-layered specimen exhibiting face wrinkling 72
failure
Figure 5.22: A multi-layered specimen that failed by shear crimping 72

x
List of Tables

Table Page

Table 2.1: Natural and synthetic fibre properties [14, 16] 7


Table 4.1: Biotex® flax fibre properties [48] 36
Table 4.2: Amorim® CoreCork NL20 cork properties [50, 65] 37
Table 4.3: ONE Super Sap® epoxy system properties [51] 37
Table 4.4: Test specimen dimensions [30, 31, 33] 38
Table 4.5: Fibre weight and volume fractions of each panel 42
Table 4.6: Theoretical facing modulus and flexural rigidity of the 43
sandwich panels produced
Table 4.7: Theoretical transverse shear rigidity of the sandwich panels 43
produced
Table 5.1: Facing and core thicknesses of the sandwich panel 50
configurations
Table 5.2: Support and loading spans used in the flexure tests 51
Table 5.3: Facing stress and face yielding failure load for the single and 52
multiple layups
Table 5.4: Facing property results according to ASTM D7249 [30] 54
Table 5.5 The maximum stress developed in the top facing of specimen, 57
S6 and the face wrinkling strength of the single layup
Table 5.6: Core shear property results according to ASTM C393 [31] 60
Table 5.7: Comparison of the different failure loads for the short beam 62
specimens
Table 5.8: The different theoretical failure loads achieved by each short 63
beam specimen
Table 5.9: Sandwich beam flexural and shear stiffness according to ASTM 65
D7250 [43]
Table 5.10: Facing compressive stress results according to ASTM C364 70
[33]

xi
Nomenclature

Latin Symbols

A Cross-sectional area of the sandwich panel (mm2)


𝐴𝑓 Area of the face sheets (mm2)
b Width of the sandwich panel (mm)
B First moment of area, as a function of 𝑧 (N)
c Total core thickness (mm)
𝑐𝑀 Thickness of each core in the multiple layer sandwich layup (mm)
𝐶1 , 𝐶2 , … , 𝐶8 Constants of integration
𝑑 Distance between centroids of the face sheets (mm)
Distance between centroids of the cores in the multiple layer
𝑑𝐶
sandwich layup (mm)
𝑑𝑇 Total sandwich thickness (mm)
D Sandwich flexural rigidity (Nm2)
E Elastic modulus (MPa)
Ec Elastic modulus of the core (MPa)
Ef Elastic modulus of the facing in the longitudinal direction (MPa)
Et Elastic modulus of the facing in the transverse direction (MPa)
𝐺𝑐 Shear modulus of the core material (MPa)
𝐺𝑓 Shear modulus of the facing (MPa)
I Second moment of area (mm4)
𝐼𝑐 Second moment of area of the core (mm4)
𝐼𝑓 Second moment of area of the faces (mm4)
L Loading span (mm)
𝑀𝑥 Bending moment in the 𝑥-direction (Nm)
𝑀𝑦𝑖𝑒𝑙𝑑 Yielding moment (Nm)
P Applied load (N)
𝑃𝐶𝑆 Core shear failure load (N)
𝑃𝐶𝑆𝐵 Core shear buckling load (N)
𝑃𝐸𝐵 Euler buckling load (N)
𝑃𝑓𝑤 Face wrinkling failure load (N)

xii
𝑃𝑓𝑦 Face yielding failure load (N)
𝑃𝐺𝐵 Collapse load for general buckling (N)
𝑃𝑢𝑙𝑡 Ultimate applied load (N)
𝑃1000 Applied force corresponding to 𝜀1000 (N)
𝑃3000 Applied force corresponding to 𝜀3000 (N)
𝑡 Facing thickness (mm)
S Support span (mm)
𝑇𝑥 Transverse shear force in the 𝑥-direction (N)
U Transverse shear rigidity (N)
𝑣 Volume fraction
𝑉 Volume (𝑚3 )
𝑤 Total deflection (mm)
𝑤𝑏 Deflection due to bending moments (mm)
𝑤𝑠 Deflection due to shear forces (mm)
𝑤𝑖 Weight fraction of facing constituent
𝑊 Weight (kg)
𝑥, 𝑦, 𝑧 Cartesian coordinate system of the sandwich panel

xiii
Greek Symbols

𝛼 Reinforcement efficiency factor


𝜀𝑥 Strain in the 𝑥-direction
𝜀1000 Magnitude of recorded strain value closest to 1000 micro-strain
𝜀3000 Magnitude of recorded strain value closest to 3000 micro-strain
𝜎𝑐 Bending stress in the core (MPa)
𝜎𝑓 Bending stress in the facing (MPa)
𝜎𝑓,𝑐 Facing compressive stress (MPa)
𝜎𝑓,𝑢𝑙𝑡 Ultimate facing stress (MPa)
𝜎𝑓𝑤 Critical wrinkling strength of the facing (MPa)
𝜎𝑓𝑦 Yield strength of the facing (MPa)
𝜎1000 Facing stress corresponding to 𝜀1000 (MPa)
𝜎3000 Facing stress corresponding to 𝜀3000 (MPa)
𝜌 Density (𝑘𝑔/𝑚3 )
𝜏 Shear stress (MPa)
𝜏𝑐 Core shear stress (MPa)
𝜏𝑓 Facing shear stress (MPa)
𝜏𝐶𝑆 Core shear strength (MPa)
𝛾 Transverse shear strain
𝛾0 In-plane core shear strain

xiv
1 Introduction and Objectives

1.1 Sandwich Panels

Sandwich panels consist of two thin, stiff facings and a low-density core. The facings,
manufactured either of composite laminates or metallic sheets, resist bending stresses
and in-plane loadings, while the core supports the shear forces and stabilises the faces,
preventing buckling or wrinkling. Sandwich panels are able to attain a high bending
stiffness and an excellent strength-to-weight ratio [1]. The advantages offered by
sandwich panels are being applied in high performance applications including
construction, aerospace and the boat building industry [2].

Sandwich panels offer a level of freedom and flexibility that is not offered by other
materials alone. Due to the high strength-to-weight ratio, sandwich panels are being
increasingly applied in the bulkheads of ships and in yacht and racing boat hulls [3].
Steel sandwich panels applied in ship bulkheads offer numerous benefits, including
weight and space savings due to larger unsupported spans, improved crashworthiness
properties and competitive prices [4]. A reduction in the ship’s weight above the
waterline improves both its stability and its manoeuvrability [3]. The use of sandwich
panels in yacht and racing boat hulls enables stiff, self-supporting hulls to be produced.
The result is light, fast and fuel-efficient boats [5].

1.2 Study Overview

The boat building industry is dominated by the use of glass-fibre reinforced polymers
(GFRP), made from non-renewable sources. The production of GFRP requires
energy-intensive processes, emitting significant amounts of greenhouse gases. Recent
studies have shown that the use of natural, renewable materials in the production of
sandwich panels offers various economic and technological advantages, while being
environmentally friendly [6]. The aim of this study is to test the suitability of sandwich
panels, having flax facings and a cork core, intended for maritime application. Flax and
cork are both natural materials, while the resin used to bond the faces to the core is
bio-based. Furthermore, both single and multiple layer sandwich panels having the same
thickness will be tested, to study the effects and possible benefits that the different layups
have on the overall sandwich properties.

1
Chapter 2 covers the literature review, to acquire knowledge on sandwich panels and
natural materials that are used for sandwich panel construction. Additionally, different
standards for the mechanical testing of such sandwich panels were also evaluated. The
theoretical analysis relating to all aspects considered in this study is presented in
Chapter 3. Chapter 4 deals with the manufacturing of the sandwich panels and testing
methods applied, following the respective American Society for Testing and Materials’
(ASTM) standards. Chapter 5 discusses the results obtained from the experimental
testing of Chapter 4 and compares the results to the theoretical values attained from
Chapter 3. Finally, Chapter 6 draws conclusions from the results obtained and
observations made while carrying out this study and offers possible improvements and
future work, extending the project presented here.

1.3 Objectives

In summary, the main objectives of this study are to:

- Review literature on the state-of-the-art and current applications of natural


materials in sandwich panels.
- Obtain the facing, core and overall sandwich properties of flax and cork
sandwich panels through the application of different ASTM standards.
- Predict the expected sandwich properties through theoretical analysis and
compare them to the experimental results.
- Analyse, discuss and conclude the possible benefits of using multiple layer
sandwich layups over conventional single layer sandwiches.

2
2 Literature Review

2.1 Introduction

As defined in ASTM C279 [7], a structural sandwich panel is ‘‘a laminar construction
comprising a combination of alternating dissimilar simple or composite materials
assembled or intimately fixed in relation to each other so as to use the properties of each
to attain specific structural advantages for the whole assembly’’ [7]. Sandwich panels
are made up of two thin, strong and stiff faces that are separated from each other by a
thick, lightweight core material. The interactions between the components of a sandwich
panel allow for excellent mechanical properties to be achieved. Properties include a high
flexural stiffness and an excellent strength-to-weight ratio [8].

The low-density core resists shear and buckling forces, while separating the two faces
by a distance to increase the moment of inertia of the sandwich panel [1]. As a result,
the bending strength and lateral stiffness of a sandwich beam are superior to those of a
single solid plate of the same total weight, fabricated of the same material as the faces.
The faces are adhesively bonded to the core to ensure efficient load transfer. Adequate
bonding and load transfer ensure that the sandwich panel is utilised to its full potential
[9]. The components of a sandwich panel are shown schematically in Figure 2.1.
Sandwich panels are known for the ability to produce lightweight components.
Applications include aerospace, marine, wind energy and transportation industries.

Adhesive joint

Face material Core material

Adhesive joint

Figure 2.1: Schematic diagram of a sandwich panel

3
2.2 Historical Background of Sandwich Structures

Sandwich structures, making use of lightweight cores, are very useful in


high-performance structural applications requiring minimum weight. The advantages of
having two stiff panels separated by a distance were first noted by a Frenchman, Duleau
in 1820. However, it was around 100 years later that the concept was first applied
commercially [1]. In 1919 sandwich panels were used for the first time in the aircraft
industry. Constructed of thin mahogany facings and an end-grain balsa wood core, these
sandwiches were used as the primary structure of the pontoons of a seaplane. Following,
in the period between World War I and World War II, the primary structure of Italian
seaplanes was constructed by gluing plywood skins to a balsa wood core [10].

Sandwich panels were mass-produced for the first time in England during World War II
for the production of the Mosquito aircraft. Parts of the airframe of the Mosquito were
constructed using veneer facings and a balsa wood core. The use of sandwich panels was
mainly due to the shortage of other commonly used materials, however, the excellent
performance displayed by this aircraft led to the acceptance and appreciation of
sandwich structures as a means of improving efficiency and performance. Further
research was then carried out, leading to the publication of the first theoretical works on
sandwich constructions in the late 1940s. New developments allowed for the production
of adhesives with a better flow during curing. The improved adhesives then led to the
production of the first all-aluminium sandwich panel in 1945, making use of a
honeycomb core [10].

2.3 Current Alternatives

The properties of sandwich structures are highly dependent on the geometry, as well as
on the materials that are chosen for the three components, namely the faces, the core and
the adhesive. The materials and geometry of a sandwich panel may be altered in order
to obtain the required mechanical properties. Besides the mechanical properties
required, material selection also takes into consideration environmental resistance,
surface finish, manufacturing method, wear-resistance and cost [1]. Many engineering
applications make use of sandwich composites composed of GFRP faces and a foam or
honeycomb core. Glass fibres and other synthetic materials used in the production of
sandwich panels are made from non-renewable sources. The production of these

4
synthetic materials generates excessive amounts of carbon dioxide, contributing to
global warming. Synthetic materials also make sandwich panels difficult to dispose of
at the end of their lifespan. Consequently, the materials commonly end up in landfills.
Natural options are available, offering economic, ecologic and technological advantages
over synthetic materials. Natural materials are derived from renewable sources and can
be recycled at the end of their service life [9].

2.3.1 Face Materials

‘‘Almost any structural material which is available in the form of thin sheets may be
used to form the faces of a sandwich panel’’ [11]. From this phrase by Allen, it is evident
that the range of materials that may be used as faces for sandwich panels is vast. In fact,
this is one of the main reasons why sandwich panels are so versatile. Face materials are
required to have a high stiffness to give the panel a high flexural rigidity, high tensile
and compressive strengths, a good surface finish, impact resistance, environmental
resistance and wear resistance. Face materials can be metallic or non-metallic materials.
Metallic faces, including materials like steel, stainless steel and aluminium alloys, have
been extensively applied in the aerospace, automotive and marine industries.
Non-metallic faces include materials like plywood, cement, reinforced plastic and
composite laminates [1].

The introduction of composite laminates as face materials has had a great impact on
sandwich structures. The high stiffness, strength, fatigue resistance, low density and ease
of shaping offered by composite laminates make them excellent alternatives for
structural applications. The anisotropic nature of composite laminate faces may seem to
be a disadvantage, however, it offers an opportunity to tailor the properties of the
sandwich panel according to the in-service loadings. Fibre materials commonly used in
the faces of sandwich panels include glass, carbon and aramid in a petroleum-based
matrix. All these are synthetic, non-renewable materials. Production methods of these
synthetic fibres are energy-intensive, raising environmental concerns [12]. The recent
increase in environmental awareness has brought about a change in the materials and
manufacturing processes used in the production of sandwich structures. There is a
growing need to develop and produce materials from renewable sources. The different
categories of natural fibres are shown in Figure 2.2.

5
Figure 2.2: Categories of natural fibres [14]

Natural fibres tend to have a high moisture uptake, limited thermal stability and show
poor bonding to polymers [13]. The drawbacks of natural fibres may be overcome
through a range of physical, chemical and additive treatments that modify the fibre
characteristics [14]. In general, the tensile strength and Young’s modulus of natural
fibres are less than those of commonly used synthetic fibres, however, the lower density
of natural fibres gives them a higher specific strength and specific modulus, making
them ideal for lightweight applications [15]. Natural fibres are being applied in
significant quantities, particularly in automotive interiors. Some properties of both
natural and synthetic fibres are given in Table 2.1, for comparison purposes.

Natural fibre-reinforced polymers incorporate animal, mineral and plant-based fibres.


Plant-based fibres are the most researched of the three categories. Bast fibres are fibres
taken from the stem of the plant and include materials like flax, hemp and jute. Flax and
hemp are of particular interest in Europe as the plants are native to the region. The fibres
are biodegradable and recyclable as they are made of a combination of cellulose,
hemicellulose and lignin. Another environmental benefit is that plant-based fibres are
carbon-positive, meaning that they absorb more carbon dioxide than they emit [14].

The mineral-based basalt fibres have also been studied as a sustainable alternative.
Basalt fibres are produced from basalt rock, the most common rock found in the earth’s
crust. No additives are added during the manufacturing process, reducing the
environmental impact. Furthermore, basalt is more chemically inert than glass, making
basalt fibres ideal for storage and transportation of corrosive liquids and gases [16].

6
Young’s Tensile Strength
Fibre Material Density (kg/m3)
Modulus (GPa) (MPa)

Jute 1460 20 – 55 200 – 800

Flax 1400 60 – 80 500 – 900


Natural

Hemp 1480 30 – 70 300 – 800

Sisal 1330 9 – 38 100 – 800

Basalt 2660 85 – 91 4500

Aramid 1440 70 – 112 2757


Synthetic

Carbon 1580 125 – 400 4127

E-Glass 2600 72 – 77 3450

Table 2.1: Natural and synthetic fiber properties [14, 16]

2.3.2 Core Materials

The core material is perhaps the most important component of a sandwich panel. The
core acts as the web in an I-beam, separating the load-bearing flanges, while the latter
carry the main tensile and compressive loads, like the faces of a sandwich panel [17].
The function of the core is to increase the thickness of the sandwich panel and thus,
increase the overall stiffness. The core also carries the shear loadings applied to the
sandwich panel. In order to perform its function efficiently, the material chosen for the
core should have a high shear strength and shear modulus. A low-density material is
required so that the core contributes as little as possible to the overall weight of the
sandwich panel. Sandwich properties, such as thermal and acoustic insulation, also
depend on the core material chosen [1].

A large selection of core materials is available, offering a wide range of densities,


lifespans, mechanical properties and costs. Two of the most commonly used,
high-quality core materials are Nomex® honeycomb and PVC foam. Nomex®
honeycomb is commonly used in lightweight, non-metallic composite constructions. It
is a commercial-grade honeycomb made of aramid fibre paper, coated with heat resistant
phenolic resin. The honeycomb structure is achieved by shaping the fibre paper into
sheets of hexagonal cell structures and then dipping it into phenolic resin to increase its

7
stability and improve the mechanical properties. Nomex® honeycomb offers a high
strength-to-weight ratio due to its low density and excellent resiliency [18]. Other
materials used to produce honeycomb cores include plain paper, aluminium and
polypropylene. The high strength-to-weight ratios offered by honeycomb structures
make them suitable for aerospace, marine, military, construction, sports and automotive
applications [19].

PVC foam is a closed-cell foam available in a wide range of densities. It is a lightweight


material with excellent mechanical and structural properties. There are two different
types of PVC foam; cross-linked PVC foam is rigid while linear PVC foam is ductile
[18]. PVC foams are commonly used in transportation, aerospace, military, wind energy
and industrial applications. The moisture resistance of PVC foams also makes them
suitable for marine applications. Besides PVC, synthetic foam cores are also produced
from polyurethane, polystyrene and polyetherimide [20].

Both Nomex® honeycomb and PVC foam are synthetic core materials. Lightweight
green core materials are available. Some of these materials, such as balsa wood, are well
established and have been used in sandwich structures for many years. Other green core
options are emerging technologies that are still being proven in widespread applications.
Green core options include bio-based polymer honeycombs, bio-based polymer foams,
nano cellulose foams and cores from trees.

Honeycomb geometries are very efficient in providing lightweight core materials. A


green alternative is producing honeycomb cores from a bio-based polymer. EconCore
produce a honeycomb core made of a 100 % bio-based polylactic acid (PLA) polymer.
The only downside of this core is that its upper-temperature range is quite low, at only
55 ºC. Paper and cardboard honeycombs are also green core options.

PLA, along with tannin, starch and flax oil, have been used to produce green foam cores.
Studies show that these bio-based polymer foams are not suitable for structural
applications due to low mechanical properties however, they may still be used as
packaging materials and thermal insulation. Nano cellulose foams are produced from
cellulose, the primary structural component of plants and the most abundant organic
compound on the planet. Through careful processing, it is possible to extract nano-
crystals from the cellulose and then assemble them into cellular, foam-like structures.
The technology is still in its early stages but it is showing promising results [14].

8
Core materials produced from trees include balsa wood and cork. Balsa wood was the
first material used as a core in load-carrying sandwich panels. Balsa has a closed-cell
structure making the wood very sensitive to humidity, with properties declining rapidly
with increasing water content. To overcome this problem, balsa is kiln-dried to remove
the water contained in the cells and it is also used in an ‘end-grain’ configuration.
Through this configuration, the sandwich panel gains an improved resistance to local
indentation and face wrinkling [1]. As a core material, balsa wood is denser than many
honeycombs and polymer foams, however, it provides good strength and stiffness and
exhibits an excellent impact resistance. The fire resistance of balsa wood is superior
to that of polymer foams and thus, it has found use in various marine applications over
the years [14].

Cork is obtained from the bark of the cork oak tree. Cork is renewable since the bark
may be stripped from the cork oak once every nine years. Europe, particularly Portugal,
produces about 80 % of the global cork supply. Cork has a closed-cell structure, mostly
made of cellulose, lignin and suberin. The cork’s closed-cell structure makes it versatile
and gives it unique properties. Cork is impermeable to gases and liquids and it is also a
good thermal, electric, sound and vibration insulator. These properties, together with its
reasonable compressive strength, make cork a suitable core material [21]. Furthermore,
cork has a mechanical performance that is comparable to that of some high-performance
foams but with a higher energy absorption capacity. Therefore, cork has better
crashworthiness properties for applications where impact loading is expected. Cork
agglomerates with lower densities have better thermal properties, an important issue in
the design of mechanically efficient structures with lightweight requirements, such as
aerospace components [22].

2.3.3 Adhesive Materials

The adhesive material is the final element making up a sandwich panel. The choice of
adhesive is primarily dependent on the mechanical requirements of the structure. Fatigue
resistance, heat resistance, strength, ageing and creep must all be taken into
consideration. Furthermore, the adhesive chosen should be able to provide sufficient
bond strength in the operating environment of the sandwich panel [1]. Adhesives used
in sandwich panel constructions are polymeric materials. There are two categories of
polymers; thermosetting resins and thermoplastics.

9
Thermosetting resins are made of polymer chains that cross-link during curing, forming
irreversible chemical bonds. Therefore, thermosetting resins cannot be re-melted and
re-formed once cured. Thermoplastic resins are not cross-linked and can easily be
recycled by re-melting and re-processing. Most polymers used in sandwich panels are
petroleum-based thermosetting resins. Common thermosetting resins used include
polyurethane and epoxy resins [23].

Polyurethane resins are the most widely used resins for bonding sandwich panels. The
resin bonds well to a variety of surfaces, including concrete. Polyurethane is mainly used
in bonding foam or balsa wood core sandwich panels, providing a good impact strength
and abrasion resistance. Polyurethane resins also have a high chemical resistance and
can be made fire-retardant and water-resistant. The adhesives contain no solvents and
are the least toxic of all resins [1].

Epoxy is a low-temperature curing resin characterised by a high strength-to-weight ratio,


good electrical properties, chemical resistance and low moisture absorption. Epoxy does
not require any solvents and this allows it to be used with almost every type of core
material. Epoxy does not create any volatile by-products during curing and exhibits good
dimensional stability and a high resistance to corrosive environments. With the right
additives, epoxy can achieve outstanding thermal resistance and electrical insulation.
Epoxy is primarily used for the production of high-performance sandwich panels with
excellent mechanical properties that are applied in the aerospace, defence, chemical
plant and automotive industries [24].

All resins mentioned so far are derived from petroleum. Petroleum is a non-renewable
source and this imposes a limitation to the polymer industry due to the continuous
depletion of crude oil, the ever-changing price of oil and various environmental concerns
regarding sustainability, gas emissions, end-of-life disposal and recyclability. Therefore,
efforts have been made towards the design and production of polymeric materials from
renewable sources. Research has shown that both thermoplastic and thermosetting
polymeric materials may be produced from natural, raw, renewable materials [14].
Bio-based polymers tend to be more expensive then petroleum-based polymers however,
prices are expected to fall as production increases. Furthermore, prices of
petroleum-based polymers are also expected to rise as crude oil reserves diminish [23].

10
Thermoplastic bio-based polymers can be made from starch, cellulose, polyesters and
lignin. Starch derived from corn, rice, potatoes or wheat can be used to produce
inexpensive polymers. Starch-based polymers tend to be brittle, sensitive to moisture
and cannot withstand elevated temperatures, however, the drawbacks can be overcome
by the addition of plasticisers that modify the mechanical properties of the polymer.
Plant cellulose can be used to produce cellulosic plastics through the acetylation process.
Cellulose is biodegradable, environmentally friendly and renewable.

Thermoplastic polyesters can be derived from both fossil fuels and renewable sources.
Polycaprolactone (PCL) is a biodegradable polyester derived from fossil oil while PLA
is a biodegradable polyester made from a bio-based source. PLA is a versatile polymer
produced through the conversion of starch or by fermenting sugars. Recently PLAs have
become economically feasible for wider engineering use and are commercially available
from a number of suppliers. Lignin is a natural, biodegradable material commonly found
in natural wood, binding the strong and stiff cellulose units together. Once separated,
the lignin can be chemically modified to produce a polymer that can be processed like
synthetic thermoplastics.

Thermosetting bio-based resins can be produced from plant oils or polyfurfuryl alcohol
(PFA). Mechanical properties comparable to those of petroleum-based counterparts are
hard to achieve and synthetic chemicals still need to be added to produce the cross-links.
Current research is focused on obtaining cross-links from biological sources. Research
in Europe is mainly centred on linseed, rapeseed and sunflower oils that are readily
available, have a low toxicity and are relatively cheap. PFA resins, derived from furfuryl
alcohol, a renewable alcohol, are thermally stable, fire retardant and chemically resistant
resins [14].

2.4 Environmental Issues of Sandwich Panels

Sandwich panels offer various environmental benefits due to the low weight, excellent
mechanical properties, good corrosion resistance and the possibility of being sustainably
sourced. In the automotive sector, sandwich panels enable weight reduction, saving fuel
throughout the lifetime of the vehicle. Besides the in-service environmental benefits of
sandwich panels, there are other environmental and social implications associated with
the production and end-of-life disposal of sandwich panels that have to be considered.
These include health and safety, the emission of volatile organic compounds, energy

11
consumption and any toxic by-products. The use of natural materials and the application
of new technologies help, in part, in addressing these issues, however, sandwich panels
still have drawbacks. Life Cycle Assessment (LCA) methods have to be carried out to
ensure that the drawbacks do not outweigh the benefits. LCA is a modelling system that
enables the quantification and validation of social, environmental and economic
impacts, allowing for the comparison of different materials and processes [14]. GaBi is
an LCA software that models every element of a product from a life cycle perspective,
allowing businesses to produce sustainable yet cost-effective products [25].

Naturally-derived fibres and polymers are thought of as a ‘greener’ option when


compared to synthetic counterparts however, this may not always be the case. LCA
methods have shown that in instances where a certain component stiffness is required, a
sandwich panel using natural fibres in a synthetic resin is less competitive than a
sandwich panel using carbon fibres in the same matrix, since the former has to be thicker
to achieve the same performance and thus, requires more resin [9].

The end-of-life disposal option chosen for a material has a great influence on the
environmental impact of its LCA. One of the main advantages of using natural materials
is their environmentally positive end-of-life options, including biodegradation and
composting. Some bio-based materials can be engineered to be biodegradable at the end
of their lifespan. Natural fibres and some bio-based polymers are degraded by
micro-organisms and then composted. The biodegradation process releases the CO2
contained within the natural materials.

Non-biodegradable materials can be either recycled or incinerated. Sandwich panels


using glass fibre-reinforcements are recycled by chopping up the material. The granules
are then reprocessed by injection moulding them into other products. Recycling natural
fibre-reinforced sandwiches using this method would result in the degradation of the
natural fibres due to the repeated heating and processing steps required. Another option
for disposing of fully bio-based sandwich composites is incineration. The benefits of
incineration depend highly on the materials making up the sandwich panel. Natural
fibres do not use petroleum-based sources and so do not release any additional CO2 when
burned. Incineration generates electricity and reduces the materials to ashes. The ashes
can then be mixed with concrete in the construction industry.

12
Sustainability is an important consideration in the design and manufacture of new
materials. Research and development of a bio-based material or product should favour
renewable resources, optimised land use, improved biodiversity, minimised
environmental impact and improved energy efficiency, in order to obtain a favourable
LCA. Recently, EcoTechnilin in the UK produced a composite sandwich panel
consisting of non-woven flax faces in a bio-resin matrix and a paper honeycomb core,
as shown in Figure 2.3. The fully bio-based sandwich panel was used as a load floor in
the Jaguar F-type convertible [14].

Figure 2.3: Flax, bio-resin and paper honeycomb sandwich by EcoTechnilin [14]

Cork sandwich composites have also shown potential. Research done so far was mostly
concentrated on cork sandwich panels having carbon or glass fibre-reinforced polymer
faces. Ferriera et al. studied the properties of sandwich panels constructed of a thick cork
core and GFRP faces. The sandwich panels offered a much higher fire resistance than
those having a synthetic core material [26]. With cork being a natural material, the focus
was recently shifted to fully bio-based cork sandwich panels.

Lakreb et al. studied the mechanical performance of sandwich panels made of a cork
agglomerate core and pinewood veneer faces. Results showed that these sandwich
panels may be applied as construction materials for panelling or partition walls [27]. A
study by Sadeghian et al. showed that sandwich composites made of a cork core and flax
fibre-reinforced polymer faces showed a structural performance comparable to that of a
sandwich panel made of a synthetic honeycomb core material and GFRP faces [28]. Mak
et al. studied the possibility of replacing GFRP faces with faces made of unidirectional
flax fibres in a pine oil epoxy matrix. These studies have shown that a sandwich panel

13
using three layers of flax fibres provided an equivalent strength and stiffness to a
sandwich panel using one layer of glass fibres. Moreover, the bio-based option exhibited
better deformability [29].

2.5 Mechanical Testing of Sandwich Panels

High performance, load-bearing components are required for high-tech applications,


such as in the aerospace and automotive industries. Sandwich components need to be
lightweight and offer some damage resistance, while meeting stiffness and strength
criteria. For these requirements to be met, structurally efficient designs are required. The
materials used for the different components of the sandwich panel, as well as the
geometry of each component, determine the efficiency of the design. Proper analysis of
the properties of a sandwich panel requires an understanding of the mechanical
behaviour of both the facing and the core materials. The overall mechanical properties
of sandwich panels can be evaluated through the corresponding ASTM standard testing
procedures. Common tests include long beam flexure [30], short beam shear [31],
flatwise tensile [7] and compressive tests [32] and edgewise compression [33].

According to ASTM D7249 [30], the long beam flexure test makes use of a
4-point loading fixture and subjects the test panel to a bending moment normal to the
plane of the sandwich panel. Acceptable failure modes are those which are internal to
one of the faces. The test can be used to evaluate the strength and stiffness of the
sandwich panel facings, as well as the sandwich panel’s flexural and shear stiffness. As
stated in ASTM C393 [31], the short beam shear test also subjects the sandwich test
panel to a bending moment normal to the plane of the panel. Acceptable failure modes
are core shear or core-to-face bond failure. Force and deflection measurements taken
while loading the panel, can be used to calculate the core shear strength or core-to-face
shear strength, whichever is lower, as well as the panel’s flexural and shear stiffness.

Following ASTM C297 [7], the flatwise tensile test subjects the sandwich panel to a
uniaxial tensile force normal to the plane of the panel. Test results can be used to
calculate the tensile strength of the core, as well as to provide information on the strength
and quality of the core-to-face bond. According to ASTM C365 [32], the flatwise
compressive test is similar to the tensile one but a compressive load is applied in this
case. The purpose of this test is to determine the modulus of the core material, as well
as the flatwise compressive strength and modulus of the sandwich panel. As stated in

14
ASTM C364 [33], the edgewise compressive test studies the load-carrying capability of
the faces of the sandwich panel. Several failure modes are possible, namely shear
crimping, general buckling, face dimpling and face wrinkling. Readings taken during
this test can be used to evaluate the properties of the sandwich panel when loaded in
compression. Different failure modes of sandwich panels are depicted in Figure 2.4.

Figure 2.4: Failure modes of sandwich beams


(a) face yielding/fracture (b) core shear failure (c) and (d) face wrinkling (e) general buckling
(f) shear crimping (g) face dimpling and (h) local indentation [1]

2.6 Conclusion

Various industries show an increasing need for lightweight structures with a high
strength-to-weight ratio. There was significant growth in the sandwich panel technology
with a rising demand for developing new materials. Typical materials used to construct
sandwich panels are mainly synthetic and not environmentally friendly [8]. The recent
increase in environmental awareness has brought about a change in the materials used,
with various bio-based options showing promising results. Green sandwich panels are
already being used in various industries including the automobile, construction and
marine industries, however, further research is required for a more widespread
application of fully bio-based sandwich panels [22].

15
3 Theoretical Analysis

3.1 Overview

The theory required to calculate the material properties of sandwich constructions will
be derived in this chapter. Laminate micromechanics and the sandwich beam theory will
both be applied. A theoretical analysis of standard ASTM testing procedures will also
be carried out, allowing for a better understanding of the behaviour of single and
multi-layered sandwich constructions.

3.2 Micromechanics of a Lamina

Micromechanics is the study of composite materials, taking into consideration the


interaction of the constituent materials on a microscopic scale and analysing the effects
on the properties of the composite material [34]. In this study, micromechanics will be
applied to analyse the fibre and matrix volume fractions of the sandwich panel facings,
that are used to predict material properties, such as the facing modulus.

Equations (3.1) and (3.3) give the upper and lower bounds of the elastic modulus of a
composite lamina, respectively. The equations assume a unidirectional lamina, as shown
in Figure 3.1. The longitudinal direction is that parallel to the fibres, while the transverse
direction is that perpendicular to the fibres. It is also assumed that there is a perfect bond
between the fibres and the matrix and that the lamina is free of any voids [35].
2
Transverse direction

1
Longitudinal direction

Figure 3.1: Schematic of a unidirectional lamina

The average tensile modulus in the fibre direction of a unidirectional composite lamina,
𝐸1𝑓 is given by equation (3.1):

𝐸1𝑓 = ∑ 𝐸𝑖 𝑣𝑖 = 𝐸𝑓𝑏 𝑣𝑓𝑏 + 𝐸𝑚 𝑣𝑚 (3.1)


𝑖

16
where 𝑣 is the volume fraction, 𝐸 is the modulus and subscript 𝑖 represents either one of
the facing constituents, namely the fibre (𝑓𝑏) or the matrix (𝑚). The approximation
given by equation (3.1) is called the parallel model, also known as the rule of mixtures.
The rule of mixtures gives the upper bound for the elastic modulus that any composite
laminate can achieve.

The average tensile modulus of a unidirectional composite lamina in the direction


transverse to the fibres, 𝐸2𝑓 is given by equation (3.2).

1 𝑣𝑖 (3.2)
=∑
𝐸2𝑓 𝐸𝑖
𝑖

𝐸𝑓𝑏 𝐸𝑚 (3.3)
𝐸2𝑓 =
𝐸𝑓𝑏 𝑣𝑚 + 𝐸𝑚 𝑣𝑓𝑏

The approximation given by equation (3.2) is called the serial model, also known as the
inverse rule of mixtures. Equation (3.3) gives the lower bound of the elastic modulus
that any composite laminate may achieve.

The rule of mixtures may be modified to predict the tensile modulus of a composite
laminate consisting of fibre orientations other than unidirectional. The modification is
called the practical rule of mixtures and it can be used to study fibre configurations such
as bi-directional symmetric and random orientations. It considers only the fibres that are
aligned with the longitudinal direction of the laminate, by introducing a reinforcement
efficiency factor. The modified rule of mixtures is given by:

𝐸1𝑓 = ∑ 𝛼𝑖 𝐸𝑖 𝑣𝑖 (3.4)
𝑖

where 𝛼 is the reinforcement efficiency factor, equal to the fraction of fibres acting in
the longitudinal direction.

Calculation of the facing modulus requires knowledge of the volume fractions of each
of the facing constituents, namely the fibre and the matrix. The volume fraction of each
facing constituent, 𝑣𝑖 is given by:

𝑉𝑖 𝑤𝑖 ⁄𝜌𝑖 (3.5)
𝑣𝑖 = =
𝑉 ∑𝑖(𝑤𝑖 ⁄𝜌𝑖 )

where 𝑉 is the volume in 𝑚3 . The density of each constituent is given by 𝜌𝑖 , while 𝑤𝑖


represents the weight fraction of each constituent.

17
The weight fraction of each constituent is given by equation (3.6):

𝑊𝑖 (3.6)
𝑤𝑖 =
𝑊
where 𝑊𝑖 represents the weight of each facing constituent and 𝑊 represents the total
weight, both in kg [1].

3.3 Sandwich Beam Theory

The Euler-Bernoulli beam theory, also known as the classical beam theory, is commonly
used in the design and analysis of engineering structures, like buildings and bridges. One
major assumption of the classical beam theory is that the beam supports the loading only
by bending and it does not take into consideration transverse shear deformations [36].
In the case of sandwich panels, transverse shear deformations must be taken into account
and so, the Timoshenko beam theory will be applied [1].

3.3.1 Flexural Rigidity of Sandwich Panels

Consider the sandwich panel of cross-sectional area A, shown in Figure 3.2.

𝑏
𝑡 𝐸𝑓

𝑋
𝐸𝑐

𝑐 𝑦 𝑑 𝑑𝑇
𝑥

𝑋 𝑧 𝑧
𝑡

Figure 3.2: Cross-sectional area of the sandwich panel along section X-X [1]

Both face sheets have an equal thickness, t and an elastic modulus, Ef. The faces are
separated by the core of thickness, c and elastic modulus, Ec. The beam has a uniform
width, b. The total thickness is given by 𝑑 𝑇 , while the distance between the centroids of
the faces is given by 𝑑 [1].

In the sandwich beam theory, when the beam bends in the x-z plane, it is assumed that
the face sheets remain perfectly bonded to the core and that the cross-sectional area
remains plane and perpendicular to the y-axis [37]. Commonly, the flexural rigidity, D

18
is equal to the product of the elastic modulus, E and the moment of inertia, I. However,
in a sandwich panel the elastic modulus varies along the z-axis and so, the equivalent
flexural rigidity of the sandwich panel is given by the sum of the individual flexural
rigidities of the two face sheets and of the core when bent about the centroidal axis of
the entire cross-section. Starting from the general equation for the flexural rigidity:

𝐷 = ∬ 𝐸𝑧 2 𝑑𝐴
𝐴

𝐷 = 2𝐸𝑓 𝐼𝑓 + 𝐸𝑐 𝐼𝑐

With reference to Figure 3.2, the flexural stiffness of a single-layered sandwich panel is
given by:
𝐸𝑓 𝑏𝑡 3 𝐸𝑓 𝑏𝑡𝑑 2 𝐸𝑐 𝑏𝑐 3 (3.7)
𝐷= + +
6 2 12
With reference to Figure 3.3, the flexural stiffness of a multi-layered sandwich panel is
given by:
2
𝐸𝑓 𝑏𝑡 3 𝐸𝑓 𝑏𝑡𝑑 2 𝐸𝑐 𝑏𝑐𝑀
3 𝐸𝑐 𝑏𝑐𝑀 𝑑𝐶 (3.8)
𝐷= + + +
4 2 6 2

𝑡
𝐸𝑓
𝑐𝑀 𝐸𝑐
𝐸𝑓 𝑑 𝑑𝑇
𝑐 𝑡 𝑑𝐶
𝑐𝑀 𝐸𝑐
𝐸𝑓
𝑡
𝑏
Figure 3.3: Cross-sectional area of a multiple layer sandwich panel

Equations (3.7) and (3.8) make use of the parallel axis theorem to give the equivalent
flexural rigidity of the sandwich panel about the centroidal axis. Considering the
single-layered sandwich panel, the first term of equation (3.7) represents the flexural
rigidity of the face sheets when bending about their individual neutral axis, while the
second term represents the stiffness of the face sheets when bending about the centroidal
axis of the entire sandwich. The third term is the flexural rigidity of the core. The neutral
axis of the core coincides with the centroidal axis of the panel, provided that the face
sheets are of the same material and are of equal thicknesses [1].

19
The faces of a sandwich panel are normally thin compared to the core i.e. 𝑡 ≪ 𝑐.
Provided that equation (3.9) applies, the first term of equation (3.7) is equal to less than
1 % of the second term [1].

𝑑 (3.9)
> 5.77
𝑡
Furthermore, the elastic modulus of the core is usually much lower than that of the faces
i.e. 𝐸𝑐 ≪ 𝐸𝑓 . Provided that equation (3.10) applies, the third term of equation (3.7) is
also equal to less than 1 % of the second term [1].

6𝐸𝑓 𝑡𝑑 2 (3.10)
> 100
𝐸𝑐 𝑐 3
Thus, equation (3.7) for the flexural rigidity of a single-layered sandwich panel can be
reduced to:
𝐸𝑓 𝑏𝑡𝑑 2 (3.11)
𝐷≈
2
3.3.2 Stresses in Sandwich Panels

Consider the general expression for the strain in the 𝑥-direction, 𝜀𝑥 [1]:

𝑀𝑥 𝑧 (3.12)
𝜀𝑥 =
𝐷
where 𝑀𝑥 is the applied constant bending moment. The strain varies linearly with z
across the cross-section. Considering Figure 3.2 and applying equation (3.12) to find the
stresses in the faces, 𝜎𝑓 and in the core, 𝜎𝑐 :
𝑀𝑥 𝐸𝑓 𝑧 𝑐 𝑐 (3.13)
𝜎𝑓 = 𝐸𝑓 𝜀𝑥 = 𝑓𝑜𝑟 < |𝑧| < ( + 𝑡)
𝐷 2 2

𝑀𝑥 𝐸𝑐 𝑧 𝑐 (3.14)
𝜎𝑐 = 𝐸𝑐 𝜀𝑥 = 𝑓𝑜𝑟 |𝑧| <
𝐷 2

The general equation for the shear stress, 𝜏 is given by [1]:

𝑇𝑥 𝐵(𝑧) (3.15)
𝜏=
𝐷
where 𝑇𝑥 is the transverse shear force and 𝐵(𝑧) is the first moment of area, given by [1]:

𝑑+𝑡
2
𝐵(𝑧) = ∫ 𝐸𝑧 𝑑𝑧
𝑧

20
𝑐
For |𝑧| < 2, the first moment of area of the core as a function of z, 𝐵𝑐 (𝑧) is:

𝐸𝑓 𝑡𝑑 𝐸𝑐 𝑐 𝑐
𝐵𝑐 (𝑧) = + ( − 𝑧) ( + 𝑧)
2 2 2 2
Substituting this into equation (3.15) gives the shear stress variation in the core, 𝜏𝑐 (𝑧):

𝑇𝑥 𝐸𝑓 𝑡𝑑 𝐸𝑐 𝑐 2 (3.16)
𝜏𝑐 (𝑧) = [ + ( − 𝑧 2 )]
𝐷 2 2 4
𝑐 𝑐
For 2 < |𝑧| < (2 + 𝑡) the first moment of area of the faces as a function of z, 𝐵𝑓 (𝑧) is:

𝐸𝑓 𝑐 𝑐
𝐵𝑓 (𝑧) = ( + 𝑡 − 𝑧) ( + 𝑡 + 𝑧)
2 2 2

Substituting this into equation (3.15) gives the shear stress variation in the faces, 𝜏𝑓 (𝑧):

𝑇𝑥 𝐸𝑓 𝑐 2 (3.17)
𝜏𝑓 (𝑧) = ( + 𝑡𝑐 + 𝑡 2 − 𝑧 2 )
2𝐷 4
𝑐
At |𝑧| = 2, equations (3.16) and (3.17) give the shear stress at the core-facing interface:

𝑇𝑥 𝐸𝑓 𝑡𝑑 (3.18)
𝜏𝑐,𝑚𝑖𝑛 = 𝜏𝑓,𝑚𝑎𝑥 = ( )
𝐷 2
The maximum shear stress in the sandwich panel occurs at the neutral axis i.e. at 𝑧 = 0.
Hence, from equation (3.16):

𝑇𝑥 𝐸𝑓 𝑡𝑑 𝐸𝑐 𝑐 2 (3.19)
𝜏𝑐,𝑚𝑎𝑥 = ( + )
𝐷 2 8
For 𝐸𝑐 ≪ 𝐸𝑓 and 𝑡 ≪ 𝑐, the above equations can be reduced to a simpler form.
Equations (3.13) and (3.14) can be approximated by:
𝑀𝑥 (3.20)
𝜎𝑓 =
𝑏𝑑𝑡
𝜎𝑐 = 0 (3.21)

Furthermore, the second term of equation (3.19) may be ignored as it offers only a slight
contribution (less than 1 %) to the final result. Thus, equation (3.19) becomes equivalent
to equation (3.18) and therefore, through this approximation, the core shear stress can
be assumed constant across the whole core thickness. Thus, for 𝐸𝑐 ≪ 𝐸𝑓 and 𝑡 ≪ 𝑐, the
core shear stress, 𝜏𝑐 is given by:
𝑇𝑥 (3.22)
𝜏𝑐 =
𝑏𝑑

21
From equation (3.17), it is evident that the shear stress across the face sheets varies from
zero at the free surface to a maximum, 𝜏𝑓,𝑚𝑎𝑥 at the core-facing interface. Provided that
𝑡 ≪ 𝑐, the stress at the face sheets can be ignored and equation (3.17) is reduced to:
𝜏𝑓 = 0 (3.23)
The results of equations (3.20) to (3.23) prove that in sandwich constructions, if 𝐸𝑐 ≪ 𝐸𝑓
and 𝑡 ≪ 𝑐, the faces carry the bending moment as tensile and compressive stresses, while
the core carries transverse forces as shear stresses, as depicted in Figure 3.4 [1].

face

core

face
𝜎𝑥 𝜏𝑥
Figure 3.4: Direct and shear stresses acting across the cross-sectional area of the sandwich
panel provided that 𝐸𝑐 ≪ 𝐸𝑓 and 𝑡 ≪ 𝑐 [1]

3.3.3 Transverse Shear Rigidity

The transverse shear rigidity, U of a general cross-section is found by equating the


potential energy of the applied loads to the strain energy of the system [1]:

1 1 (3.24)
𝑇𝑥 𝛾 = ∫ 𝜏𝑥𝑦 (𝑧). 𝛾𝑥𝑧 (𝑧) 𝑑𝑧
2 2
The transverse shear strain, 𝛾, is given by [1, 11]:

𝑇𝑥 𝑇𝑥 (3.25)
𝛾= =
𝑈 𝐺𝑐 𝑑
where 𝐺𝑐 is the shear modulus of the core. For a sandwich panel, the shear modulus
of the faces, 𝐺𝑓 is very large and provided that 𝐸𝑐 ≪ 𝐸𝑓 and 𝑡 ≪ 𝑐, substituting
equations (3.22) and (3.25) into equation (3.24) gives:
𝑐
1 1 2 𝑇𝑥 𝑇𝑥 (3.26)
𝑇𝑥 𝛾 = ∫ . 𝑑𝑧
2 2 −𝑐 𝑏𝑑 𝐺𝑐 𝑑
2

Integrating equation (3.26) gives:


1 𝑇𝑥2 𝑐 (3.27)
𝑇𝛾=
2 𝑥 2𝐺𝑐 𝑏𝑑 2

22
Substituting equation (3.25) into equation (3.27) and simplifying gives:
𝐺𝑐 𝑏𝑑 2 (3.28)
𝑈=
𝑐

3.4 Long Beam Flexure Test

The long beam flexure test was designed in accordance with ASTM D7249 [30]. The
test allows for the determination of the facing properties of a flat sandwich construction
subjected to flexure in such a manner that the applied moments produce curvature of the
sandwich facing planes, resulting in compressive and tensile forces in the facings [30].
The standard 4-point loading configuration is shown in Figure 3.5. The total deflection
of the beam, 𝑤 has two components; the deflection due to bending moments, 𝑤𝑏 and the
deflection due to shear forces, 𝑤𝑠 :

𝑤 = 𝑤𝑏 + 𝑤𝑠 (3.29)

Figure 3.5: Loading configuration for 4-point bending as given in ASTM D7249 [30]

3.4.1 Deflection due to bending moments

Different beam sections, as shown in Figure 3.6, will be considered to find the transverse
shear forces, 𝑇𝑥 and bending moments, 𝑀𝑥 acting along the length of the beam.

𝑃 𝑃
1 2 2 2 3
𝑥

1 2 3 𝑃
𝑃
2
2

Figure 3.6: Sections taken along the length of the beam

23
𝑆−𝐿
Section 1-1, for 0 < 𝑥1 < : 𝑃
2 𝑇1 =
2
𝑥1 𝑃
𝑀1 = 𝑥
𝑀1 2 1
𝑃
𝑎𝑡 𝑥1 = 0; 𝑇1 = ,𝑀 = 0
2 1
𝑃 𝑇1
𝑆−𝐿 𝑃 𝑃
2 𝑎𝑡 𝑥1 = ; 𝑇1 = , 𝑀1 = (𝑆 − 𝐿)
2 2 4

𝑆−𝐿 𝑆+𝐿
Section 2-2, for < 𝑥2 < :
2 2
𝑇2 = 0
𝑃 𝑃
𝑥2 𝑀2 = (𝑆 − 𝐿)
2 4
𝑀2 𝑆−𝐿 𝑃
𝑎𝑡 𝑥2 = ; 𝑇2 = 0, 𝑀2 = (𝑆 − 𝐿)
2 4
𝑆+𝐿 𝑃
𝑃 𝑇2 𝑎𝑡 𝑥2 = ; 𝑇2 = 0, 𝑀2 = (𝑆 − 𝐿)
2 4
2

𝑆+𝐿
Section 3-3, for < 𝑥3 < 𝑆: 𝑃
2 𝑇3 = −
𝑃 2
𝑃
2 𝑃
𝑥3 2 𝑀3 = (𝑆 − 𝑥3 )
2
𝑀3 𝑆+𝐿 𝑃 𝑃
𝑎𝑡 𝑥3 = ; 𝑇3 = − , 𝑀3 = (𝑆 − 𝐿)
2 2 4
𝑇3 𝑃
𝑃 𝑎𝑡 𝑥3 = 𝑆; 𝑇3 = − , 𝑀3 = 0
2
2

The results are used to plot the shear and bending moment diagrams, Figure 3.7.

𝑃ൗ
2

𝑥
− 𝑃ൗ2
𝑀
𝑃
(𝑆 − 𝐿)
4

0 (𝑆 − 𝐿)⁄2 (𝑆 + 𝐿)⁄2 𝐿
𝑥

Figure 3.7: Shear and moment diagrams for a beam under 4-point bending

24
The shape of the deflected beam can be approximated by equation (3.30), assuming
elastic bending [38]:
𝑑2 𝑤𝑏 𝑀𝑥 (3.30)
2
=−
𝑑𝑥 𝐸𝐼
𝑆−𝐿
For 0 < 𝑥1 < , substituting for 𝑀1 in equation (3.30):
2

𝑑2 𝑤𝑏1 𝑃𝑥1
2 = − 2𝐸𝐼
𝑑𝑥1
𝑑 𝑤𝑏1 𝑃𝑥12 (3.31)
=− + 𝐶1
𝑑𝑥1 4𝐸𝐼
𝑃𝑥13 (3.32)
𝑤𝑏1 = − + 𝐶1 𝑥1 + 𝐶2
12𝐸𝐼
At 𝑥1 = 0, 𝑤𝑏1 = 0 and so, from equation (3.32), 𝐶2 = 0

𝑆−𝐿 𝑆+𝐿
For < 𝑥2 < , substituting for 𝑀2 in equation (3.30):
2 2

𝑑2 𝑤𝑏2 𝑃(𝑆 − 𝐿)
2 =− 4𝐸𝐼
𝑑𝑥2
𝑑 𝑤𝑏2 𝑃(𝑆 − 𝐿)𝑥2 (3.33)
=− + 𝐶3
𝑑𝑥2 4𝐸𝐼
𝑃(𝑆 − 𝐿)𝑥22 (3.34)
𝑤𝑏2 = − + 𝐶3 𝑥2 + 𝐶4
8𝐸𝐼
𝑆 𝑑 𝑤𝑏2
At 𝑥2 = 2 , = 0 and so, from equation (3.33):
𝑑𝑥2

𝑃𝑆
𝐶3 = (𝑆 − 𝐿)
8𝐸𝐼
𝑆−𝐿 𝑑 𝑤𝑏1 𝑑 𝑤𝑏2
At 𝑥1 = 𝑥2 = , = . Equating (3.31) and (3.33) and substituting for 𝐶3 :
2 𝑑𝑥1 𝑑𝑥2

𝑃(𝑆 − 𝐿)2 𝑃(𝑆 − 𝐿)2 𝑃𝑆(𝑆 − 𝐿)


− + 𝐶1 = − +
16𝐸𝐼 8𝐸𝐼 8𝐸𝐼

𝑃
𝐶1 = (𝑆 − 𝐿)(𝑆 + 𝐿)
16𝐸𝐼
𝑆−𝐿
At 𝑥1 = 𝑥2 = , 𝑤𝑏1 = 𝑤𝑏2. Equating equations (3.32) and (3.34) and substituting
2

for 𝐶1 , 𝐶2 and 𝐶3 :

𝑃(𝑆 − 𝐿)3 𝑃 𝑃(𝑆 − 𝐿)3 𝑃𝑆(𝑆 − 𝐿)2


− + (𝑆 − 𝐿)2 (𝑆 + 𝐿) = − + + 𝐶4
96𝐸𝐼 32𝐸𝐼 32𝐸𝐼 16𝐸𝐼

𝑃
𝐶4 = − (𝑆 − 𝐿)3
96𝐸𝐼

25
The largest bending moment occurs at the mid-span. Substituting 𝑥2 , 𝐶3 , 𝐶4 and
𝐷 = 𝐸𝐼, in equation (3.34), the maximum theoretical deflection due to bending
moments for a beam loaded in 4-point bending is:
𝑃 (3.35)
𝑤𝑏,𝑚𝑎𝑥 = (2𝑆 3 − 3𝑆𝐿2 + 𝐿3 )
96𝐷
3.4.2 Deflection due to shear forces

The deflection due to the transverse shear forces, 𝑤𝑠 is given by [1]:


𝑀𝑥 𝛾0 𝑐𝑥 (3.36)
𝑤𝑠 = − + 𝐶5
𝑈 𝑑
where 𝛾0, the in-plane core shear strain and 𝐶5 are constants, fixed for each loading
configuration. The unknown constants can be found by applying boundary conditions.

𝑆−𝐿
For 0 < 𝑥1 < , substituting 𝑀1 in equation (3.36):
2

𝑃𝑥1 𝛾0 𝑐𝑥1
𝑤𝑠 = − + 𝐶5
2𝑈 𝑑

At 𝑥1 = 0, 𝑤𝑠 = 0 giving 𝐶5 = 0

𝑆+𝐿
For < 𝑥3 < 𝑆, substituting 𝑀3 in equation (3.36) gives:
2

𝑃 𝛾0 𝑐𝑥3
𝑤𝑠 = (𝑆 − 𝑥3 ) −
2𝑈 𝑑

At 𝑥3 = 𝑆, 𝑤𝑠 = 0 giving 𝛾0 = 0

𝑆
Maximum deflection takes place at the mid-span (𝑥2 = 2). Substituting for 𝑥2 , 𝛾0 and

𝐶5 in equation (3.36) shows that the maximum deflection due to shear forces for a beam
in 4-point bending is:
𝑃 (3.37)
𝑤𝑠,𝑚𝑎𝑥 = (𝑆 − 𝐿)
4𝑈
3.4.3 Total Deflection

The total maximum theoretical deflection for 4-point bending, 𝑤𝑚𝑎𝑥 can be found by
substituting equations (3.35) and (3.37) into equation (3.29):

𝑤𝑚𝑎𝑥 = 𝑤𝑏,𝑚𝑎𝑥 + 𝑤𝑠,𝑚𝑎𝑥


𝑃 𝑃 (3.38)
𝑤𝑚𝑎𝑥 = (2𝑆 3 − 3𝑆𝐿2 + 𝐿3 ) + (𝑆 − 𝐿)
96𝐷 4𝑈

26
3.4.4 Calculation of Facing Properties

3.4.4.1 Facing Ultimate Stress

In this section and in the following sections, some of the symbols used in the ASTM
standards were changed for coherency throughout the study. From ASTM D7249 [30],
the facing ultimate stress, 𝜎𝑓,𝑢𝑙𝑡 can be calculated by:

𝑃𝑚𝑎𝑥 (𝑆 − 𝐿) (3.39)
𝜎𝑓,𝑢𝑙𝑡 =
2𝑏𝑡(𝑑 𝑇 + 𝑐)

provided that the thickness of the facings is small in comparison to the core thickness.

3.4.4.2 Effective Facing Chord Modulus

The modulus of each face sheet, 𝐸𝑓 can be calculated using equation (3.40):

σ3000 − σ1000 (3.40)


𝐸𝑓 =
ε3000 − ε1000

where σ is the facing stress corresponding to the recorded strain value ε, closest to
1000 με and 3000 με, respectively. The modulus has to be calculated separately for each
facing; compressive for the top facing and tensile for the bottom facing.

3.4.4.3 Sandwich Flexural Stiffness

The sandwich flexural stiffness, 𝐷 can be calculated using equation (3.41):

(𝑆 − 𝐿) 𝑑𝑇 𝑃3000 − 𝑃1000 (3.41)


𝐷 = ( )
4 (𝜀𝑡,3000 − 𝜀𝑡,1000 ) + (𝜀𝑏,3000 − 𝜀𝑏,1000 )

where subscripts 𝑡 and 𝑏 represent the top and bottom facings, respectively [30].
3.4.5 Theoretical Failure Modes

The only acceptable failure modes for the long beam flexure test, are those that are
internal to one of the facings, namely face yielding and face wrinkling. Failure due to
indentation beneath the loading rollers is also possible, however, this is not a desirable
mode of failure and rubber pressure pads are placed between the loading bars and the
test specimen to avoid any local damage to the faces. Core shear is also a possible mode
of failure, however, this is also not desirable. The specimen geometry specified by
ASTM D7249 [30] promotes failure that is internal to one of the facings [30, 39].

27
3.4.5.1 Face Yielding Failure

Face yielding, depicted in Figure 3.8, occurs when the axial stress in the face sheet in
compression exceeds the yield strength of the facings, 𝜎𝑓𝑦 [39]. From equations (3.20)
and (3.21), the contribution of the core to the overall bending strength of the sandwich
construction may be neglected and it may be assumed that the facings carry the total
bending moment applied. The load at which face yielding failure occurs, 𝑃𝑓𝑦 can thus
be found by equating the maximum moment being experienced by the beam, 𝑀𝑥,𝑚𝑎𝑥 , to
the maximum moment that the beam can withstand before yielding, 𝑀𝑦𝑖𝑒𝑙𝑑 [40]:

𝑀𝑥,𝑚𝑎𝑥 = 𝑀𝑦𝑖𝑒𝑙𝑑

𝑃𝑓𝑦 (𝑆 − 𝐿)
= 𝜎𝑓𝑦 𝑏𝑑𝑡
4

4𝜎𝑓𝑦 𝑏𝑑𝑡 (3.42)


𝑃𝑓𝑦 =
𝑆−𝐿

Figure 3.8: Face yielding failure [39]

3.4.5.2 Face Wrinkling Failure

Face wrinkling is a form of local instability that involves short-wavelength buckling of


the face sheet in compression, as shown in Figure 3.9 [41]. It occurs when the stress in
the face sheet exceeds the critical wrinkling strength of the facing, 𝜎𝑓𝑤 [39]. The critical
wrinkling strength is given by the Hoff and Maunter formula [42]:

3 (3.43)
𝜎𝑓𝑤 = 0.5 √𝐸𝑓 𝐸𝑐 𝐺𝑐

Figure 3.9: Different modes of face wrinkling failure [39]

The load at which face wrinkling occurs, 𝑃𝑓𝑤 can be found by equating the Hoff and
Maunter formula to the yield strength of the facings, 𝜎𝑓𝑦 :

28
𝜎𝑓𝑦 = 𝜎𝑓𝑤
𝑃𝑓𝑤 (𝑆 − 𝐿) 3
= 0.5 √𝐸𝑓 𝐸𝑐 𝐺𝑐
4𝑏𝑑𝑡

2𝑏𝑑𝑡 3 (3.44)
𝑃𝑓𝑤 = √𝐸𝑓 𝐸𝑐 𝐺𝑐
(𝑆 − 𝐿)

3.5 Short Beam Shear Test

The short beam shear test is designed in accordance with ASTM C393 [31]. The test
allows for the determination of the core shear properties of flat sandwich panels
subjected to flexure in a way that the applied moments produce curvature of the
sandwich panel facings. The standard 3-point loading configuration is shown in
Figure 3.10.

𝑃
𝑆ൗ 𝑆ൗ
2 2

𝑃ൗ 𝑆 𝑃ൗ
2 2

Figure 3.10: Loading configuration for 3-point bending as given in ASTM C393 [31]

Repeating the procedure followed for the 4-point bending configuration, the two
components making up the total deflection will be considered separately.

3.5.1 Deflection due to bending moments

One beam section is sufficient to produce the shear and bending moment diagrams for
this configuration.

𝑆
For 0 < 𝑥 < 2:
𝑃 𝑃
𝑇= 𝑎𝑛𝑑 𝑀 = 𝑥
𝑥 2 2
𝑀 𝑃
𝑎𝑡 𝑥 = 0; 𝑇1 = ,𝑀 = 0
2

𝑇 𝑆 𝑃 𝑃𝑆
𝑃 𝑎𝑡 𝑥 = ; 𝑇1 = , 𝑀 =
2 2 4
2

29
The loading is symmetric, giving the shear and bending moments of Figure 3.11.

𝑃ൗ
2

𝑥
− 𝑃ൗ2

𝑀
𝑃𝑆ൗ
4

0 𝑥
𝑆 ⁄2 𝐿
Figure 3.11: Shear and moment diagrams for a beam under 3-point bending

Approximating the shape of the deflected beam by equation (3.30), substituting for 𝑀:

𝑑 2 𝑤𝑏 𝑀 𝑃𝑥
2
=− =−
𝑑𝑥 𝐸𝐼 2𝐸𝐼
𝑑 𝑤𝑏 𝑃𝑥 2 (3.45)
=− + 𝐶6
𝑑𝑥 4𝐸𝐼
𝑃𝑥 3 (3.46)
𝑤𝑏 = − + 𝐶6 𝑥 + 𝐶7
12𝐸𝐼
Applying boundary conditions gives:

𝑃𝑆 2 𝐶7 = 0
𝐶6 =
16𝐸𝐼
Maximum deflection due to the bending moments, 𝑤𝑏,𝑚𝑎𝑥 takes place at mid-span.
𝑆
Substituting 𝑥 = 2 , 𝐶6, 𝐶7 and 𝐷 = 𝐸𝐼 into equation (3.46) gives:
𝑃𝑆 3 (3.47)
𝑤𝑏,𝑚𝑎𝑥 =
48𝐷

3.5.2 Deflection due to shear forces

Applying equation (3.36) to find the deflection due to the transverse shear forces:
𝑀𝑥 𝛾0 𝑐𝑥 (3.48)
𝑤𝑠 = − + 𝐶8
𝑈 𝑑
Applying the boundary conditions, 𝑤𝑠 = 0 at 𝑥1 = 0 and at 𝑥3 = 𝑆, gives:
𝛾0 = 𝐶8 = 0

30
𝑆
Maximum deflection takes place at the mid-span. Substituting 𝑀𝑥 at 𝑥 = 2 , 𝛾0 and 𝐶8

into equation (3.48) gives:


𝑃𝑆 (3.49)
𝑤𝑠,𝑚𝑎𝑥 =
4𝑈

3.5.3 Total Deflection

The theoretical total maximum deflection, 𝑤𝑚𝑎𝑥 can be found by substituting equations
(3.47) and (3.49) into equation (3.29):

𝑤𝑚𝑎𝑥 = 𝑤𝑏,𝑚𝑎𝑥 + 𝑤𝑠,𝑚𝑎𝑥

𝑃𝑆 3 𝑃𝑆 (3.50)
𝑤𝑚𝑎𝑥 = +
48𝐷 4𝑈

3.5.4 Calculation of the Core Shear Properties

3.5.4.1 Core Shear Strength

Following ASTM C393 [31], the core shear strength, 𝜏𝐶𝑆 can be calculated by:

𝑃𝑚𝑎𝑥 (3.51)
𝜏𝐶𝑆 =
(𝑑 𝑇 + 𝑐)𝑏

3.5.4.2 Facing Stress

The facing stress, 𝜎𝑓 can be calculated by the equation:


𝑃𝑚𝑎𝑥 𝑆 (3.52)
𝜎𝑓 =
2𝑡 (𝑑𝑇 + 𝑐)𝑏

3.5.5 Theoretical Failure Modes

According to ASTM C393 [31], the only acceptable failure modes for this test are core
shear failure or core-to-facing bond failure. Indentation beneath the rollers is also
possible however, rubber pressure pads are placed between the loading bars and the test
specimen to avoid damaging the facings [31].

3.5.5.1 Core Shear Failure

Core shear failure, shown in Figure 3.12, occurs when the applied load produces a
transverse shear force that is greater than the shear strength of the core. The load at

31
which core shear failure occurs, 𝑃𝐶𝑆 can be found by equating the maximum transverse
shear stress in the core, 𝜏𝑐,𝑚𝑎𝑥 , to the core shear strength, 𝜏𝐶𝑆 [39]. From Figure 3.11:

𝑇𝑥,𝑚𝑎𝑥 𝑃𝐶𝑆 1
𝜏𝑐,𝑚𝑎𝑥 = = .
𝐴 2 𝑏𝑑

𝑃𝐶𝑆
𝜏𝑐,𝑚𝑎𝑥 = = 𝜏𝐶𝑆
2𝑏𝑑

𝑃𝐶𝑆 = 2𝑏𝑑𝜏𝐶𝑆 (3.53)

Figure 3.12: Core shear failure [39]

3.6 Flexural Stiffness, Transverse Shear Rigidity and Core Shear Modulus

ASTM D7250 [43] uses results obtained from both ASTM D7249 [30] and ASTM C393
[31] to determine the sandwich panels’ flexural and shear stiffness. The facing modulus
of the sandwich panels, 𝐸𝑓 can be found from ASTM D7249 [30] as explained in section
3.4.4.2. Provided that the facings are identical, ASTM D7250 [43] can then be applied
to find the sandwich flexural stiffness 𝐷, using the equation:

𝐸𝑓 (𝑑 3𝑇 − 𝑐 3 )𝑏 (3.54)
𝐷=
12

Following, using the applied loads, 𝑃 and the corresponding mid-span deflections, 𝑤𝑚𝑎𝑥
of the short beams loaded under 3-point loading according to ASTM C393 [31], the
shear rigidity 𝑈, can be found by equation (3.55):

𝑃𝑆 (3.55)
𝑈=
2𝑃𝑆 3
4 (𝑤𝑚𝑎𝑥 − 96𝐷 )

Finally, the core shear modulus 𝐺𝑐 , can be calculated by equation (3.56) [43]:

𝑈 (𝑑 𝑇 − 2𝑡) (3.56)
𝐺𝑐 =
(𝑑 𝑇 − 𝑡)2 𝑏

32
3.7 Edgewise Compressive Strength Test

The edgewise compressive strength test is designed in accordance with ASTM C364
[33]. The test allows for the determination of the compressive properties of flat structural
sandwich constructions of unsupported length 𝐿, in a direction parallel to the sandwich
facing plane [33]. The edgewise compressive strength of a short sandwich panel
provides a basis for judging the load-carrying capacity of the faces [44].

Figure 3.13: Loading configuration for the edgewise compressive strength test

3.7.1 Calculation of the Facing Compressive Stress

Consider the loading configuration shown in Figure 3.13. As shown in section 3.3.2, it
can be assumed that the faces of a sandwich column carry the applied in-plane loads and
bending moments, provided that 𝐸𝑐 ≪ 𝐸𝑓 and 𝑡 ≪ 𝑐, while the core resists the shear and
buckling forces [44]. From ASTM C364 [33], the ultimate load carried by the sandwich
column can be used to calculate the facing compressive stress, 𝜎𝑓,𝑐 , by applying the
following equation:
𝑃 𝑃 (3.57)
𝜎𝑓,𝑐 = =
𝐴𝑓 𝑛𝑏𝑡

where 𝐴𝑓 is the area of the face sheets and 𝑛 is the number of facings.

3.7.2 Theoretical Failure Modes

Various failure modes are possible for the edgewise compressive strength test, namely
general buckling, face sheet yielding and face wrinkling [39].

33
3.7.2.1 General Buckling

General buckling consists of two types of buckling that interact together at transition
values of the strut slenderness ratio, resulting in a combined failure mode. The two types
of buckling are Euler buckling, involving bending of the sandwich column, and shear
buckling, involving core shear. The two types of buckling are depicted in Figures 3.14a
and 3.14b.

Figure 3.14a: Euler buckling [39] Figure 3.14b: Shear buckling [39]

The combined collapse load for general buckling, 𝑃𝐺𝐵 is given by:

1 1 1 (3.58)
= +
𝑃𝐺𝐵 𝑃𝐸𝐵 𝑃𝐶𝑆𝐵
where 𝑃𝐸𝐵 , the Euler buckling load is given by equation:

4𝜋 2 𝐷 (3.59)
𝑃𝐸𝐵 = 2
𝐿
and 𝑃𝐶𝑆𝐵 , the core shear buckling load, is approximately equal to the shear rigidity of
the core [45]:
𝐺𝑐 𝑏𝑑2 (3.60)
𝑃𝐶𝑆𝐵 = 𝑈 = ≈ 𝐺𝑐 𝑏𝑐
𝑐
3.7.2.2 Face Sheet Yielding

Face sheet yielding, also referred to as face sheet microbuckling, occurs when the
compressive stress in the face sheet exceeds the compressive yield strength of the face,
𝜎𝑓𝑦 . The face sheet yielding failure mode is shown in Figure 3.15a. The critical load at
which face sheet yielding occurs in edgewise compressive testing, 𝑃𝑓𝑦 is given by [39]:

𝑃𝑓𝑦 = 𝑛𝑏𝑡𝜎𝑓𝑦 (3.61)

34
3.7.2.3 Face Wrinkling

As explained in section 3.4.5.2, face wrinkling is a form of local instability involving


short wavelength elastic buckling. The face wrinkling failure mode is shown in
Figure 3.15b. The critical load at which face wrinkling failure, 𝑃𝑓𝑤 occurs can be found
by equating the critical face wrinkling stress 𝜎𝑓𝑤 , given by the Hoff Maunter formula
(equation (3.43)), to ultimate the stress developed in the face sheets, 𝜎𝑓,𝑢𝑙𝑡 given by
equation (3.57):

𝜎𝑓,𝑢𝑙𝑡 = 𝜎𝑓𝑤

𝑃𝑓𝑤 3
= 0.5 √𝐸𝑓 𝐸𝑐 𝐺𝑐
𝑛𝑏𝑡
3 (3.62)
𝑃𝑓𝑤 = 0.5 𝑛𝑏𝑡 √𝐸𝑓 𝐸𝑐 𝐺𝑐

Figure 3.15a: Face sheet yielding [39] Figure 3.15b: Face wrinkling [39]

3.8 Conclusion

The equations in this chapter, either derived from first-principles or obtained from
ASTM standards, will be used for determining the facing and core shear properties of
the sandwich panels. The same tests also allow for the calculation of overall sandwich
properties, such as the flexural stiffness and shear rigidity. The edgewise compression
behaviour will be used as a basis for judging the load-carrying capacity of the faces.
Different sandwich layups will also be analysed, allowing for a comparison between the
properties of the single and multi-layered sandwich constructions.

35
4 Sandwich Panel Construction and Testing

4.1 Overview

The study will be based on sandwich panels having flax facings and a cork core, as a
greener alternative to traditional materials used in the maritime industry. Single and
multiple layer sandwich panels were fabricated and tested. The following chapter
presents the properties of the sandwich panel components, gives details on the sandwich
panel fabrication process and explains the ASTM standard testing procedures applied.

4.2 Sandwich Panel Materials

Materials such as glass fibre and carbon fibre reinforced plastics are common in the boat
building industry, mainly in yacht hull structures. These materials are expensive and
have energy-intensive production processes [46]. A shift is being made towards
materials that can be extracted and processed with low carbon dioxide emissions,
providing eco-friendly alternatives, while still being structurally viable [47]. The aim of
this study is to analyse the properties and suitability of greener alternatives for the
maritime industry. All materials used in this study are thus natural and bio-based.

The fabric used in the facings is the 2/2 twill Biotex® flax fibre fabric. The woven flax
fabric has superior mechanical properties when compared to other natural alternatives,
such as jute, and can be processed using methods applied to carbon and glass fibres [48].
When compared to glass fibre, flax generates less pollution during production, resulting
in minimal health hazards [49]. Flax fibres have already been used in sandwich panels
in boat building applications, exhibiting an adequate stiffness, vibration damping and
impact and abrasion-resistant properties [47]. The properties of the flax fibres used in
this study are given in Table 4.1. As flax is a natural material, slight variations from the
values given in the table are expected.

Density, 𝜌𝑓𝑏 (kg/m3) 1500

Tensile modulus, 𝐸𝑓𝑏 (GPa) 50

Tensile strength (MPa) 500

Table 4.1: Biotex® flax fibre properties [48]

36
The core is made of Amorim® CoreCork NL20 cork. Cork is a lightweight, sustainable
material offering vibration and thermal insulation. Another property of cork is its low
water absorption, making it ideal for maritime applications [50]. Studies by Castro et al.
have shown that cork offers a higher shear strength, in comparison to other core
materials, limiting the region of crack propagation. Furthermore, sandwich panels with
cork cores have a high energy absorption capacity, displaying better crashworthiness
properties under impact loading; an important consideration for maritime applications
[22]. Properties of the cork core are given in Table 4.2.

Density, 𝜌𝑐 (kg/m3) 170 – 235

Modulus, 𝐸𝑐 (MPa) 6.0

Shear modulus, 𝐺𝑐 (MPa) 5.9

Shear strength, 𝜏𝐶𝑆 (MPa) 0.9

Table 4.2 Amorim® CoreCork NL20 cork properties [50, 65]

The resin used in this study is the ONE high bio-based laminating epoxy, a
general-purpose, renewable, liquid epoxy resin with a bio-based content of 30 %. ONE
is the industry leader for USDA BioPreferredSM certified epoxy systems [51]. It is a
Super Sap® formula, employing green chemistry techniques that require less energy.
Greenhouse gas emissions from the production of these epoxies are reduced by 33 %
over conventional petroleum-based epoxies. The raw materials used in the production
of these epoxy resins are co-products or waste products of other industrial processes,
such as those from a bio-refinery [52]. Properties of the resin are given in Table 4.3.

Density, 𝜌𝑚 (kg/m3) 1090

Tensile modulus, 𝐸𝑚 (GPa) 2.7

Tensile strength (MPa) 53.2

Table 4.3 ONE Super Sap® epoxy system properties [51]

37
4.3 Sandwich Panel Design

The panels produced were sized according to the number of required test specimens and
the required geometry, as specified in the corresponding standards. Furthermore, the
standard configuration of ASTM C393 [31] is 3-point loading, while that of ASTM
D7249 [30] is 4-point loading. For edgewise compression testing, ASTM C364 [33]
states that the width of the specimens shall be at least 50 mm but not less than twice the
total thickness. The unsupported length shall be not greater than eight times the total
thickness [33]. Table 4.4 lists the standards applied for each test, along with the
specimen dimensions. For all tests, the specimen is rectangular in cross-section, with the
thickness of the specimen equal to the thickness of the sandwich construction.

Standard Specimen length (mm) Specimen width (mm)

ASTM D7249 [30]


600 75
(standard configuration)

ASTM C393 [31]


200 75
(standard configuration)

ASTM C364 [33] 75 50

Table 4.4: Test specimen dimensions [30, 31, 33]

Single layup sequence:


3 layers 2/2 twill Biotex® flax fibre fabric
8 mm
8 mm Amorim® CoreCork NL20
3 layers 2/2 twill Biotex® flax fibre fabric

Figure 4.1a: Single layup (not to scale)


Multiple layup sequence:
2 layers 2/2 twill Biotex® flax fibre fabric
4 mm
4 mm Amorim® CoreCork NL20
2 layers 2/2 twill Biotex® flax fibre fabric
4 mm
4 mm Amorim® CoreCork NL20
2 layers 2/2 twill Biotex® flax fibre fabric
Figure 4.1b: Multiple layup (not to scale)

38
For each test, both single and multiple layer sandwich panels of the same overall
thickness were constructed, in order to analyse the effect of the layup sequence on the
overall sandwich properties. The layup sequences considered are shown in Figures 4.1a
and 4.1b. As can be seen from the figures, both layups have a total of six layers of flax
and a total of 8 mm of cork.

4.4 Sandwich Panel Fabrication

The sandwich panels were fabricated using a combination of hand layup and vacuum
bagging techniques. In vacuum bagging, a uniform pressure is applied to the sandwich
panel during the curing process. The pressure removes the air and excess resin, resulting
in an optimised fibre-to-resin ratio and a panel with a higher strength-to-weight ratio
[53]. Furthermore, the applied pressure consolidates the components making up the
sandwich panel, improving the efficiency of force transmission [54].

Figure 4.2: Preparation of cork and flax by drilling holes and tapering the edges

All panels were fabricated slightly larger than required to have some allowance for the
blade cutting thickness. The first step was to prepare the required materials. The flax and
cork were cut to the required size as shown in Figure 4.2. To allow for the resin to go
through the cork, 2 mm holes were drilled every 2 cm. The edge of the cork was tapered
at an angle of 45 º to allow for the bag film to tighten properly around the edges once

39
pressure is applied, preventing potential air pockets from forming. The mass of flax and
cork used in each panel was noted. The room temperature and relative humidity during
the fabrication of each panel were also noted.

Figure 4.3: Wetting the flax with resin

A flat melamine surface was used for the fabrication of the sandwich panels. The surface
was initially cleaned with a scraper to remove any residual resin from previous use. A
release wax was applied to the surface to ensure easy removal of the panel, post-curing.
The resin and hardener mixture was prepared following the mix ratio given in the
guidelines, and a layer of resin was applied to the waxed flat surface. The first layer of
flax was laid on top and resin was applied as shown in Figure 4.3, until the flax layer
was completely wetted. A roller was used over the flax layer to help distribute the resin
evenly throughout. The procedure was repeated for the subsequent layers of flax and
cork, as required for both the single and multiple layup sequences.

Figure 4.4: Vacuum bagging layup sequence

40
Once the flax and cork were all laid up, the peel-ply was applied on top of the panel.
The peel-ply aids in obtaining a smooth surface and facilitates demoulding. A perforated
film was then placed on top of the peel-ply, to allow trapped air and excess resin to be
removed [55]. The excess resin is absorbed by the breather, added on top of the
perforated film. A pump connector was placed on top of the breather and then, the panel
together with the peel-ply, perforated film and breather, were covered with a bag film.
The bag was sealed at the edges, using sealant tape to form an airtight seal. A clearance
border of a few centimetres was allowed between the panel and the sealant tape. The
vacuum bagging layup sequence described is depicted in Figure 4.4.

Figure 4.5: Two panels curing under pressure

The vacuum pump was connected to the pump connector by making a small hole in the
bag film. The pump was set to 50 kPa below atmospheric pressure and switched on,
making sure that there is no air release [56]. Figure 4.5 shows two of the sandwich
panels, curing under the pressure applied by the vacuum pump. The sandwich panels
were allowed to cure under such conditions for 6 hours, after which the vacuum pump
was switched off. The peel-ply, perforated film and breather were removed the following
day. The process was repeated until all the required panels were fabricated.

Once cured, the panels were cut into the required test specimen geometry, using a band
saw. The edges of the test specimens were smoothed out using a belt sander to remove
any blade marks and minor imperfections that resulted from cutting. Some of the single
and multiple layer test specimens are shown in Figure 4.6.

41
Figure 4.6: Single and multiple layer test specimens

A separate panel was manufactured for each of the three tests, for both single and
multi-layered sandwiches; a total of six panels. In the following sections, the single and
multi-layered specimens are labelled by the letters “S” and “M”, respectively. Table 4.5
gives the resulting fibre weight and volume fractions, along with the temperature and
relative humidity recorded during the fabrication process of each panel. The fibre weight
fractions were calculated by equation (3.6), while the fibre volume fractions were
calculated by equation (3.5).

Room Relative Fibre weight Fibre volume


Test Method Layup temperature humidity fraction fraction
(℃) (%) Eq. (3.6) Eq. (3.5)

ASTM S 18.9 60.4 0.251 0.196


D7249 [30] M 19.1 69.2 0.246 0.191
ASTM S 0.250 0.195
18.2 63.5
C393 [31] M 0.238 0.185
ASTM S 0.300 0.237
17.5 58.7
C364 [33] M 0.269 0.211

Table 4.5: Fibre weight and volume fractions of each panel

42
Table 4.6 gives the resulting lower and upper bounds for the theoretical facing modulus,
𝐸𝑓 , calculated by equations (3.3) and (3.1), respectively. The practical rule of mixtures
(equation (3.4)) was applied using α = 0.5 [1], to predict the theoretical facing modulus
resulting from the 2/2 twill flax fibre fabric used in this study. All equations used in the
prediction of the theoretical facing modulus make use of the fibre volume fraction, as
given in Table 4.5. The theoretical flexural rigidity, 𝐷 was then computed for the single
and multiple sandwich layups, by applying equations (3.7) and (3.8), respectively.

Lower bounds α = 0.5 [1] Upper bounds

Ef D Ef D Ef D
Test Method Layup (GPa) (Nm2) (GPa) (Nm2) (GPa) (Nm2)
(3.3) (3.7 & 3.8) (3.4) (3.7 & 3.8) (3.1) (3.7 & 3.8)

ASTM S 3.31 15.6 5.98 28.1 12.0 56.2


D7249 [30] M 3.30 12.8 5.88 22.7 11.8 45.4

ASTM S 3.31 15.6 5.96 28.0 11.9 56.0


C393 [31] M 3.27 12.7 5.72 22.1 11.4 44.2

ASTM S 3.48 16.4 6.97 32.7 13.9 65.5


C364 [33] M 3.37 13.1 6.33 24.5 12.7 48.9

Table 4.6: Theoretical facing modulus and flexural rigidity of the sandwich panels produced

The theoretical transverse shear rigidity, U was computed by equation (3.28), for both
single and multiple layups. As can be noted from the resulting values, given in Table
4.7, the theoretical transverse shear rigidity is practically equal for both layups.

U (N)
Layup
Eq. (3.28)

Single 4763

Multiple 4777

Table 4.7: Theoretical transverse shear rigidity of the sandwich panels produced

43
4.5 Testing

The single and multiple layer sandwich test specimens produced were then tested
according to the respective standards. Test results allow for the calculation and
comparison of the mechanical properties of the two different sandwich layups. A screw-
driven Instron® 4206-606 testing machine was used. Different testing jigs were attached
to the testing machine for the three different experimental procedures carried out.

4.5.1 Long Beam Flexure Test

The long beam flexure test was carried out as specified by ASTM D7249 [30]. The
standard configuration of this test method is 4-point loading, subjecting the panels to a
bending moment in such a way that compressive and tensile forces are produced in the
top and bottom facings, respectively. Readings of the mid-span deflection and strain
developed in the facings with respect to the applied load, allow for the calculation of
facing properties, namely the ultimate facing stress, 𝜎𝑓,𝑢𝑙𝑡 , the effective facing chord
modulus, 𝐸𝑓 , and the overall flexural stiffness of the sandwich panel, 𝐷 [30]. Strain
gauges were attached to the test specimens at the centre of each of facing, aligned
lengthwise.

Figure 4.7: Long beam flexure set-up, 4-point standard configuration

The mid-span deflection was recorded using a Vernier height gauge, using the set-up
shown in Figure 4.7. The height gauge was lowered at 5 mm intervals and force and
strain readings were taken when the lower surface of the test specimen touched the

44
measuring jaw of the Vernier height gauge. The standard configuration has a support
span, S of 560 mm and a loading span, L of 100 mm.

When loading the specimen into the testing fixture, pressure pads were inserted between
the test specimen and the loading and support bars in order to prevent any local damage
and failure by indentation. The standard recommends a constant crosshead speed of
6 mm/min, so as to produce failure within 3 to 6 minutes. At this speed, testing of each
specimen took between 15 to 18 minutes and failure was not achieved. The tests had to
be stopped due to set-up limitations that did not allow for mid-span deflections greater
than 100 mm. The resulting large deflections are shown in Figure 4.8. Even though
failure was not achieved, it was noted that for most specimens, the applied load had
reached a fairly constant maximum value when testing was stopped.

Figure 4.8: Long beam, 4-point loading deflection

The long beam specimens did not fail within the specified time frame, even though
specimen sizing and testing were carried out as specified in ASTM D7249 [30]. The
large deflections exhibited by the specimens can be linked to the fact that the sandwich
panels are made of natural and bio-based materials, indicating that the standard’s
guidelines and parameters are limited to the testing of materials used in the commercial
production of sandwich panels. Updated specimen geometry and testing parameters may
need to be specified for the testing of sandwich panels made from natural and bio-based
materials.

45
Figure 4.9: Long beam, 3-point loading deflection

As can be deduced from sections 3.4.1 and 3.5.1, 3-point loading may result in a larger
bending moment than that produced by 4-point loading, depending on the length chosen
for the loading and support spans. For this reason, two specimens, one single and one
multiple, were tested under 3-point loading conditions, with a support span of 520 mm.
The change in loading configuration increased the bending moment at the mid-span of
the beam from 𝑀 = 0.115𝑃 under 4-point loading, to 𝑀 = 0.13𝑃 under 3-point loading.
The modified set-up and loading configuration offered no limitations, allowing both
single and multi-layered specimens to reach failure. The deflection of the long beam
specimen under 3-point loading is shown in Figure 4.9.

4.5.2 Short Beam Shear Test

The short beam shear test was carried out following ASTM C393 [31] guidelines. The
test method allows for the determination of sandwich core shear properties by subjecting
the panels to a bending moment. The support span specified by the standard promotes
failure by core shear [31].

The testing jig required for 3-point bending was attached to the testing machine and the
specimens were loaded shown in Figure 4.10. A pressure pad was inserted between the
test specimen and the loading bar in order to prevent any local damage which may lead
to failure by indentation. The crosshead displacement speed was set to 6 mm/min in
order to achieve failure within 3 to 6 minutes. The support span, as specified by ASTM

46
C393 [31] for the standard configuration, is 150 mm. Readings of the applied load and
the mid-span deflection were recorded synchronously by the testing machine.

Figure 4.10: Short beam shear test loading configuration

4.5.3 Edgewise Compression Test

The edgewise compression test was carried out according to the guidelines given in
ASTM C364 [33]. The test allows for the determination of the compressive properties
of the sandwich panel, namely the facing compressive stress, by subjecting the test
specimen to an end-loading. The facing compressive stress provides a basis for analysing
the load-carrying capacity of the two different sandwich panel layups under
consideration. The testing jig required for edgewise compression was fitted in the
Instron® 4206-606 testing machine and the test specimen was secured into place, as
shown in Figure 4.11.

In order to prevent any localised end failures, it was ensured that the ends of the test
specimens are right, flat, smooth and free of burrs, so that the applied load is distributed
uniformly to each of the facings. The crosshead displacement was set to 0.5 mm/min as
specified in the standard, to result in failure within 3 to 6 minutes. Readings of the
applied loading with respect to the end compression were recorded synchronously by
the testing machine.

47
Figure 4.11: Edgewise compression test loading configuration

4.6 Conclusion

Chapter 4 gives an overview of the materials chosen for this study. Both single and
multi-layered sandwich panels were produced, using a combination of hand layup and
vacuum bagging techniques. The sandwich panels were cut into the required test
specimen geometry and tested according to the relevant standards. The results obtained,
together with the application of theory, were used to calculate various facing and core
properties. The following chapter presents the results and calculated mechanical
properties, as well as a comparison between the different sandwich layups.

48
5 Results and Discussion

5.1 Overview

The readings obtained from the testing procedures explained in the previous chapter
allow for the calculation of different sandwich properties. Single and multi-layered
sandwich panels of the same thickness were tested to analyse the effects on the overall
sandwich properties and performance. Application of the ASTM standards allowed for
the experimental determination of various material and overall sandwich properties that
can be compared to properties given in datasheets or derived through theory. The
following chapter discusses these results, together with any formulated observations
made during testing.

5.2 Results

The ASTM standards applied in this study state that at least five specimens are to be
tested to obtain valid results. Therefore, six single and six multi-layered specimens were
tested for each different testing procedure.

Prior to testing, the single and multi-layered specimens were viewed under an optical
microscope to measure the resulting facing and core thicknesses of the two layups.
Photomicrographs of the top and bottom facings of the single-layered panels are shown
in Figures 5.1a and 5.1b, respectively. Figures 5.2a and 5.2b show photomicrographs of
the multi-layered sandwich panels, along with facing and core thickness measurements,
taken at four different points across the width of the specimen. The overall sandwich
panel thickness was measured at different points along the length, using Vernier calipers.
The resulting sandwich thickness parameters are given in Table 5.1.

In this study, the sandwich panels were tested under different loading conditions, namely
flexure and edgewise compression. Each experimental procedure required different
spans and specimen geometry, that were also used in the calculations. The different
spans and dimensions are summarised in Table 5.2. The spans of the standard
configurations applied in flexure testing were set as specified in the respective standards,
namely ASTM D7249 [30] for the long beam flexure tests and ASTM C393 [31] for the
short beam shear tests.

49
Figure 5.1a: Photomicrograph of the top section of the single-layered panel

Figure 5.1b: Photomicrograph of the bottom section of the single-layered panel

Single layup Multiple layup

t (mm) 1.59 1.13

c (mm) 7.23 8.38

d (mm) 8.82 9.51

dT (mm) 10.4 10.6

dC (mm) - 4.76

cM (mm) - 3.63

Table 5.1: Facing and core thicknesses of the sandwich panel configurations

50
Figure 5.2a: Photomicrograph of the top section of the multi-layered panel

Figure 5.2b: Photomicrograph of the bottom section of the multi-layered panel

Loading
S (mm) L (mm) b (mm)
Configuration

Long beam flexure 4- point 560 100 75

ASTM D7249 [30] 3-point 520 - 75

Short beam shear


3-point 150 - 75
ASTM C393 [31]

Edgewise compression
- - 53 50
ASTM C364 [33]

Table 5.2: Support and loading spans used in the flexure tests
51
5.2.1 Long Beam Flexure Tests according to ASTM D7249 [30]

The long beam specimens were tested under 4-point loading according to ASTM D7249
[30], as explained in section 4.5.1. The mid-span deflection and the strain were recorded
with respect to the applied load. Figures 5.3 and 5.4 plot the applied load against the
mid-span deflection for the single and multi-layered specimens, respectively. The
figures also plot the theoretical load-deflection behaviour of the long beams under
4-point loading, assuming an elastic, perfectly plastic model. As can be noted, the
experimental curves lie between the theoretical lower and upper bounds.

The theoretical model assumes an initial linear elastic behaviour. Once the maximum
theoretical load is reached, the material produces a plastic response. In plastic behaviour,
the strain increases indefinitely at constant stress and therefore, the theoretical model
displays a constant maximum load [57]. The load-deflection relationships plotted in
Figures 5.3 and 5.4 were predicted by equation (3.38), substituting the lower and upper
bounds of the flexural rigidity, as given in Table 4.6. The maximum theoretical load was
predicted by equation (3.42), for face yielding failure. From Figures 5.3 and 5.4, it can
be noted that the single-layered specimens are able to withstand higher maximum loads
than the multi-layered specimens, as predicted theoretically. The face yielding failure
load was calculated using the facing stress, 𝜎𝑓 , as obtained experimentally from
application of ASTM C393 [31] in section 5.2.2. The facing stress and the theoretical
face yielding failure load are summarised in Table 5.3, for both sandwich layups.

Figure 5.5 shows the mid-section of test specimen S4. The lighter colour at the centre of
the bottom facing indicates that the matrix reached its yield point. Further loading would
have resulted in face yielding, however testing of the long beams under 4-point loading
had to be stopped at a mid-span deflection of 100 mm due to set-up limitations, as
explained in section 4.5.1.

Facing stress (MPa) Theoretical face yielding failure load (N)

Eq. (3.52) Eq. (3.42)

Single layup 48.0 440

Multiple layup 52.6 369

Table 5.3: Facing stress and face yielding failure load for the single and multiple layups

52
The maximum load reached by each specimen was used to calculate the ultimate facing
stress, 𝜎𝑓,𝑢𝑙𝑡 , by equation (3.39). The strain readings closest to 1000 μԑ and 3000 μԑ were
used to determine the facing moduli, 𝐸𝑓,𝑡 and 𝐸𝑓,𝑏 , where subscripts 𝑡 and 𝑏 represent
the top and bottom facings, respectively. The facing moduli and the flexural rigidity, 𝐷,
were calculated by equations (3.40) and (3.41), respectively, as given in Table 5.4.
0.5

0.4
Theoretical -
Lower Bound
Theoretical -
Applied Load (kN)

0.3 Upper Bound


S1

S2
0.2
S3

S4
0.1
S5

0.0
0 20 40 60 80 100 120
Mid-Span Deflection (mm)
Figure 5.3: Applied load (kN) vs. mid-span deflection (mm) for the single-layer, long beam
flexure tests under 4-point loading
0.5

0.4
Theoretical -
Lower Bound
Theoretical -
Applied Load (kN)

0.3 Upper Bound


M1

M2
0.2
M3

M4
0.1
M5

0.0
0 20 40 60 80 100 120
Mid-Span Deflection (mm)
Figure 5.4: Applied load (kN) vs. mid-span deflection (mm) for the multi-layer, long beam
flexure tests under 4-point loading

53
Matrix
yielding

Figure 5.5: Matrix yielding, following 4-point bending

𝜎𝑓,𝑢𝑙𝑡 (MPa) 𝐸𝑓,𝑡 (GPa) 𝐸𝑓,𝑏 (GPa) 𝐷 (Nm2 )


Specimen
Eq. (3.39) Eq. (3.40) Eq. (3.40) Eq. (3.41)

S1 44.8 5.10 5.46 28.9

S2 38.4 5.61 5.54 30.6

S3 40.0 5.48 5.04 28.8

S4 47.0 5.23 5.43 29.2

S5 39.8 5.07 5.50 28.9

Sample Mean 42.0 5.30 5.39 29.3

Standard Deviation 3.66 0.237 0.203 0.736

Coeff. of Variation (%) 8.73 4.47 3.77 2.51

M1 46.3 6.07 5.66 25.2

M2 49.7 7.92 7.15 32.3

M3 49.6 6.11 6.29 26.6

M4 47.2 5.83 5.94 25.3

M5 45.3 6.03 5.80 25.4

Sample Mean 47.6 6.39 6.17 26.9

Standard Deviation 1.98 0.860 0.596 3.03

Coeff. of Variation (%) 4.15 13.5 9.66 11.3

Table 5.4: Facing property results according to ASTM D7249 [30]


(t = top facing, b = bottom facing)

54
The theoretical facing modulus, 𝐸𝑓 , was calculated in section 4.4 by application of the
practical rule of mixtures [1], that takes into account the fact that the facings are made
of a woven twill fabric. Figure 5.6 presents a visual comparison of the theoretical and
experimental facing moduli, of both the single and multi-layered sandwich panels. The
experimental values plotted are the sample means of each layup, as given in Table 5.4.
As can be noted from Figure 5.6, the top and bottom facing moduli of each layup
compare well to each other, as well as to the theoretical values. Contrary to theoretical
predictions, the experimental facing moduli of the multi-layered panels are slightly
higher than those of the single-layered ones. The higher facing modulus of the
multi-layered sandwiches may be attributed to the larger adhesive surface area between
the flax and the cork, in this layup.

8
7
Facing Modulus (GPa)

6
5
4
3
2
1
0
Single Multiple
Top, Experimental (GPa) 5.30 6.39
Bottom, Experimental (GPa) 5.39 6.17
Theoretical (GPa) 5.98 5.88

Figure 5.6: Comparison of the theoretical and experimental values for the facing modulus, 𝐸𝑓

The experimental values of the facing moduli were used in the calculation of the
sandwich flexural rigidity, 𝐷 in accordance with ASTM D7250 [43]. For both sandwich
layups, the lower of the top and bottom facing moduli calculated in this section was
used, in order to obtain more conservative results for the flexural rigidity. Section 5.2.3
compares the theoretical sandwich flexural rigidity calulcatued in section 4.4, to the
experimental values obtained from both ASTM D7249 [30] and ASTM D7250 [43].

Five of the long beam specimens were tested under a 4-point loading configuration, the
standard configuration of ASTM D7249 [30]. Flexure of these specimens had to be
stopped at a mid-span deflection of 100 mm due to set-up limitations and failure was not
reached. Besides this, most of the specimens had reached a fairly constant load value, as
can be noted from Figures 5.3 and 5.4. As explained in section 4.5.1, the sixth specimen
was tested under 3-point loading to produce a higher bending moment at the mid-span

55
of the beam. The increase in bending moment did, in fact, result in failure of both single
and multi-layered specimens, as can be noted from the load-deflection curves plotted in
Figures 5.7 and 5.8. Even though strain gauges had already been attached to the top and
bottom facings of specimens S6 and M6, the strain was not recorded due to interference
between the 3-point loading fixture and the strain gauge attached to the top facing.
Figures 5.7 and 5.8 also plot the theoretical lower and upper bounds for the load-
deflection behaviour, as given by equation (3.50) for 3-point loading. The maximum
theoretical load was predicted by equation (3.42), for face yielding failure. As can be
noted, both experimental curves lie between the theoretical lower and upper bounds.

0.4

0.3
Applied Load (kN)

0.2

0.1

0.0
0 20 40 60 80 100 120 140 160
Mid-Span Deflection (mm)
S6 Theoretical - Lower Bound Theoretical - Upper Bound

Figure 5.7: Applied load (kN) vs. mid-span deflection (mm) for the single-layer, long beam
flexure specimen, S6 under 3-point loading
0.4

0.3
Applied Load (kN)

0.2

0.1

0
0 20 40 60 80 100 120 140 160
Mid-Span Deflection (mm)
M6 Theoretical - Lower Bound Theoretical - Upper Bound

Figure 5.8: Applied load (kN) vs. mid-span deflection (mm) for the multi-layer, long beam
flexure specimen, M6 under 3-point loading

56
The increase in bending moment at the mid-span of the beams resulted in face wrinkling
of the top facing of specimen S6, as shown in Figures 5.9a and 5.9b. Face wrinkling
failure indicates that the compressive stress developed in the facing during bending
exceeded the face wrinkling strength of the facing, as can be confirmed by the values
given in Table 5.5. The upper bound for the face wrinkling strength of the single layup
was computed by the Hoff and Maunter formula [42], equation (3.43). Specimen S6 also
showed signs of matrix yielding on the bottom facing. The different types of failure
experienced by the two facings can be explained by the fact that under 3-point loading,
the top facing was in compression, while the bottom facing experienced tension.

Face wrinkles

Figure 5.9a: Top view of specimen S6 Figure 5.9b: Side view of specimen S6
showing signs of face wrinkling failure showing signs of face wrinkling failure

Facing Compressive Stress (MPa) Face Wrinkling Strength (MPa)


Eq. (3.39) Eq. (3.43)

47.2 37.6

Table 5.5: The maximum stress developed in the top facing of specimen, S6 and the face
wrinkling strength of the single layup

From Figures 5.3 and 5.4, as well as from Figures 5.7 and 5.8, it can be noted that the
maximum loads reached by the multi-layered specimens are generally lower than those
of the single-layered specimens. Under 3-point loading, the long beam, single-layered
specimen S6 reached a maximum load of around 380 N, while that attained by the
multi-layered specimen M6 is approximately 300 N. The higher maximum loads reached
by the single sandwiches can be linked to the higher facing thickness of this layup. The
single-layered specimens have a mean facing thickness of 1.59 mm, while the average
facing thickness of the multi-layered specimens is 1.13 mm.
57
The multi-layered specimen, M6 suffered extensive damage under 3-point loading,
besides having reached a lower maximum load. The top facing has visible signs of face
wrinkling failure, while the bottom facing underwent extensive face yielding under
tension, leading to fracture of both the fibres and the resin, as shown in Figure 5.10a.
Fracture of the bottom facing led to crack propagation through the whole core thickness,
as shown in Figure 5.10b. Crack propagation was delayed by the intermediate flax layer,
leading to debonding of the bottom core layer from the flax laminate.

Face fracture

Crack propagation

Figure 5.10a: Top view of specimen Figure 5.10b: Side view of specimen M6,
M6, showing face fracture showing crack propagation

5.2.2 Short Beam Shear Tests following ASTM C393 [31]

The short beam specimens were tested under a 3-point loading configuration following
ASTM C393 [31], as explained in section 4.5.2. Figures 5.11 and 5.12 plot the applied
load against the mid-span deflection as recorded by the Instron® 4206-606 testing
machine, for the single and multi-layered specimens, respectively. The figures also plot
the theoretical load-deflection behaviour as given by equation (3.50) for 3-point loading,
assuming an elastic, perfectly plastic model. The theoretical failure load was predicted
by equation (3.53), for core shear failure. As can be noted from the graphs, the
experimental curves of the multi-layered specimens fall between the theoretical lower
and upper bounds. On the other hand, the experimental curves of single-layered
specimens fall outside the theoretical upper bound, meaning that for a particular applied
load, the short beam specimens displayed a lower mid-span deflection than that
predicted by equation (3.50). With reference to this equation, a lower mid-span
deflection indicates that the single-layered specimens attained a higher flexural and/or
shear rigidity than that predicted by theory.

58
From Figures 5.11 and 5.12 it can also be observed that, contrary to theoretical
predictions, most of the single-layered specimens attained much higher maximum loads
than the multi-layered specimens. The specimen geometry specified by ASTM C393
[31] promotes core shear failure, therefore, this observation indicates that the single-
layered panels will attain a higher core shear strength than the multi-layered panels.

1.6

1.4
Theoretical -
1.2 Lower Bound
Theoretical -
Upper Bound
1.0
Applied Load (kN)

S1

0.8 S2

0.6 S3

S4
0.4
S5
0.2
S6
0.0
0 5 10 15 20 25
Mid-Span Deflection (mm)
Figure 5.11: Applied load (kN) vs. mid-span deflection (mm) for the single layer, short
beam flexure tests under 3-point loading

1.6

1.4
Theoretical -
Lower Bound
1.2
Theoretical -
Upper Bound
1.0 M1
Applied Load (kN)

0.8 M2

0.6 M3

0.4 M4

M5
0.2
M6
0.0
0 5 10 15 20 25
Mid-Span Deflection (mm)
Figure 5.12: Applied load (kN) vs. mid-span deflection (mm) for the multiple layer, short
beam flexure tests under 3-point loading

59
Following ASTM C393 [31], the maximum load reached by each specimen, 𝑃𝑚𝑎𝑥 was
used to calculate the core shear strength, 𝜏𝐶𝑆 by equation (3.51) and the facing stress, 𝜎𝑓
by equation (3.52). The results are given in Table 5.6. Values obtained for the core shear
strength confirm that this sandwich property is, in fact, higher for the single-layered
specimens, even though both single and multiple sandwich layups have the same total
core thickness.

𝜏𝐶𝑆 (MPa) 𝜎𝑓 (MPa)


Specimen 𝑃𝑚𝑎𝑥 (N)
Eq. (3.51) Eq. (3.52)
S1 1429 1.08 50.9
S2 1401 1.06 49.9
S3 1305 0.99 46.4
S4 1188 0.90 42.3
S5 1305 0.99 46.4
S6 1470 1.11 52.3
Sample Mean 1350 1.02 48.0
Standard Deviation - 0.078 3.69
Coefficient of Variation (%) - 7.68
M1 1148 0.80 53.3
M2 1180 0.83 54.8
M3 1132 0.79 52.6
M4 1071 0.75 49.8
M5 1099 0.77 51.1
M6 1156 0.81 53.7
Sample Mean 1131 0.79 52.6
Standard Deviation - 0.028 1.84
Coefficient of Variation (%) - 3.51

Table 5.6: Core shear property results according to ASTM C393 [31]

60
Figure 5.13 provides a visual comparison of the core shear strength of both single and
multiple sandwich layups, to the shear strength of the cork core given in the datasheet
[50]. The core shear strength of the single-layered panels is very comparable to the shear
strength of the cork, with the difference between the two values being just
11.8 %. Slight variations are expected since cork is a natural material. On the other hand,
the fact that the mean core shear strength of the multi-layered specimens is lower than
that claimed by the manufacturer indicates that these panels may not have failed by core
shear. Furthermore, the multi-layered panels generally experienced higher facing
stresses than the single-layered panels, as can be observed from the results given in Table
5.6. The higher facings stresses that developed in the multi-layered panels is another
indication that these specimens may have experienced face failure rather than core shear.
Core Shear Strength (MPa)

1.2

0.8

0.6

0.4

0.2

0
Single Multiple
Experimental (MPa) 1.02 0.79
Theoretical (MPa) 0.90 0.90

Figure 5.13: Comparison of the core shear strength, 𝜏𝐶𝑆 [50]

The failure modes exhibited by the short beam specimens were similar to those
experienced by the long beam specimens. Short beam specimens S6, M5 and M6
experienced fracture of the bottom facing, with a few other specimens showing early
signs of face yielding on the same facing. Moreover, most of the specimens exhibited
signs of face wrinkling failure on the top facing.

A closer look at the short beam specimens that suffered face fracture revealed an
interesting observation. The single-layered specimen experienced fracture of both the
fibres and the resin, while the multi-layered specimens experienced only resin fracture.
The different modes of face fracture, shown in Figures 5.14a and 5.14b, indicate that the
multi-layered specimens are able to withstand higher facing stresses than the single-
layered ones, as confirmed by the values obtained for the facing stress in Table 5.6. The

61
higher facing stresses attained by the multi-layered panels may be attributed to the larger
adhesive surface area between the flax and the cork in this layup. Furthermore, the
intermediate flax layer of the multi-layered specimens limited crack propagation to just
one core thickness, whereas face fracture of the single-layered specimen resulted in
crack propagation across the whole core thickness.

Figure 5.14a: Fibre and resin failure Figure 5.14b: Resin fracture of the
of the single-layered specimen, S6 multi-layered specimen, M6

ASTM C393 [31] states that the only acceptable failure modes for the short beam flexure
tests are debonding and core shear. Even though the specimens were sized and tested as
specified by the same standard, different modes of failure were observed. Table 5.7 lists
the theoretical failure loads of the short beam specimens under core shear, face wrinkling
and face yielding. The face wrinkling load is based on the upper limit of the theoretical
facing modulus. As can be noted from the table, the theoretical failure loads for face
wrinkling are lower than those for core shear, for both single and multi-layered sandwich
layups, meaning that, theoretically, the short beam specimens considered in this study
should initially undergo face wrinkling failure under 3-point loading.

Core Shear Failure Face Wrinkling Face Yielding


Layup Load (N) Failure Load (N) Failure Load (N)

Eq. (3.53) Eq. (3.44) Eq. (3.42)

Single 1191 1053 1348


Multiple 1284 796 1132

Table 5.7: Comparison of the different failure loads for the short beam specimens

62
Comparing the maximum loads reached by each specimen (Table 5.6) to the different
theoretical failure loads (Table 5.7) gives an indication of the failure modes experienced
by each specimen. From Table 5.8 it is clear that the failure loads achieved are in line
with the observations made, as all specimens exceeded the theoretical face wrinkling
load, with some specimens also exceeding the face yielding failure load.

Specimen M5 did not exceed the theoretical face yielding failure load even though it
experienced face fracture. The maximum load reached by this specimen is fairly close
to the theoretical face yielding failure load, however, premature face yielding and resin
fracture could indicate that this particular specimen has a slightly higher resin volume
fraction than its counterparts, making it more brittle and thus, weaker. Furthermore, it is
expected that the resin fractures first, followed by the fibres, as the latter have a higher
flexural strength.

Specimen Core Shear Failure Face Wrinkling Failure Face Yielding Failure

S1 X X X

S2 X X X

S3 X X

S4 X

S5 X X

S6 X X X

M1 X X

M2 X X

M3 X X

M4 X

M5 X

M6 X X

Table 5.8: The different theoretical failure loads achieved by each short beam specimen

63
Another observation that can be made from Table 5.8 is the fact that nearly all single-
layered specimens exceeded the theoretical core shear failure load, while none of the
multi-layered specimens did. As predicted previously, from the results given in Table
5.6, the two layups experienced different failure modes. The fact that failure of the multi-
layered specimens was dominated by face wrinkling and face yielding rather than core
shear, can be attributed to the lower top and bottom facing thicknesses of this layup.

The specimen geometry specified by ASTM C393 [31] should promote core shear
failure and debonding. As observed, not all specimens managed to reach one of the
required failure modes. A similar problem was encountered with the long beam
specimens, following test method ASTM D7249 [30]. The ASTM standards used in this
study may be applied in the testing of fibre composite sandwich structures, however,
that fact the it was difficult to reach the required type of failure in both short and long
beam flexure tests, could indicate that the standards are limited to the testing of sandwich
panels made of fibres that are brittle in nature. The natural fibre composite sandwiches
tested in this study are not brittle, as evident by the observed fraying and fibre pull-out
at the fracture sites. A revision of the specimen geometry and testing parameters
specified by the ASTM standards would enable application to sandwich panels made
from natural materials. Ensuring that the required type of failure is achieved would
minimise the problems encountered during testing and allow for a more accurate
calculation of sandwich properties.

5.2.3 Application of ASTM D7250 [43]

The sandwich panels’ flexural and shear rigidity were determined experimentally by
application of test method ASTM D7250 [43]. The load-deflection data recorded from
the short beam flexure tests was used.

The sandwich flexural stiffness, 𝐷 was calculated by equation (3.54). The flexural
stiffness depends on the thickness parameters of the panel and on the facing modulus,
𝐸𝑓 , as determined experimentally from test method ASTM D7249 [30] in section 5.2.1.
The shear rigidity, 𝑈 and the core shear modulus, 𝐺𝑐 were calculated by equations (3.55)
and (3.56), respectively. Following the procedure specified by the standard, the
transverse shear rigidity of each specimen was calculated at ten different force levels,
evenly spaced over the force range. The mean value was then found and used in the
calculation of the core shear modulus, as given in Table 5.9.
64
𝐷 (Nm2 ) 𝑈 (N) 𝐺𝑐 (MPa)
Specimen
Eq. (3.54) Eq. (3.55) Eq. (3.56)
S1 9573 11.9
S2 9479 11.7
S3 7628 9.4
24.9
S4 6752 8.4
S5 9632 11.9
S6 7781 9.6
Sample Mean - 8474 10.5
Standard Deviation - 1243 1.54
Coeff. of Variation (%) - 14.7
M1 3695 4.6
M2 4296 5.3
M3 4381 5.4
23.8
M4 4193 5.2
M5 4546 5.6
M6 4879 6.0
Sample Mean - 4332 5.35
Standard Deviation - 393 0.49
Coeff. of Variation (%) - 9.1

Table 5.9: Sandwich beam flexural and shear stiffness according to ASTM D7250 [43]

Both ASTM D7249 [30] and ASTM D7250 [43] allowed for the determination of the
sandwich flexural stiffness. The two sets of experimental values obtained from these test
methods are plotted in Figure 5.15, for both the single and multiple sandwich layups.
The figure also plots the theoretical flexural stiffness, as calculated by the practical rule
of mixtures for 𝛼 = 0.5 [1], given in section 4.4. For both single and multiple sandwich
layups, the experimental results for the flexural rigidity are very comparable to each
other, as well as to the theoretical value.

65
The difference between the two sets of experimental results is less than 18 %, for both
single and multiple sandwich layups, indicating that evaluation of the flexural stiffness
using the facing modulus gives comparable results. From equation (3.54) for the flexural
rigidity, it can be noted this sandwich property is proportional to the facing modulus.
From application of ASTM D7249 [30], it was inferred that the multiple sandwich panels
have a facing modulus that is higher than that of the single-layered specimens. Besides
this, the flexural rigidity of the single-layered panels is generally higher than that of the
multiple, as can be observed from Figure 5.15. Flexural rigidity increases with core
thickness and so, the higher flexural rigidity attained by the single sandwiches can be
linked to the fact that even though both sandwich layups have the same total core
thickness, the core of the single layup is continuous, while that of the multiple layup is
interrupted by the intermediate flax layer [1].

35
Sandwich Flexural Stiffness (Nm2)

30

25

20

15

10

0
Single Multiple
D7249 (Nm2) 29.3 26.9
D7250 (Nm2) 24.9 23.8
Theoretical (Nm2) 28.0 22.1

Figure 5.15: Comparison of the theoretical (α = 0.5) and experimental values


calculated for the sandwich flexural stiffness

The experimental transverse shear rigidity (equation (3.55)) is dependent on the


load-deflection data, the flexural rigidity and the support span values. The
load-deflection data recorded from the 3-point flexure tests of the short beam specimens
was applied, along with the flexural rigidity derived from application of ASTM D7250
[43], as given in Table 5.9. Figure 5.16 compares the experimental and theoretical values
of transverse shear rigidity, for both single and multiple layups. The theoretical shear
rigidity was calculated by equation (3.28). As can be noted from Figure 5.16, the mean
experimental value obtained for the multiple layup compares very well to the theoretical,
with the difference between the two values being just 10.2 %.
66
The equation for the theoretical transverse shear rigidity, equation (3.28), predicted an
almost equal shear rigidity for the two sandwich layups. Contrary to this prediction, the
experimental transverse shear rigidity of the single layup is approximately twice that of
the multiple, as can be noted from Figure 5.16. The reason behind this discrepancy was
found by taking a closer look at the deflection data of the short beam specimens. For the
same applied load, the mid-span deflection of the single-layered specimens was much
lower than that recorded for the multi-layered panels. The difference in deflection
behaviour between the two layups was not observed in the long beam flexure tests.

10000
Transverse Shear Rigidity (N)

8000

6000

4000

2000

0
Single Multiple
Experimental (N) 8474 4332
Theoretical (N) 4763 4777

Figure 5.16: Comparison of the theoretical and experimental transverse shear rigidity

A study carried out by J. Arbaoui [58], analyses the effect of core thickness and multiple
layers on the mechanical properties of polypropylene honeycomb multi-layer sandwich
structures. In this study, the experimental transverse shear rigidity of single and double
sandwich layups of same overall core thickness differed by 19 %. A study carried out
by Ferreira et al., [26] on the experimental characterisation of cork agglomerate core
sandwich panels for wall assemblies in buildings, applied various ASTM standards
including D7250 [43]. The same standard was applied in a study carried out by Jiang et
al. [59], on a new approach to manufacturing bio-composite sandwich structures from
mycelium-based cores. Both studies investigate the properties of bio-based sandwich
panels and apply ASTM D7250 [43] to determine only the flexural rigidity, with no
mention of the transverse shear rigidity. The fact that both studies omit the transverse
shear rigidity could indicate that these researchers also encountered discrepancies in the
calculated shear rigidity, similar to those observed in this study.

67
Test method ASTM D7250 [43] lists two different procedures for the calculation of the
sandwich panels’ flexural and shear rigidity. The first method is used when the facing
modulus is unknown and makes use of the load-deflection data collected from flexure
tests conducted under two loading configurations, on two different test specimens. The
calculations require the displacement values of the two specimens at the same force
values. In this study, the force range of the short beam specimens under 3-point loading
is not comparable to the force range of the long beam specimens under 4-point loading
and therefore, the first method stated in ASTM D7250 [43] could not be applied.

The second method listed in ASTM D7250 [43], for the calculation of the sandwich
panels’ flexural and shear rigidity, may be used when the facing modulus is known. In
this study, the facing modulus was determined experimentally from test method ASTM
D7249 [30]. The calculation makes use of the load-deflection data collected from flexure
testing carried out under one loading configuration.

The results given in Table 5.9 were calculated according to the second method, using
the deflection data collected from flexure of the short beam specimens following ASTM
C393 [31]. The force-displacement data of the short beam specimens was preferred over
that of the long beam specimens, since the force-displacement data of the short beams
was recorded synchronously by the Instron® 4206-606 testing machine. Furthermore,
the short beam specimen data was also favoured due to the linear relationship between
the applied load and the mid-span deflection.

The experimental transverse shear rigidity, as given in Table 5.9, was used in the
calculation of the core shear modulus, by equation (3.56). The value obtained for the
core shear modulus of the multiple layup is 5.3 MPa, which compares excellently to the
shear modulus of the cork specified in the datasheet, that of 5.9 MPa [50]. On the other
hand, there is a large discrepancy between the core shear modulus of the single layup
and the value specified by the manufacturer. A similar observation was made in the study
carried out by J. Arbaoui, where the experimental value calculated for the core shear
modulus of the single sandwiches having a 20 mm core differs by 33.9 % from that
specified by the manufacturer. On the other hand, the discrepancy between the
theoretical and experimental values obtained for the double sandwiches having 10 mm
cores is just 0.3 % [58].

68
5.2.4 Edgewise Compression Tests following ASTM C364 [33]

The edgewise compression specimens were loaded as explained in section 4.5.3. The
shortening of the end subjected to the edgewise loading was recorded by the Instron®
4206-606 testing machine. Figures 5.17 and 5.18 plot the applied load against end
shortening for the single and multi-layered specimens, respectively.

10

8
S1
7
Applied Load (kN)

S2
6
S3
5
S4
4
S5
3
S6
2

0
0 1 2 3 4 5
End Shortening (mm)

Figure 5.17: Applied load (kN) vs. end shortening (mm) for the single layer,
edgewise compression tests
10

8
M1
7
M2
Applied Load (kN)

6
M3
5
M4
4
M5
3
M6
2

0
0 1 2 3 4 5
End Shortening (mm)

Figure 5.18: Applied load (kN) vs. end shortening (mm) for the multiple layer,
edgewise compression tests

69
A common feature that can be noted from the plots is that all specimens, both single and
multiple, exhibited a plateau at an applied load of approximately 2 kN. Such behaviour
indicates that the strain in the specimen kept increasing with little to no increase in stress,
meaning that some type of failure took place. At this point, failure is likely to be local,
such as resin breakage, delamination or core crushing. Following this failure region, the
specimen regained its strength, withstanding increasing applied loads up to eventual
failure. A similar load-deflection curve was observed in an edgewise compression study
carried out by W. A. Samad et al. [60].

𝜎𝑓,𝑐 (MPa)
Specimen 𝑃𝑚𝑎𝑥 (N)
Eq. (3.57)
S1 8.26 51.9
S2 8.53 53.6
S3 8.42 52.9
S4 8.62 54.2
S5 8.19 51.5
S6 8.13 51.1
Sample Mean 8.36 52.5
Standard Deviation - 1.22
Coefficient of Variation (%) - 2.33
M1 9.13 53.8
M2 8.80 51.9
M3 8.61 50.7
M4 9.02 53.2
M5 8.70 51.3
M6 8.87 52.3
Sample Mean 8.85 52.2
Standard Deviation - 1.15
Coefficient of Variation (%) - 2.21

Table 5.10: Facing compressive stress results according to ASTM C364 [33]

70
The maximum loads reached by each specimen were used to calculate the facing
compressive stress, 𝜎𝑓,𝑐 of each specimen, by equation (3.57). As can be noted from the
results given in Table 5.10, the single and multi-layered panels attained facing
compressive stresses that are practically equal, even though the multi-layered panels
have thinner top and bottom facings. The equal facing compressive stresses can be linked
to the fact that the in-plane loading resulting from edgewise compression is being taken
up by six flax layers in both layups.

It was also observed that the multi-layered sandwich panels were able to withstand
maximum loads that are around 6 % higher than those attained by the single-layered
sandwich panels. The higher maximum loads attained by the multiple layup may be
attributed to the larger adhesive surface area between the flax and the cork in this layup.

Different failure modes were observed for the single and multi-layered panels. In
general, the single-layered sandwiches failed by Euler buckling, while the multi-layered
sandwiches failed by shear crimping. Figures 5.19 and 5.20 show two single specimens
that failed by Euler buckling. Even though these two columns exhibited different
buckled shapes, both failed by Euler buckling. Face wrinkling was observed in both
sandwich layups, however, the extent of this type of failure was more visible in the
multi-layered sandwich panels. The reason why this type of failure was more evident in
the multi-layered sandwiches is the fact that the multiple layup has thinner top and
bottom facings. Figure 5.21 shows a buckled multi-layered specimen that also
experienced face wrinkling failure.

Most of the multi-layered panels exhibited shear crimping failure, as shown in Figure
5.22. Shear crimping is short-wavelength buckling, initiated by core shear. From
equation (3.60), it follows that the shear crimping buckling load is a function of core
thickness. Therefore, the reason why the dominant mode of failure observed in the
multi-layered sandwich panels is shear crimping, is that even though both sandwich
layups have the same total core thickness, the thickness of each core layer of the multiple
specimens is half that of the single.

71
Figure 5.19: A single-layered Figure 5.20: Another single-layered
specimen exhibiting Euler buckling specimen that failed by Euler buckling

Figure 5.21: A buckled Figure 5.22: A multi-layered specimen


multi-layered specimen exhibiting face that failed by shear crimping
wrinkling failure

5.3 Conclusion

Chapter 5 gave an overview of the resulting material and sandwich properties obtained
from the application of the different ASTM standard testing procedures. Some degree
of variation is expected when working with natural materials, however, the standard
deviation in the results obtained is generally low. The low standard deviation indicates
high repeatability across all specimens, for the three different standard testing
procedures carried out.

72
The main aim of this study was to identify any advantages of multiple sandwich layups
over single layups of the same thickness, in terms of specimen characteristics. As was
observed, the intermediate flax layer had little effect on sandwich properties when
loaded under flexure. The multi-layered sandwich panels did attain slightly higher facing
moduli than the single, however, the latter managed to withstand maximum loads that
are between 15 to 30 % higher than those attained by the multiple. On the other hand,
when loaded under edgewise compression, the multi-layered specimens managed to
withstand maximum loads that are approximately 6 % higher than those managed by the
single sandwiches, with only slight differences in the facing compressive modulus.

Another interesting observation made during testing is the fact that different failure
modes were predominant in the single and multiple sandwich layups. The thinner facings
and the larger adhesive surface area between the flax and the cork in the multiple layup
both had an effect on the type of failure exhibited and reported for the different sandwich
structure configurations.

73
6 Conclusions and Suggestions for Future Work

6.1 Conclusions

The aim of this study was to determine the characteristic properties of green sandwich
panels, comprising Biotex® 2/2 twill 200 g flax fibre facings and Amorim® CoreCork
NL20 as core, bonded together by ONE Super Sap® high bio-based laminating epoxy.
Single and multiple sandwich layups of the same overall thickness were considered, in
order to investigate the effects that the different layups have on the properties of the
structure. The sandwich panels were tested under different loading conditions, namely
flexure and edgewise compression. Testing and calculations were carried out in
accordance with the applicable ASTM standards.

An in-depth literature review of the different materials that are currently being used in
the construction of sandwich panels was carried out. In recent years, sandwich panels
made from natural materials have received significant interest, mainly due to increased
environmental awareness and the need to develop materials from renewable sources,
through environmentally sustainable technologies. Sandwich panels made from natural
or bio-based materials are currently being used in various sectors, including the
automotive, construction and maritime industries. Environmental issues regarding the
production and disposal of sandwich panels were also addressed [14].

Following the literature review, to obtain the sandwich structures’ characteristic


properties, composite laminate micromechanics and the sandwich beam theory were
outlined by consulting established literature. Micromechanics allowed for the theoretical
prediction of facing properties, while the sandwich beam theory was used to calculate
sandwich properties, such as the flexural and transverse shear rigidity. A theoretical
analysis of the different tests was also carried out. The load-deflection relationship for
the short and long beam flexure tests was derived, assuming an elastic, perfectly plastic
model. The different failure modes expected from each loading configuration were
analysed and the theoretical failure loads calculated. The maximum loads achieved by
the specimens compared very well to theoretical predictions, with most of the
experimental curves falling between the theoretical lower and upper bounds.

74
Characterisation tests adopting ASTM standard practices, were carried out to
experimentally determine the properties of the single and multiple sandwich layups. The
long beam flexure tests were carried out according to ASTM D7249 [30], while the short
beam shear tests were conducted as specified in ASTM C393 [31]. The single-layered
specimens achieved higher maximum loads than the multiple, however, the latter
exhibited larger mid-span deflections before failing. The higher maximum loads
achieved by the single specimens may be explained by the thicker facings and
continuous core of this layup. The larger mid-span deflections exhibited by the
multi-layered specimens may be attributed to the thinner top and bottom facings, that
resulted in an increased flexibility but a lower bending strength, when compared to the
single-layered specimens. The load-deflection data of both short and long beam flexure
tests was used in the calculation of the flexural and transverse shear rigidity of the
sandwich panels, by application of ASTM D7250 [43].

Edgewise compression testing was carried out on the single and multi-layered
specimens, following test method ASTM C364 [33]. Both layups developed equal facing
compressive stresses, however the multi-layered sandwich panels were able to withstand
higher maximum loads than the single-layered specimens. Therefore, under edgewise
compression, the performance of the multiple layup is superior to that of single.

A few problems were encountered during testing. Both short and long beam flexure tests
were carried out according to standard specifications. Besides this, the long beam
specimens subjected to 4-point loading, exhibited very large deflections and did not
manage to achieve failure, due to setup limitations. The short beam specimens did fail,
however, the failure modes achieved do not conform to those required by the standard.
The shortcomings encountered during testing indicate that updated specimen geometry
and testing parameters need to be specified for the testing of sandwich panels made from
natural materials. In addition, it must be commented that these standards have originally
been formulated for use with “brittle” type fibres. It is quite evident from this study that
some modification is required to these “sandwich” standards, to enable complete
applicability to natural and bio-based materials, based on structure property
characterisation.

75
6.2 Suggestions for Future Work

Further research is required to gain more knowledge on the flax and cork sandwich
constructions considered in this study. It is important to find the optimal sandwich
configuration for the specific application, to ensure that the sandwich structure is able
to deliver the required performance.

Sandwich panels having a different facing and/or core thickness may be investigated
under the test methods applied in this study. Increasing the core thickness will increase
the bending strength and flexural rigidity of the sandwich panel, with little increase in
weight [1]. Furthermore, alternative sandwich layups and facing materials may be
considered. Natural facing materials include hemp, sisal or jute [14]. The response of
the different sandwich panels can then be compared to the results obtained from this
study, to determine which alternative is the most suitable for the required application.

It is also worth investigating the behaviour of the flax and cork sandwich panels under
impact loading. Cork has a high damage tolerance to impact loads, as well as excellent
vibration damping characteristics [21]. The dimensional recovery capability of cork
gives rise to its exceptional energy absorbing properties [61]. Furthermore, the
intermediate flax layer of the multiple layup may offer significant advantages over the
single layup, under impact loading. A study by Wang [62], on multi-layered corrugated
sandwich panels subjected to impact loading, showed that the intermediate facing layer
enhanced the energy absorption characteristics of the sandwich construction. Impact
testing will allow for a quantitative measurement of the damage resistance of the
sandwich layups under consideration. Marine vessels are continually subjected to impact
loads from waves, that may result in structural damage if the vessel does not possess
sufficient impact strength [63]. Therefore, knowledge of the damage resistance of a
sandwich construction is of significant importance in the maritime industry.

Boat hulls and other structural members are subject to high cyclic stresses, caused by
waves, cargo handling, propulsion forces, steering forces and changes in loading
conditions. The cyclic loadings make vessels and other marine vehicles prone to fatigue
[64]. Therefore, knowledge of the fatigue strength of the sandwich panels is of utmost
importance. Studying the behaviour of the flax and cork sandwich panels under cyclic
loading will also allow for an analysis of the resulting fatigue failure mechanisms.

76
The ASTM “sandwich” standards may be applied in the testing of fibre composite
sandwich structures. However, the large deflections and different failure mechanisms
exhibited by the bio-based materials used in this study, made it difficult to achieve the
failure modes specified by the standards. Natural fibres are generally highly flexible and
elastic [14], unlike the brittle type fibres that the ASTM “sandwich” standards are
intended for. A revision of these standards would enable complete application to
sandwich panels made from natural materials, based on structure property
characterisation.

A Life Cycle Assessment (LCA) is a cradle-to-grave methodology that can be used to


quantify the carbon footprint of a product, by assessing all environmental impacts, from
extraction of the raw materials, production processes, transportation, use and eventual
disposal of the product [14]. An LCA should be carried out, to ensure that the bio-based
flax and cork sandwich panels considered in this study are in fact a green and sustainable
alternative.

77
7 References

[1] D. Zenkert, An Introduction to Sandwich Construction, United Kingdom: CPI


Antony Rowe, 1997, pp. 1-51, 107-110.

[2] “Sandwich Panels,” Rock West Composites, [Online]. Available:


https://www.rockwestcomposites.com/plates-panels-angles/sandwich-panels.
[Accessed 13th March 2020].

[3] T. Bitzer, “Honeycomb Marine Applications,” Journal of Reinforced Plastics


and Composites, vol. 13, no. April, pp. 355-360, 1994.

[4] P. Kujala and A. Klanac, “Steel Sandwich Panels in Marine Applications,”


Brodo Gradnja, vol. 56, no. 4, pp. 305-314, 2005.

[5] L. Larsson and R. Eliasson, Principles of Yacht Design, Great Britian: McGraw-
Hill , 2000.

[6] “End-of-life boat disposal – a looming issue for the composites industry,”
Linset, 9th September 2013. [Online]. Available:
https://linset.it/it/news/scheda.php?id=71&st=1&k=End-of-life-Boat-Disposal-
Looming-Issue. [Accessed 14th March 2020].

[7] ASTM International, “C297 Standard Test Method for Flatwise Tensile Strength
of Sandwich Constructions,” United States, 2004.

[8] L. Reis and A. Silva, “Mechanical Behaviour of Sandwich Structures using


Natural Cork Agglomerates as Core Materials,” Journal of Sandwich Structures
and Materials, vol. 11, pp. 487-500, 2009.

[9] P. Sadeghian, D. Hristozov and L. Wroblewski, “Experimental and Analytical


Behavior of Sandwich Composite Beams: Comparison of Natural and Synthetic
Materials,” Journal of Sandwich Structures & Materials, 2016.

[10] T. Bitzer, Honeycomb Technology: Materials, Design, Manufacturing,


Applications and Testing, Dublin, CA: Springer Science & Business Media,
1997.

78
[11] H. G. Allen, Analysis and Design of Structural Sandwich Panels, Oxford:
Pergamon Press, 1969.

[12] A. Manalo, T. Aravinthan, A. Fam and B. Benmokrane, “State-of-the-Art


Review on FRP Sandwich Systems for Lightweight Civil Infrastructure”.

[13] J. P. Torres, R. Hoto, J. A. García-Manrique and J. Andres, “Manufacture of


Green-Composite Sandwich Structures with Basalt Fiber and Bioepoxy Resin,”
Advances in Materials Science and Engineering, 2013.

[14] R. Quarshie and J. Carruthers, “Technology Overview - Biocomposites,”


NetComposites, 2014.

[15] J. Duflou, Y. Deng, W. Dewulf and K. Van Acker, “Do fiber-reinforced


polymer composites provide environmentally benign alternatives? A life-cycle-
assessment-based study.,” Materials Research Society, vol. 37, no. April , pp.
374 - 383 , 2012.

[16] H. Mankodi, “New Reinforced Material For Textile Composite - Basalt Fiber,”
Technical Textile, [Online]. Available:
https://www.technicaltextile.net/articles/new-reinforced-material-2514.
[Accessed 22nd August 2019].

[17] D. Bannister, “materialstoday,” 18 February 2014. [Online]. Available:


https://www.materialstoday.com/composite-processing/features/an-introduction-
to-core-materials-part-1/. [Accessed 25th August 2019].

[18] “TYPES OF CORE MATERIALS,” Fibermax Composites, [Online]. Available:


https://www.fibermaxcomposites.com/shop/index_files/typesofcore.html.
[Accessed 25th August 2019].

[19] “Nomex™ Honeycomb,” Core Composites, [Online]. Available:


https://www.corecomposites.com/products/honeycomb/nomex-
honeycomb.html. [Accessed 25th August 2019].

79
[20] “PVC Foam,” Composites One, [Online]. Available:
https://www.compositesone.com/product/core-materials/pvc-foam/. [Accessed
25th August 2019].

[21] L. Gil, “Cork Composites: A Review,” Materials, vol. 2009, no. 2, pp. 776 -
789, 2009.

[22] O. Castro, J. Silva, T. Devezas, A. Silva and L. Gil, “Cork Agglomerates as an


Ideal Core Material in Lightweight Structures,” Materials and Design, 2009.

[23] T. Garrison, A. Murawski and R. Qui, “Bio-Based Polymers with Potential for
Biodegradability,” Polymers, vol. 14, p. July, 2016.

[24] “Epoxy resins – an introduction,” Materials Today, 10th June 2012. [Online].
Available: https://www.materialstoday.com/composite-
processing/features/epoxy-resins-an-introduction/. [Accessed 28th August
2019].

[25] “Improve your Product Sustainability Performance,” GaBi-Software, [Online].


Available: http://www.gabi-software.com/international/index/. [Accessed 16th
September 2019].

[26] R. Ferreira, D. Pereira, A. Gago and J. Proenca, “Experimental characterisation


of cork agglomerate core sandwich panels for wall assemblies in buildings.,”
Journal of Building Engineering, vol. 5, pp. 194-210, 2016.

[27] N. Lakreb, B. Bezzazi and H. Pereira, “Mechanical behavior of multilayered


sandwich panels of wood veneer and a core of cork agglomerates.,” Materials &
design, pp. 627-636, January 2015.

[28] P. Sadeghian, L. Wroblewski and D. Hristozov, “Long-term tensile properties of


natural fibre-reinforced polymer composites: comparison of flax and glass
fibres,” Composites Part B: Engineering, vol. 95, pp. 82-95, 2015.

[29] K. Mak, A. Fam and C. MacDougall, “Flexural Behavior of Sandwich Panels


with Bio-FRP Skins Made of Flax Fibers and Epoxidized Pine-Oil Resin,”
Journal of Composites for Construction, vol. 19, no. 6, 2015.

80
[30] ASTM International, “D7249 Standard Test Method for Facing Properties of
Sandwich Constructions by Long Beam Flexure,” United States, 2006.

[31] ASTM International, “C393 Standard Test Method for Core Shear Properties of
Sandwich Constructions by Beam Flexure,” United States, 2009.

[32] ASTM International, “C365 Standard Test Method for Flatwise Compressive
Properties of Sandwich Cores,” United States, 2003.

[33] ASTM International, “C364 Standard Test Method for Edgewise Compressive
Strength of Sandwich Constructions,” United States, 1999.

[34] R. M. Jones, Mechanics of Composite Materials, New York: Taylor & Francis
Group, 2010, p. 12.

[35] D. Camilleri, “Micromechanical Behaviour of a Lamina,” in Engineering


Analysis of Composite Structures, University of Malta, 2019, pp. 14-21.

[36] Kelly, “The Elementary Beam Theory,” [Online]. Available:


http://homepages.engineering.auckland.ac.nz/~pkel015/SolidMechanicsBooks/P
art_I/BookSM_Part_I/07_ElasticityApplications/07_Elasticity_Applications_04
_Beam_Theory.pdf. [Accessed 18th October 2019].

[37] A. Petras, Design of Sandwich Structures, Cambridge: Robinson College, 1998.

[38] R. C. Hibbeler, Mechanics of Materials, 9th ed., Prentice Hall, 2014, p. 577.

[39] K. Shivakumar and H. Chen, “Structural Performance of Eco-Core Sandwich


Panels,” in Major Accomplishments in Composite Materials and Sandwich
Structures, Springer, 2009, pp. 381-407.

[40] E. Azzopardi, “Mechanical Testing of Sandwich Panels,” 2012.

[41] L. A. Carlsson and G. A. Kardomateas, “Structural and Failure Mechanics of


Sandwich Composites,” in Solid Mechanics and its Applications, Springer, p.
11.

81
[42] L. Fagerberg, “Wrinkling of Sandwich Panels for Marine Applications,”
Stockholm, 2003.

[43] ASTM International, “D7250 Standard Practice for Determining Sandwich


Beam Flexural and Shear Stiffness,” United States, 2012.

[44] “Mechanical Testing of Sandwich Panels,” HEXCEL, United States, 2007.

[45] N. A. Fleck and I. Sridhar, “End Compression of Sandwich Columns,”


Composites, pp. 353-359, 2001.

[46] S. Mallaiah, K. Vinayak Sharma and M. Krishna, “Development and


Comparative Studies of Bio-based and Synthetic Fiber Based Sandwich
Structures,” International Journal of Soft Computing and Engineering (IJSCE),
vol. 2, no. 1, p. 332, March 2012.

[47] M. L. Jay, “Biocomposites Encourage Innovative Applications,” Composites


Manufacturing, 1 September 2018. [Online]. Available:
http://compositesmanufacturingmagazine.com/2018/09/biocomposites-
encourage-innovative-applications/. [Accessed 7th February 2020].

[48] “200g 2x2 Twill Biotex Flax Fibre 1590mm Wide,” Easy Composites, [Online].
Available: https://www.easycomposites.co.uk/#!/fabric-and-
reinforcement/natural-flax-and-jute-fibres/dry-flax-and-jute-fibres/flax-22-
200g.html. [Accessed 11th February 2020].

[49] J. Zhu, H. Zhu, J. Njuguna and H. Abhyankar, “Recent Development of Flax


Fibres and Their Reinforced Composites Based on Different Polymeric
Matrices,” Materials, vol. 6, no. 11, pp. 5171-5198, 2012.

[50] “CoreCork NL20`,” Core Composites, [Online]. Available:


https://www.corecomposites.com/suppliers/amorim-cork-composites/corecork-
nl20.html. [Accessed 11th February 2020].

[51] “ONE – High Biobased Laminating Epoxy,” Entropy Resins, [Online].


Available: https://eu.entropyresins.com/product/one-high-biobased-laminating-
epoxy/. [Accessed 11th February 2020].

82
[52] “Why Super Sap,” Entropy Resins, [Online]. Available:
https://entropyresins.com/why-use-super-
sap/?doing_wp_cron=1581456449.4830420017242431640625. [Accessed 11th
February 2020].

[53] “Vacuum Bagging Equipment and Techniques for Room-Temp Applications,”


Fibre Glasst, [Online]. Available: https://www.fibreglast.com/product/vacuum-
bagging-equipment-and-techniques-for-room-temp-
applications/Learning_Center. [Accessed 13th February 2020].

[54] E. J. Barbero, Introduction to Composite Materials Design, USA: Taylor and


Francis Group, 1999.

[55] “Vacuum Bagging,” Net Composites, [Online]. Available:


https://netcomposites.com/guide/repair/vacuum-bagging/. [Accessed 13th
February 2020].

[56] R. Brown, Handbook of Polymer Testing: Physical Methods, CRC Press, 1999,
p. 131.

[57] “Mechanical Behaviour of Materials,” EuroMotor, [Online]. Available:


https://www.euromotor.org/mod/resource/view.php?id=20521. [Accessed 9th
April 2020].

[58] J. Arbaoui, “Effect of Core Thickness and Intermediate Layers on Mechanical


Properties of Polypropylene Honeycomb Multi-Layer Sandwich Structures,”
Archives of Metallurgy and Materials , vol. 59, no. 1, p. 16, 2014.

[59] J. Lai, G. Mclntyre, D. Walczyk and R. Bucinell, “A New Approach to


Manufacturing Biocomposite Sandwich Structures: Mycelium-Based Cores,” in
Proceedings of the ASME 2016 International Manufacturing Science and
Engineering Conference, Blacksburg, Virginia, USA, 2016.

[60] W. A. Samad, A. A. Warsame and A. Khan, “An Experimental Study on the


Edgewise Compressive Failure of Paper Honeycomb Sandwich Panels with
Respect to Various Aspect Ratios,” Dubai, United Arab Emirates, 2018.

83
[61] F. Sarasini, J. Tirillo, L. Lampani, T. Valente, P. Gaudenzi and C. Scarponi ,
“Dynamic Response of Green Sandwich Structures,” Procedia Engineering,
vol. 167, pp. 232 - 244, 2016.

[62] D. Wang, “Cushioning properties of multi-layer corrugated sandwich


structures,” Journal of Sandwich Structures and Materials , vol. 11, no. 1, pp.
57 - 66, 2009.

[63] S. Surendran and K. Kumar, “Design and analysis of composite panel for
impact loads in marine environment,” Ships and Offshore Structures, vol. 8, no.
5, pp. 597 - 606 , 2013.

[64] W. Fricke, “Fatigue and Fracture of Ship Structures,” in Encyclopedia of


Maritime and Offshore Engineering, Hamburg University of Technology
(TUHH), Hamburg, Germany, John Wiley & Sons, Ltd., 2017.

[65] “Cork Wood Properties,” The Falcon Eye, [Online]. Available:


https://thefalcon-eye.com/common-wood-materials-properties/cork-wood-
general-mechanical-and-thermal-properties/. [Accessed 20th February 2020].

84
8 Appendix – Material Datasheets

8.1 Biotex® 2/2 twill 200 g flax fibre [48]

85
86
8.2 Amorim® CoreCork NL20 [50]

87
8.3 ONE Super Sap® High Bio-Based Laminating Epoxy [51]

88

You might also like