You are on page 1of 115

Fabrication and Characterization of Bio-based Polyurethane

Foams Reinforced with Natural Fibers

By

Syed Muhammad Husainie

A thesis submitted in conformity with the requirements for the degree of Master of Applied
Science
Mechanical and Industrial Engineering University of Toronto

© Copyright by Syed Muhammad Husainie 2020


Fabrication and Characterization of Bio-based Polyurethane Foams
Reinforced with Natural Fibers
Syed Muhammad Husainie
Master of Applied Science
Mechanical and Industrial Engineering
University of Toronto
2020

Abstract
With the world evolving towards the use of polymers, there has been an increasing demand to
develop high performance, eco-friendly polyurethane (PU) foams for sustainable
manufacturing. Herein, three novel studies were carried out to fabricate bio-based PU with
natural fillers. The goal of these studies was to enhance the mechanical properties of foam
while maintaining high quality and scalability. Initially, natural fillers including chitin, chitosan,
lignin, polysaccharide, and hazelnut were embedded into free-rise PU foam to improve the
mechanical properties. Consequently, this study helped identify the ideal formulation
techniques to fabricate molded foam plaques embedded with hazelnut, chitin, cellulose, and
eggshell. These composites were tested against the pure formulation to measure
improvements in properties. They were characterized using mechanical, thermal, and
morphological analyses to measure the scale-up potential of natural fiber-based PUs. In the last
study, alternative reagents and formulation adjustments were made to see how they impact
foam properties.

ii
Acknowledgments
I’d like to start by thanking my parents for their endless support throughout my masters. I
would also like to thank my uncle, aunt, brother, cousins, and friends, and without them I
wouldn’t be here.
I would like to thank my supervisor, Prof. Hani Naguib who provided me with this exciting and
incredible research opportunity. I am truly grateful for the guidance and advice he offered me
throughout my masters. It has been a pleasure working with you and being mentored in the
diverse field of high-performance polymers by you.
Lastly, I would like to acknowledge Evoco Ltd. for their financial support and services during my
masters. I would also like to thank members of the SAPL lab. It was an amazing experience
working with you all. I am thankful to all of you for making these two years so rememberable
and enjoyable. I’d like to give a special mention to Shahriar Ghaffari, Muhammad Anwer, and
Zia Saadatnia for welcoming me to the lab and guiding me through my research studies
whenever I needed help.

iii
Table of Contents
Abstract ................................................................................................................................................... ii

Acknowledgments ................................................................................................................................... ii

Table of Contents ....................................................................................................................................iv

List of Figures.......................................................................................................................................... vii

List of Tables .............................................................................................................................................x

........................................................................................................................................ 1

Introduction............................................................................................................................................. 1

1.1 Preface ............................................................................................................................................... 1

1.2 Problem Statement & Motivation ..................................................................................................... 2

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

1.4 Thesis Organization ........................................................................................................................... 3

........................................................................................................................................ 5

Background & Literature Review ............................................................................................................. 5

2.1 Introduction ....................................................................................................................................... 5

2.2 Foaming Technology.......................................................................................................................... 7

2.3 Environmental Impact of Foaming .................................................................................................... 8

2.4 Background on Thermoset Polyurethane Foaming ........................................................................... 9

2.5 Fillers and Fibers for Polymer Composites ...................................................................................... 15

2.6 Literature Review of Natural Fillers Incorporated into Polyurethane Foam ................................... 16
2.6.1 Eggshell waste for polyurethane synthesis .............................................................................................. 16
2.6.2 Walnut and hazelnut shells for polyurethane synthesis .......................................................................... 18
2.6.3 Soybean oil based rigid polyurethane (SBOP) foam ................................................................................. 20
2.6.4 Natural fibre based polyurethane micro foams ....................................................................................... 21
2.6.5 Micro fibrillated cellulose to synthesize polyurethane foams ................................................................. 22
2.6.6 Lignin based rigid polyurethane foam reinforced with pulp fiber ............................................................ 23

iv
2.6.7 Glycerol and cellulose fiber modified water-blown soy polyol-based polyurethane foams .................... 24
2.6.8 Incorporation of hemp fiber in polyurethane .......................................................................................... 25
2.6.9 Potential of flax fiber as natural filler ....................................................................................................... 26
2.6.10 Potential of kenaf fiber as natural filler.................................................................................................. 26

2.7 Conclusion ....................................................................................................................................... 27

...................................................................................................................................... 28

A Comparative Study on the Mechanical Properties of Different Natural Fiber Reinforced Free-rise
Polyurethane foam Composites ............................................................................................................ 28

Abstract ................................................................................................................................................. 28

3.1 Introduction ..................................................................................................................................... 29

3.2 Experimental ................................................................................................................................... 34


3.2.1 Materials selection ................................................................................................................................... 34
3.2.2 Preparation and analyses of natural fillers used for polyurethane composites ....................................... 34
3.2.3 Polyurethane composite synthesis........................................................................................................... 35
3.2.4 Characterization of polyurethane foam composites ................................................................................ 35

3.3 Results and Discussion..................................................................................................................... 37


3.3.1 Morphology of polyurethane foam composites....................................................................................... 37
3.3.2 Mechanical properties of polyurethane foam composites ...................................................................... 40
3.3.3 Thermal analysis using Differential Scanning Calorimetry (DSC).............................................................. 47

3.4 Conclusions ...................................................................................................................................... 50

...................................................................................................................................... 51

Natural fibers as reinforcement for closed-molded polyurethane foam plaques: evaluating the
mechanical, morphological, and thermal properties ............................................................................ 51

Abstract ................................................................................................................................................. 51

4.1 Introduction ..................................................................................................................................... 52

4.2 Experimental ................................................................................................................................... 55


4.2.1 Materials selection ................................................................................................................................... 55
4.2.2 Preparation of natural filler for addition in polyurethane ....................................................................... 55
4.2.3 Steps to fabricate polyurethane composites............................................................................................ 58

v
4.2.4 Characterization of rigid polyurethane foams.......................................................................................... 58

4.3 Results and Discussion..................................................................................................................... 60


4.3.1 Mechanical testing of polyurethane foam composites ............................................................................ 60
4.3.2 Thermal analyses of polyurethane foam composites .............................................................................. 70
4.3.3 Morphological analyses of polyurethane foam composites .................................................................... 72

4.4 Conclusions ...................................................................................................................................... 75

...................................................................................................................................... 76

Influence of reagents and additives on the mechanical properties of closed-molded polyurethane


foams ..................................................................................................................................................... 76

Abstract ................................................................................................................................................. 76

5.1 Introduction ..................................................................................................................................... 76

5.2 Experimental ................................................................................................................................... 78


5.2.1 Materials selection ................................................................................................................................... 78
5.2.2 Preparation and analyses of natural filler for polyurethane fabrication .................................................. 78
5.2.3 Procedure for polyurethane foam composite synthesis .......................................................................... 79
5.2.2 Characterization of rigid polyurethane foam composites ........................................................................ 79

5.3 Results and Discussion..................................................................................................................... 81

5.4 Conclusions ...................................................................................................................................... 85

...................................................................................................................................... 86

6.1 Conclusions and Future Work ......................................................................................................... 86


6.1.1 Concluding remarks .................................................................................................................................. 86
6.1.2 Fundamental scientific contribution ........................................................................................................ 87
6.1.3 Future work .............................................................................................................................................. 88

References .................................................................................................................................... 90

vi
List of Figures
Figure 1.1: Distribution of polyurethane global consumption in 2016 .......................................... 1
Figure 2.1: Schematic representation of polyurethane foam matrix (in red) with air bubbles
trapped inside ................................................................................................................................ 6
Figure 2.2: Property of foam materials used in different applications......................................... 11
Figure 2.3: (a) Easy gas transfer pathway due to large cells, (b) Difficult gas transfer pathway
due to high cell density ................................................................................................................ 12
Figure 2.4: Stages of foam synthesis ............................................................................................ 13
Figure 3.1: Chemical structure of (a) chitin [11], (b) chitosan [11], (c) hazelnut [22], (d) lignin
[23], (e) polysaccharide [24] ......................................................................................................... 31
Figure 3.2: Steps for Polyurethane foam Composites .................................................................. 35
Figure 3.3: Scanning Electron Microscope (SEM) images for polyurethane composites with 1%,
2.5%, and 5% fiber loading ........................................................................................................... 39
Figure 3.4: Mechanical results for polyurethane foam composites displaying (a) Tensile strength
of each composite for 1%, 2.5%, and 5% fiber loading, (b) Ultimate elongation % of each
composite for 1%, 2.5%, and 5% fiber loading, (c) Tensile strength of PU composites after aging
for 1%, 2.5%, and 5% fiber loading (d) Ultimate elongation % of each composite after aging for
1%, 2.5%, and 5% fiber loading .................................................................................................... 45
Figure 3.5: Mechanical results for polyurethane foam composites displaying (a) Tear strength of
each composite for 1% fiber, 2.5%, and 5% fiber loadings, (b) Ball drop resiliency % of each
composite for 1%, 2.5%, and 5% fiber loading, (c) Density of PU composites after aging for 1%,
2.5%, and 5% fiber loading (d) Hardness of each composite after aging for 1%, 2.5%, and 5%
fiber loading ................................................................................................................................. 46
Figure 3.6: DSC heat-cool cycle for (a) Chitin (b) Chitosan (c) Lignin (d) Polysaccharide (e)
Hazelnut at 1%, 2.5%, and 5% fiber loadings ............................................................................... 49
Figure 4.1: Chemical structure of (a) hazelnut [38] (b) cellulose [39], (c) chitin [40], (d) eggshell
[41] ............................................................................................................................................... 56
Figure 4.2: Reactions of natural fillers involved in polyurethane foaming ................................... 57
Figure 4.3: Steps to fabricate polyurethane foam composite ...................................................... 58

vii
Figure 4.4: Tensile strength of polyurethane foam composites before aging for (a) 1% filler, (b)
2.5% filler, (c) 5% filler .................................................................................................................. 62
Figure 4.5: Tensile strength of polyurethane foam composites after aging for (a) 1% filler, (b)
2.5% filler, (c) 5% filler .................................................................................................................. 62
Figure 4.6: Ultimate elongation of polyurethane foam composites before aging for (a) 1% filler,
(b) 2.5% filler, (c) 5% filler ............................................................................................................ 64
Figure 4.7: Ultimate elongation of polyurethane foam composites after aging for (a) 1% filler,
(b) 2.5% filler, (c) 5% filler ............................................................................................................ 64
Figure 4.8: Young’s modulus of polyurethane foam composites before aging for (a) 1% filler, (b)
2.5% filler, (c) 5% filler .................................................................................................................. 65
Figure 4.9: Young’s modulus of polyurethane foam composites after aging for (a) 1% filler, (b)
2.5% filler, (c) 5% filler .................................................................................................................. 65
Figure 4.10: Ball drop resiliency % of polyurethane foam composites for (a) 1% filler, (b) 2.5%
filler, (c) 5% filler........................................................................................................................... 67
Figure 4.11: Split tear results measuring the tear loading strength of polyurethane foam
composites for (a) 1% filler, (b) 2.5% filler, (c) 5% filler ............................................................... 68
Figure 4.12: Compression set% of polyurethane foam composites measuring permanent
deformation in thickness of (a) 1% filler, (b) 2.5% filler, (c) 5% filler ........................................... 69
Figure 4.13: Hardness of polyurethane foam composites for (a) 1% filler, (b) 2.5% filler, (c) 5%
filler .............................................................................................................................................. 70
Figure 4.14: Differential Scanning Calorimetry (DSC) results used to measure the glass transition
temperature of polyurethane foam composites .......................................................................... 71
Figure 4.15: Thermogravimetric Analysis (TGA) results measuring the thermal degradation of
polyurethane foam composites ................................................................................................... 71
Figure 4.16: Ultrapycnometer results measuring the porosity% of polyurethane foam
composites(a) 1% filler, (b) 2.5% filler, (c) 5% filler ...................................................................... 73
Figure 4.17: Average cell size of polyurethane foam composites measured using Scanning
Electron Microscope (SEM) .......................................................................................................... 73
Figure 4.18: Scanning Electron Microscope (SEM) of polyurethane foam composites ................ 74

viii
Figure 4.19: Scanning Electron Microscope (SEM) images of eggshell, hazelnut, and chitin fiber
...................................................................................................................................................... 75
Figure 5.1: Polyurethane foam synthesis reaction ....................................................................... 79
Figure 5.2: Reactions involved for additives within polyurethane ............................................... 81
Figure 5.3: Mechanical properties of polyurethane foam plaques (a) Normalized tensile
strength, (b) Normalized ultimate elongation, (c) Young’s modulus, (d) Normalized tear
strength, (e) Ball drop resilience %............................................................................................... 84
Figure 6.1: Effects of additives on (a) tensile strength, (b) ultimate elongaiton %, (c) Resilience
%, (d) tear strength....................................................................................................................... 88

ix
List of Tables
Table 2.1: Thermoplastic vs. Thermoset foaming properties ......................................................... 8
Table 3.1: Average cell size ( m) of polyurethane foam composites ........................................... 40
Table 3.2: Glass transition temperature of PU foam composites ................................................ 48

x
Introduction
1.1 Preface
In recent years, there is an increasing demand for the use of polyurethane (PU) based products
as they increasingly impact day to day lives. According to an article measuring the forecasted
PU global market size, the worldwide consumption of PU is predicted to be over 79 billion USD
by 2021 [1]. The figure below shows the distribution of how PU was consumed globally in 2016
[1]:

Flexible foam
Rigid foam
18% Molded
Elastomers
31% Adhesives and sealants
3% Coatings
6% Global consumption of
PU in 2016
6%

11%
25%

Figure 1.1: Distribution of polyurethane global consumption in 2016

Despite the numerous advantages of PU products, they are also known to pose significant
environmental impacts due to bioaccumulation and harm to natural organisms. Presently, the
majority of the PU is fabricated using petroleum-based substances which are derived from coal
and oil. Not only does this pose an environmental burden, but problems also arise due to an
increase in price of petroleum and limited resources [2]. Moreover, the use of synthetic fibers
which are embedded into PU to improve mechanical properties are known to pose drastic

1
environmental impacts through microplastic ingestion [3], [4]. Consequently, the use of
renewable resources for fabricating PU foam is becoming an increasing need to match
government regulations and sustainable manufacturing practices. Additionally, natural fibers
are viewed as potential replacements to traditional ones for improving the mechanical
properties of PU as research has shown high potential of natural fibers for use in PU foam [5]
[6], [7].

1.2 Problem Statement & Motivation


Conventionally, PU foams have been synthesized with non-renewable resources including crude
oil and coal-based products. This does not only impact the environment negatively with
potential to cause environmental pollution through release of harmful gases, but also causes an
increase in prices due to the gradual depletion of resources. Moreover, traditional PU foam
properties are enhanced using synthetic fibers such as carbon fiber and glass fiber. However,
the accumulation of microplastics in the environment has given rise to toxicity and suffocation
to natural organisms through ingestion and/or irritation. Consequently, global regulations have
directed the increased need to produce more sustainable means of polymer synthesis in order
to mitigate the environmental and price constraints.

1.3 Thesis Objectives


The objectives of this project are to fabricate novel bio-based PU foams for high performance
applications. Different sources of natural fiber have been studied, in order to embed them into
PU foam and measure improvements in the mechanical properties.

The main objectives for the thesis have been divided into three components as follows:

1. The first component of the study will be to research on a range of natural fillers that can
be incorporated into PU foam. This study will serve to identify which natural fillers are
compatible with PU foam with the help of free rise foam synthesis. Targeting free rise

2
foams helps in understanding its curing behaviour over time and whether there is
scorching or shrinkage etc. taking place. Ultimately, if the filler is not compatible with
foam, collapse, low curing, or other defects arise to account for that. Moreover, free rise
foam also helps in measuring density. If the foam density is not at the required range, it
indicates that the formulation has to be adjusted.

2. Once free-rise foam synthesis has been used to study filler-matrix compatibility, foam is
fabricated using a closed mold. The foam plaques are tested according to their
mechanical, thermal, and morphological properties (similar to free rise foams) to
identify improvements compared to pure PU. Improvements seen in PU foam will be the
basis for large-scale applications such as production of consumer-based products etc.

3. PU formulation depends on the nature of polyol and isocyanate. Additionally, it also


depends on the types of reagents used, including catalyst, surfactant, blowing agent,
and cross linker. The last component of this study deals with fabricating and testing PU
with small adjustments and additions of reagents known to enhance PU properties. The
objective will be to fabricate PU using renewable resources, which gives optimal
mechanical properties.

1.4 Thesis Organization


The thesis is divided into five chapters based on the objectives mentioned in section 1.3. This
chapter provides an overview of the project scope, with more systematic discussions in
subsequent chapters.
Chapter 2 provides thorough background research on PU foams. Fundamental science of PU
production and the fillers is provided. Additionally, some similar past studies are discussed to
gain an understanding of the current developments in bio-based PUs.
In chapter 3, various natural fillers were studied in order to see whether they are compatible
with foam or not. The fillers were prepared through mechanical grinding and sieved to less that
300 microns. The resulting fibers were mixed at high rates into polyol and formulated as free

3
rise buns. Following 24-hour curing, they were visually inspected and mechanically tested to
determine improvements in properties compared to pure PU.
Chapter 4 was a more thorough study of natural fillers. Upon studying the results obtained in
chapter 4 and identifying the most promising fillers, closed-molded plaques were produced by
incorporating natural fillers into them. Filler was added into foam at weight amounts of 1%,
2.5%, and 5%. The resulting plaques were tested according to their mechanical, thermal, and
morphological properties to identify the possibility of adding natural fillers to PU for sustainable
scaled-up applications.
Chapter 5 focuses on a parametric analysis of the main factors influencing PU synthesis, with
respect to reagents. The reagents involved in PU synthesis include the catalyst, surfactant,
blowing agent, and crosslinker. Each reagent has a critical function that it offers such as
maintaining high temperature, reducing the matrix viscosity for ease of flow, and to enable gas
formation needed for foaming. The goal of this study was to modify foam chemistry to
understand how each reagent affects foam properties and obtain an optimal formulation
where the foam produced does not shrink or scorch upon curing.

4
Background & Literature Review

2.1 Introduction
Polymers have become an enticing and essential component for industrial applications due to
their structural properties. The polymer industry gained significant attention in the mid
twentieth century as a result of their light weight and wide property spectrum [8]. Over the
decades, the use of PU foam has become widespread in furniture, construction, textile,
automotive, and transportation industries. Nevertheless, a significant constraint that has been
a constant cause of concern is the depletion of resources needed for polymer synthesis.
Additionally, the bioaccumulation and persistence are other issues for PU foams which have
caused a shortage of landfill space and the need to synthesize more sustainable foam.
Use of foam products has caused the build-up of polymeric materials in the environment, which
have resulted in more rigid global regulations and criteria for a cleaner and greener
environment. Research into more eco-friendly composite materials is the focus of modern-day
studies. Amongst the activities that are involved in this, the use of natural fibers has been an
enticing field for research and testing. Natural fibers offer a high cost-performance potential
because they are not only cheaper than synthetic fibers, but they are also abundantly available
and have high specific strength [9].
Many different types of foams can be fabricated with the help of different types of precursor
chemicals. Some of the common foams available today include polyethylene foams, closed cell
PU foams, and open cell PU foams. Each foam has specific physical and chemical properties that
enable them to be used effectively in the relevant application.
Once foam has been selected, research is carried out entailing the study of foam characteristics
based on varying matrix, catalyst, or surfactant composition etc. Selecting the appropriate
materials is vital for effective foam synthesis. After material selection, the filler is carefully
analyzed by looking at the properties it exhibits. The filler must possess chemical bonds which
are compatible with matrix used for fabrication, in order to develop strong interfacial bonds.

5
Consequently, filler functionalization becomes necessary upon poor composite properties. The
filler can naturally exist with excessive functional groups and intermingled bonds which inhibit
its reactivity with the matrix. As a result, the filler is functionalized by reacting it with chemicals
which free up functional groups for reaction with the matrix [10].
Following successful fabrication of foam, extensive mechanical testing is performed on the
materials to measure tensile strength, modulus, tear strength, and other properties such as the
density of the material, hardness, resilience, and abrasion resistance. Characterization is
another major step that goes in hand with testing. The main objective of characterization is to
analyze the material structure and see how the matrix/ filler has reacted. Poor reaction is
indicated by tears and cracks throughout the matrix.
The term foam refers to gaseous voids which are dispersed within a dense continuous medium
of material [11]. The process typically takes place as free gas molecules are converted into
spherical bubbles, resulting in dissipations due to temperature and pressure variations.
Foaming is typically described as a way of transitioning from a homogeneous state to a meta-
stable or non-homogeneous state. In specific cases however, the conditions for foaming are not
adequately met, which leads to foam collapse and shrinkage. The main goal for foaming
reactions is to turn an unstable foaming reaction into a stable, valuable final product, before
gaseous bubbles can escape or cause foam shrinkage. The figure below shows the structure
following a successful foaming reaction:

Figure 2.1: Schematic representation of polyurethane foam matrix (in red) with air bubbles trapped inside

6
A critical parameter that is typically focused on is the critical bubble radius which determines
whether the bubble in the foaming reaction will stay intact or collapse. The critical bubble
radius is calculated by using the following equation:

𝑅𝑐𝑟 = 2𝜎/(𝑃𝐵 − 𝑃) 1

If the bubble radius is small compared to the critical bubble radius, it causes bubbles to cluster
and collapse. In addition to this, if the surrounding resistance to the bubble is low, buoyancy
takes over which results in the bubbles travelling up and dissipating. The nature of polyol
(dense and viscous) helps bubble nucleation by stifling it in the surrounding viscous phase. This
in turn helps provide the bubble with a longer lifespan, giving the polymer an opportunity of
curing into a permanent foamed structure.

2.2 Foaming Technology


A polymer is typically defined by long chained monomers combined together with the help of
functional groups and van der Waals forces. Depending on the network structure and thermal
variations, the polymer is categorized into thermoplastics and thermosets. In thermosets,
elevated temperatures result in the formation of a network structure which is no longer
thermo-reversible. However, in the case of thermoplastics, these polymers are capable of
transitioning between solid to liquid states at specific temperature limits [8]. This is because
thermoplastics possess chain lengths which are held together by van der Vaal forces and chain
entanglements which enable the polymer to behave visco-elastically. The table below gives a
comparison of thermoplastic and thermoset foaming differences [13]:

7
Table 2.1: Thermoplastic vs. Thermoset foaming properties

Gas inclusion Foaming Stabilization


Thermoplastics Dissolution, decomposition Supersaturation Cooling

Thermoset Chemical reaction Gas evolution Polymerization, X-linking, cooling

While thermoplastics require different parent polymers in order to inherit specific properties, a
wide range of properties can be obtained using the same polymer for thermosets. Processes
such as extrusion and injection molding are used to produce thermoplastic PU foams, whereas
polymerization chemical reactions are utilized in order to produce thermosets [8].
Different sets of foaming techniques are utilized in industries at present times. Among them,
reactive foaming makes up approximately 58% of the foaming market [11]. This is the process
incorporated for this thesis, involving reactions such as the addition reaction between polyol
and isocyanate in order to produce PU foam. The other reagents added to the reaction,
including the blowing agent, catalyst, surfactant etc. help influence and obtain foam with
properties such as low or high density, hardness, resiliency %.
Another foaming process is “soluble foaming” which involves the addition of a physical blowing
agent into the polymer through the process of extrusion and injection molding [8]. A screw
embedded into the machine is designed to create a high-pressure atmosphere which helps
achieve homogenous mixing of the gas, followed by a lower pressured environment for the
foaming reaction. Saturating a polymer bead to enable gas expansion is another method which
is commonly used to obtain foam products, and it is typically incorporated with closed molded
foaming to obtain the final product.

2.3 Environmental Impact of Foaming


Since the early 20th century, blowing agents have been commonly used in order to fabricate
foams. CFCs and HCFCs were widely used early on [12], [13], [14] due to their effective
properties in foam fabrication. However, subsequent research showed that stable CFCs and
HCFCs release chlorine gas upon exposure to ultraviolet light (UV) which resulted in the

8
depletion of ozone layer and global warming. Consequently, renewable means of making PU
foam with the help of environmentally friendly blowing agents has been an important goal in
present times. The main challenge with this is that the nature of the blowing agent needs to be
very specific e.g. large molecular weight with low diffusivity. Other challenges with blowing
gases besides CFCs are that they have high heat capacities and low energy efficiencies which
impede successful foam synthesis. The quintessential goal for selecting an ideal blowing agent
involves forming a foam which has ideal material density as well as the required performance
characteristics. As a result, more research and testing of PU foams with different additives is
being carried out, with an objective of fabricating high-quality PU foams that are more
sustainable for the environment.
Aside from reagents involved, eco-friendly composite materials that can degrade on their own
when exposed to the natural environment is also the focus in current research. Natural fibers
offer a high cost-performance potential because they are not only cheaper and more
sustainable compared to synthetic fibers, but also because they are abundantly available with
high specific strengths [9]. Replacement of synthetic fibers will reduce the toxicity to living
organisms through the bioaccumulation and persistence of microplastics.

2.4 Background on Thermoset Polyurethane Foaming


PU fabrication was discovered in the late 1930s. The initial development focussed on PU
products obtained from aliphatic di-isocyanate and diamine forming polyurea. The aliphatic
groups consist of carbon groups that form open bonds. In the following years, with more
research and development, the use of glycol along with isocyanate was realized. The glycol
group consists of hydroxyl bonds that are able to react with the -NCO bonds of isocyanate,
creating PU linkage. Additionally, different types of polyols are available which help create PU
with different properties. For instance, polyester based polyols were discovered which form
different structural configurations compared to polyether-based polyols. In present times,
polyether- based polyols are replacing polyester-based polyols because of their low cost, ease
of handling, and improved hydrolytic stability [9].

9
The main reaction that takes place for foam fabrication is the reaction between isocyanate and
polyol. The -NCO group of isocyanates oxidizes the -OH group of the polyol resulting in the
urethane linkage as shown in the reactions below. The type of isocyanate influences the
properties of the final PU foam that is fabricated. For instance, selecting PMDI as the isocyanate
helps result in good thermal resistance and flame retardance, or using aromatic isocyanate can
help produce rigid PU [9] .
Polyol is another essential component for fabricating PU. It consists of hydroxyl groups and
other functional groups such as ester, ether, amide, and acrylic bonds that help achieve PU of
various functionalities. Using highly branched polyester polyol (PEP) results in rigid PU with
good heat and chemical resistance [16] whereas less branched PEP results in PU with good
flexibility (at low temperature) and low chemical resistance. Subsequently, using low molecule
weight polyols produces rigid PU while high molecular weight long chain polyols yield flexible
PU. Nevertheless, PEP is susceptible to hydrolysis due to the ester groups, which causes
deterioration of the mechanical properties. Nevertheless, this disadvantage can be overcome
with the help of carbodiimides which prevent significant breakdown of the PEP compounds due
to reaction with water [17].
Other polyols include polyether polyols (PETP) which are less expensive than PEP. PETP is
produced through the reaction of ethylene or propylene oxide + alcohol/ amine + acid/base
catalyst to give PETP and other products. PU fabricated with this polyol shows high moisture
permeability and low Tg, which makes them good for applications such as coatings and paints
[18].
In addition to isocyanate and polyol, additives also play a vital role to make PU foam. This
includes the catalyst, which helps speed up the foaming reaction by reducing the activation
energy needed to initiate the reaction effectively. In turn, the curing time is also reduced. Some
of the common types of catalysts include several aliphatic and aromatic amines, organometallic
compounds, or alkali metal salts [19].
Due to their vastly enticing characteristics, PU foams have become the center for much
attention and research. PU plastics have opened doors to unique advantages as a result of their
enhanced strength to weight ratio, and malleability with respect to properties. These diverse

10
properties have been the main driving force for further research and development in order to
improve foaming technologies. Through the years, many different techniques of manufacturing
foam have been employed, including extrusion foaming, reactive foaming, injection molding,
oven heat foaming, etc. With the help of advanced fabrication techniques, novel and unique
foam have been produced, such as rigid PU foam and flexible PU foam. The steady growth in PU
foams has resulted in their successful replacement in different wooden and metal product
applications [11]. The figure below shoes the wide property ranges for different foam materials
used in products [20], [21]:

7 3

Log Thermal Conductivity (W/moK)


Log Young's Modulus (MN/m2)

5 2

3 1

1 0

-1 -1

-3 -2

-5 -3
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
PU foam

PU foam
PS foam

PS foam
Steel

Steel
SIC

SIC
Silica

Silica
Al

Al
Gold

Gold
PU

PU
PS

PS
LDPE

LDPE
Cook

Cook
Air

Air

Log Material and their Density (kg/m2) Log Material and their Density (kg/m2)

Figure 2.2: Property of foam materials used in different applications

The nature of PU fabricated depends on several factors due to the vast reagent and foaming
options. Characteristics are determined through structural parameters including cell density,
cell expansion, distribution, and open cell content. The cells formed are controlled by the
nature of polymer used and the functional groups present, and so, manipulating these
parameters helps to achieve foam with the desired functionality [22], [23]. Manipulating the
polymer chemistry can help distinguish and obtain properties such as higher toughness, lower
density, high tensile strength, low hardness, etc. Fabricating foam with high cell size helps in
synthesizing foam which has better energy absorption and heat convection properties. On the
other hand, maintaining high cell density helps in the formation of harder foam with more

11
polymer dependent heat transfer. Similarly, a weak cell wall (which gives rise to open cell
structures) results in the reduction of the foam’s mechanical properties but provides good fluid
absorption features [11].
Another important additive for foams is the blowing agent which helps to produce a cellular
structure via the foaming process. Blowing agents such as water helps form gas during the
reaction which gets trapped within the foam cells. This increases the thermal and acoustic
insulation, in addition to the relative stiffness of the foam. [19] studied the addition of hemp
filler and the addition of water as a blowing agent helped to increase the number of cells in the
foam. This was a result of higher carbon dioxide gas evolving from the reaction through water
addition, and more crosslinking because of fiber addition. Blowing agent added to the foaming
reaction can serve a wide range of effects on the properties of foam. Although any gas can
serve as a blowing agent, each one has a unique nature and may not be easily compatible with
foam. The blowing agent controls different properties of foam such as the thermodynamic &
transport properties amongst other applications. It controls the different states of foam
synthesis, from raw materials to cooling the polymer matrix. The following figure shows how
the travel pathway of gas is inhibited due to higher cell count, a consequence of effective
blowing agent and high closed cell content:

a b
Figure 2.3: (a) Easy gas transfer pathway due to large cells, (b) Difficult gas transfer pathway due to high cell density

12
The steps include implementation of gas into the polymer during mixing, homogenization,
foaming, and finally evacuation of the blowing agent and replacement with air. The following
figures show this process [8]:

Matrix Gas Elastomer Elastomer


(Matrix + Gas) (Matrix + Air)
Volume

Addition of Mixing Foaming Air replacing


gas through
components Reaction permeation
for foaming
reaction

Figure 2.4: Stages of foam synthesis

Research has shown how the foaming process has significant levels of variability. This variability
has sparked the need for understanding the foaming process at a deeper level and reducing the
variable nature of foam formulations. Below show some of the reactions during PU fabrication:

13
1. Reaction of polyol and isocyanate:

2. Reaction of polyether-based polyol and isocyanate:

3. Reaction of polyester-based polyol and isocyanate:

A wide range of foams can be fabricated with the help of precursor chemicals, polymer,
processes etc. Some of the common foams available today include polyethylene foam, closed
cell PU foams, and open cell PU foams. Each foam has specific physical and chemical properties
that enable it to be used effectively in relevant applications. Once the foam has been selected,

14
its characteristics are studied based on varying matrix, catalyst, or surfactant composition etc.
Each reagent can have a significant impact on foam properties. Moreover, even with exact
formulation and mixing, foaming reactions can have large deviations due to the randomness of
the reaction. As a result, thorough testing and repetition of the foam chemistry is vital to
ensure effective foaming. After material selection, the filler is carefully selected by analyzing
the properties it exhibits. The filler must possess chemical bonds that have an affinity to react
well with the matrix or have properties (such as hydrophobicity) which can improve the overall
hydrolytic stability and other targeted properties. Although the filler can naturally exist with
excessive functional groups intermingled together, which inhibited its reactivity with the
matrix, treatment with chemicals and functionalization can improve matrix-filler interactions
and composite properties [10].
Despite environmental regulations, the use of foamed products has increased dramatically
compared to other polymers over the past decades. The wide range of properties and ease of
processing make foams ideal to meet large scale customer demands. Since the mid 20th
century, foam has been widely used in industrial applications such as packing, automotive,
construction, textile, sports, and medical industries etc. [11].

2.5 Fillers and Fibers for Polymer Composites


Fillers are commonly used in materials to reinforce their structural integrity and provide them
with higher mechanical strength. The filler is targeted to maintain effective interfacial bonding
with the matrix which results in a stronger material. Additionally, the filler helps impart
significant structural support to the composite material, giving it high durability. There are
many different types of fillers that are incorporated into polymers in the present market.
Although synthetic fibers such as carbon, steel, and glass fibers provide high strength, the
growing need for more eco-friendly products has increased interest for research and
development of natural fiber based composite materials. Natural fibers are abundantly
available, and they also provide high specific strength. Moreover, density of natural fibers are
of the same order as those of plastics and almost half of glass fibers [9]. This makes them
suitable alternatives to traditional fillers, due to light weight and high strength. The mechanical

15
properties of fibers are determined through cellulose content and microfibrillar angle.
Additionally, the degree of polymerization also has an impact on the mechanical properties of
the composite [9]. Natural fibers with the desired functional groups are important for effective
matrix filler interactions. For instance, cellulose and polysaccharide contain hydroxyl groups
which may react with the isocyanate to form strong interfacial bonds during the polymerization
reaction. Some of the common natural fibers that have been used in the past as filler include
banana fiber, jute, hemp, coir, flax, kapok fabric, and ramie.
An important aspect to consider when working with fibers is their successful mixing into the
composite. A conventional method for producing fiber reinforced composites is through
structural reaction injection molding (SRIM). This method involves embedding chopped fibers
into the mold to mix with the matrix. Another method involves spraying the reinforcement with
the matrix system followed by compression molding. However, these methods are known to be
expensive, and so, there is a need for more cost- effective methods, such as the long fiber
injection process developed by Krauss Maffei Technologies GmbH [9] which provides effective
cost-performance benefit process for fabrication.

2.6 Literature Review of Natural Fillers Incorporated into Polyurethane


Foam
2.6.1 Eggshell waste for polyurethane synthesis
PU foams exhibit a wide range of valuable properties including low apparent density, good
mechanical properties, high weatherability, low thermal conductivity, and excellent damping
abilities. They are used in automotive, furniture, construction, and cosmetic industries for use
as insulating elements for construction, heat engineering, and protective packaging. [25]
focused on the fabrication of rigid PU with eggshell waste as filler. To achieve their set criteria,
raw materials were selected based on background research and testing. Natural filler (eggshell
waste) was added to foam for reinforcement, and the. Composite fabrication for eggshell
involved grinding eggshells using the centrifugal mill followed by filtering of the fillers using a
200-micrometer mesh sieve [25]. The remaining filler was dried at 105oC and 24 hours to

16
remove any traces of water and deactivate the biological functionality of the organic
substances contained within the filler. The polyol was mixed with catalyst, surfactant, water,
and ES filler. This solution was directly reacted with isocyanate after which the mixture was
stirred at 1000 rpm for 8 seconds and poured into an open mould where foam growth took
place in the vertical direction. The resulting material was annealed for 30 min at 70oC and
conditioned at 22oC at 50% relative humidity for 24 hours before being removed from the
mould. The foam was left to cure for two weeks before cutting and testing.

The composites were characterized using infrared spectroscopy, Xenotest Alfa,


Thermogravimetric analysis (TGA), Powder X-ray diffraction (PXRD), and SEM. Infrared
spectroscopy was used to analyse the chemical constitution of the RPUFs. The FTIR machine
gave a descriptive analysis of the bonds present within the foam and thus, it helped to
determine the degree of phase separation within the foam material. The aging test involved the
use of light exposure and weathering test instrument. It involved 2 hours of exposure to UV
light followed by 18 min of spraying the sample with distilled water. The process also involved
pre-set temperature and humidity conditions.

Aside from FTIR, Thermogravimetric Analyses (TGA) was used to measure the thermal
degradation of the materials used in the experiments while PXRD was applied to analyze the
crystallinity of the obtained materials. SEM helped to analyze the cellular structure of the foam
and measure cell density. Other parameters such as the compressive strength, cell density,
friability, toxicity, and bacteria adhesion were also measured.

The results showed improvements in the RPUF’s properties e.g. flame resistance, aging
resistance etc. The eggshell contains about 95% calcium carbonate and 5% organic materials
which provided improvements in the mechanical properties and thermal stability of the filler.

The results were used to confirm that the foam produced met the physical requirements of
rigid PU foam. FTIR showed the presence of -N-H groups, N=C=O groups, and C-O groups which
indicated the presence of eggshell filler and PU linkage. The following figure shows the FTIR
results for rigid PU foam with eggshell waste:

17
FTIR tests were carried out before and after the aging tests to analyze the aging process and
how it affected the foam microstructure. However, FTIR did not show a significant shift in the
intensity of the characteristic FTIR signals. However, there was a change in the sample colour
which suggested the formation of menaquinone-imide derivatives. Furthermore, looking at the
TGA analyses, the mass change helped to examine the temperature of maximum degradation
for the used polyols. This change helped identify the temperature of degradation for the filler,
matrix precursors, filler etc. The results showed almost 42% weight loss for the eggshells at a
temperature of 700oC, after the degradation process was finished, indicating high degradation
resistance of the foam.

SEM results showed that as the eggshell content increased, the average cell size decreased but
the number of cells increased. Additionally, increasing the eggshell content increased friability
of the foam but it had no significant impact on the bacteria adhesion and growth.

2.6.2 Walnut and hazelnut shells for polyurethane synthesis


[26] involved the addition of walnut and hazelnut shells for fabricating water blown PU foams.
13.5% and 18% compositions of the filler were added to foam after they were grinded using a
MUKF-10 mill. Thermogravimetric Analyses (TGA), Differential Scanning Calorimetry (DSC),
Fourier Transform Infrared Spectroscopy (FTIR), and mechanical testing were used to
characterize PU foam and the results indicated that degradation of foam in nitrogen
atmosphere occurred in two stages:

1. The first stage was between 150oC and 300oC which involved the degradation of
urethane bonds (dependent on the isocyanate and alcohol used in foam fabrication).

2. Then second stage occurred at 300oC to 400oC which involved the degradation of soft
segments in PU.

The results also indicated that adding filler into foam allowed it to have higher thermal stability.
Degradation in air showed that it occurred in four stages with three of them being in between
50-400oC. Through the TGA and DSC curves obtained in this study, it was observed that the filler

18
content for both walnut shell and hazelnut shell had no significant effect on thermal stability of
the tested materials.

Differential scanning calorimetry (DSC) and Dynamic mechanical analyses (DMA) were
performed to study the influence of walnut and hazelnut addition on PU foam structure and
properties. DSC is a technique to measure the difference in amount of heat required to increase
the temperature of a sample compared to raising the temperature of a reference material. It
also gives information on phase transition temperatures which is useful for industrial
applications where climate difference is seen. The results from DSC are a curve of heat flux
versus temperature or versus time. This curve can be used to calculate the enthalpies of
transition. In DMA a sinusoidal stress is applied and the strain in the material is measured. This
approach is used to measure the glass transition temperature of the material. Experiments
showed that hazelnut and walnut do not have significant influence on the glass transition
temperature of free rise PU foam. The storage modulus (SM) (which describes the stiffness of
the material and its ability to store energy) showed that the SM values for modified foams were
higher than the SM values obtained for neat foam.

Similarly, FTIR analyses showed that the structure of PU was not altered significantly by the
presence of WS or HS. Nevertheless, the R index and degree of phase separation (DPS)
increased with the increase of the amount of both the tested fillers while the degree of phase
mixing decreased.

When comparing the values for physical and mechanical measurement of the filler-based PU, it
was seen that adding filler material to the foam helped to increase its density compared to neat
PU, but there was not a high difference in density when the filler was 13.5% or 18%. Hardness
measurements showed that the addition of walnut or hazelnut did not have any influence on
this parameter. Additionally, flexibility measurements indicated that the addition of WS and HS
helped to increase flexibility but at 18% composition, the flexibility reduced compared to the
value at 13.5% filler.

19
2.6.3 Soybean oil based rigid polyurethane (SBOP) foam
[27] detailed on PU foam synthesis by substituting polypropylene based polyol with soybean
oil-based polyol (SBOP). The main goal of the research was to analyze the effect of adding
biodegradable polyol to fabricate foam and measure the properties of PU with and without
biodegradable PU. The experiment was carried out by producing soybean-based PU and
comparing the results to traditional, petroleum derived polypropylene based polyol. N, N-
Dimethyl cyclohexylamine was used as the catalyst in foaming along with other additives such
as n-pentane. The mixture was first homogenized using a 10mm drill with 2.8-inch diameter
mixing blade for 40-50s at a speed of 2500 rpm. PMDI isocyanate was then added and mixed
for 10 seconds, enabling the reaction to commence effectively.

Many different aspects of the foam were analyzed using mechanical testing and
characterization techniques and SEM was used to study the cell density of the foam materials
and the closed cell content of the foam. Other tests including compressive strength, DSC, and
DMA measurements were used to calculate the mechanical strength and characteristic
properties of the foam, including the glass transition temperature. Aside from these properties,
the viscoelastic property of the material was studied. Viscoelasticity of the material is the
property that exhibits both viscous and elastic characteristics when undergoing deformation.
This force is measured by using oscillatory force (stress) that is applied to the material and the
resulting displacement (strain) measured.

The results gave many discoveries for SBOP. Firstly, they showed that isocyanate conversion is
slower after 20 seconds for 100% SBOP. Additionally, foam density increased with increase in
SBOP substitution but the two surfactants that were tested did not have significant impacts on
the SBOP substituted foam. Nevertheless, soybean helped rigid PU attain smaller average cell
size which added to the useful properties of PU. It provided the foam with good insulating
performance through high closed cell content. Other important results included high resistance
to flow which showed that the resulting foam matrix was more viscous. Moreover, surfactant B
resulted in lower cell size compared to surfactant A, showing high heat resistance. Lastly, the

20
compressive strength and glass transition temperature increased with soy substitution due to
morphological improvements and thermally resistant matrix components.

2.6.4 Natural fibre based polyurethane micro foams


[28] involved the production of PU composites with woven flax and jute fabrics. Approximately
20-40 mm length fibers were used with an average of 37-20 micrometer diameter. The woven
fibre fabrics were pretreated with Tinozym AL/ Ultrawon CN- solution to dewax the textiles and
make them suitable for use in research. The fibres were dried at 80oC for 2 hours and the foam
was produced using compression molding at 60oC for 10 minutes.

Different aspects of the mechanical strength were analyzed such as the flexural strength of the
material and the flexural modulus of the material. The flexural strength measures the ability of
the material to withstand bending forces applied to its longitudinal axis whereas the modulus is
a measure of the ratio of stress to strain (the resistance during bending). An impact tensile
pendulum test was used to measure the modulus of the composite matrix using a dynamic
mechanical thermal analysis. The temperature was maintained between 25-230oC at a heating
rate of 1oC/ min and the damping wave showed that the amplitude decreased over time, as
expected.

To measure the results at high accuracy, fiber content and fraction of micro voids were
recorded during tests. The results showed that the Young’s modulus increased with increasing
fiber content, but it decreased with an increase in micro void content. Comparing flax and jute
fiber, it was shown that flax fiber-based composites were approximately 12-17% higher
respectively compared to jute fiber, mainly due to higher strength of flax fiber compared to jute
fiber. Additionally, fiber-based composites showed an increase in shear modulus by
approximately 44% for 29% fiber loading, compared to pure PU foam. For woven flax, the log
decrement after 90oC was decreased and the peak in the alpha relaxation region was also
decreased from 2.64 to 0.45 during the reinforcement effect. Increasing the micro void content
helped reduce the shear modulus and increase log decrement. After relaxation, the effect of
micro void content on the shear modulus and log decrement was negligible. Lastly, analyzing
the impact strength showed that increasing micro void content reduced force and increased

21
deflection of the foam. Additionally, increasing the filler percentage helped improve force but
reduce deflection forces. For both jute and flax fiber composites, a higher micro void content
was seen to increase the loss energy and damping index. In short, the results showed that
woven flax fiber improved and performed at a higher strength compared to the woven jute
fiber.

2.6.5 Micro fibrillated cellulose to synthesize polyurethane foams


[29] focused on the application of cellulose into PU micro and nanofibers. Cellulose was
compression molded into PU at both micro and nanoscale, in order to compare the reinforcing
effect of micro and nanosized cellulose fibrils in the PU matrix. The composites were prepared
using a high-pressure homogenizer.

The micro-fibrils were prepared as mats using a plastic mesh. Thickness of the film was varied
within the range of 50-250 micrometers by altering the cellulose concentrations in water
cellulose slurry. Similarly, nanofibers were also produced using a similar process but instead of
using a plastic mesh, a special membrane was used which made preparation more efficient and
effective. Compression molding was then used to combine the PU with fibers using a
temperature of 175oC and 100 bars for 1.5 minutes. Micro-fiber-based composites were
identified apart from nanofiber-based composites because they were visually translucent
whereas nanofiber was transparent.

The results were analyzed using mechanical testing and characterization. Visual examination,
optical microscopy, and SEM (operated at 6V) was incorporated to test the PU produced for
these tests. The fibril-based foams were prepared using 0.025% cellulose slurry at 500 bars. The
slurry was passed through the homogenizer 15, 30, 45, and 60 times. Visual inspection showed
that the change in dispersion of cellulose suspension was easily observed. Additionally, optical
microscopy showed that nanoscale fibrils were smaller and more fibrillated compared to
microscale ones (as expected).

Mechanical properties of the composites were also improved with an increase in cellulose fiber
content. This helped strengthen the hypothesis that cellulose and PU matrix are compatible due
to their hydrophilic natures. As the loading percentage of nanocomposites decreased, it

22
resulted in lower mechanical properties due to poor sample preparation dispersion of the fibers
in PU matrix at lower loading. Nevertheless, there were large improvements in the storage
modulus of the nanocomposites compared to neat PU and cellulose fiber composites. This was
mainly due to the strong hydrogen bonding between fiber and matrix, in addition to more
surface contact between the matrix and fiber. Additionally, the cellulose fibrils helped to give
more flexibility and temperature stability than neat PU at room temperature because of a
percolating network (movement and filtering of fluids though porous materials). Lastly, DMTA
testing was done to observe the behaviour of composites with increasing temperature, and it
was seen that there was no molecular interaction between the PU and cellulose.

2.6.6 Lignin based rigid polyurethane foam reinforced with pulp fiber
In this study, petroleum-based polyol was replaced with different amounts of lignin to prepare
lignin-based PU foam (LRPF). This foam was further reinforced with different weight ratios
(1,2,5%) of pulp fiber. Pulp and paper industry generate approximately 55 million tons of lignin
each year. However, lignin products are limited to approximately 2%. Rigid PU is a closed cell
structure that is typically made using MDI and polyol precursors. The aliphatic and phenolic
hydroxyl functionalities of lignin provide good reacting sites towards isocyanates. Research has
shown that PU prepared with low molecular weight lignin is more flexible than PU prepared
with high molecular weight lignin.

The lignin obtained in this research was thoroughly characterized to understand the
composition of alkaline lignin. This was done by initial treatment with 72% sulfuric acid for 1
hour at 30oC followed by dilution to 3% sulfuric acid using deionized water. The solution was
then processed at 121oC for 1 hour after which it was cooled, filtered, dried, and weighed. The
resulting structure was analyzed using chromatography and spectroscopy.

Bio foam without lignin was prepared as control following the same procedure as lignin. For the
composite, pulp fiber was added to the polyol manually and mixed for 10 mins to ensure
complete penetration. The results were based on mechanical testing and microscopic
characterization. The PNMR spectra showed that the maximum lignin content that was added
to bio foam was 37.19%. Additionally, 15% lignin did not show an appreciable difference in the

23
reaction or structure compared to the pure structure. The SP/MAS NMR spectra results helped
to show that the pulp fiber and MDI reaction took place. Characterization results showed that
the cellular shape became more inhomogeneous and less regular as the lignin content
increased, mainly due to the extra formation of large cells. Moreover, lignin did not disperse
well in the PU foam once the content of lignin reached 37.19%.

Adding lignin or pulp was also seen to drop the density and compressive strength of the
samples. Pulp fibers influenced the reactivity of the components in the system, which
ultimately affected foam expansion and density. Moreover, LRPF reinforced with increasing
pulp fiber content led to an increase in open cell content. The compressive strength for lignin-
based composite was also decreased. The main reasons for these was due to uneven mixture of
lignin and polyol because of agglomeration and due to lower amount of hydroxyl groups in
lignin, leading to low cross-linking density and compressive strength of LRPF. Analyzing the
thermal stability of these bio foams, all samples had a narrow degradable temperature range
which was around 300-400oC. The presence of pulp fiber elevated the initial decomposition
temperature. However, pure PU had a slightly higher decomposition rate compared to
composites with low lignin and pulp fibers because of more urethane linkage. Moreover, lignin
showed higher thermal stability compared to neat PU when the lignin composition is very high.
In short, adding pulp fiber generally improved the thermal stability of bio foam.

2.6.7 Glycerol and cellulose fiber modified water-blown soy polyol-based


polyurethane foams
This study involved analyzing the effects of reinforcing agents, glycerol, and refined cellulose
fiber on PU made from soybean oil. Soy based PU foams have gained importance in present
times because of growing economic and environmental constraints. Along with soy-based
polyol and isocyanate, water was added to the PU to form urea and CO2, which expanded the
matrix polymer. Additionally, materials such as glycerol were reported to reinforce foam and
improve its strength by providing strong cross linking for the PU foam and thus enhancing the
rigidity of foam systems [30].

24
The experimental procedure involved the use of Dabco which is a silicone surfactant. This
reduced the surface tension between the chemicals by aiding in the dispersion of hydrophilic
blowing agent (water) into the hydrophobic soy polyol system. Additionally, a tin catalyst was
used to accelerate the curing reaction for urethane formation. The entire process involved
mixing the ingredients for five minutes followed by the addition of isocyanate which was mixed
for another minute. In order to see the effect of cellulosic fiber on foaming, the blowing agent
was substituted by a dilute aqueous suspension of mechanically disintegrated fiber. The
concentration of the refined fiber was extremely low, and the fibers were obtained through a
combination of high shear and impact. The foams were characterized by density, FTIR, and SEM
measurements. Additionally, the rise height (used to measure the ease of foaming) and fiber
time (used to measure the time to materialize) was also recorded. The results showed the there
was a reduction in the ease of foaming due to an increase in glycerol level. There was also an
enhancement in foaming at higher NCO index. Increasing the NCO index resulted in an increase
in urea formation and decrease in urethane groups. Moreover, the fiber time at higher NCO
index was lower than the fiber time at lower NCO index because of the enhanced blowing
tendency at higher NCO index.

Results showed that adding fibers and mixing for 20 minutes increased the density of PU foam.
Urethane formation predominated over blowing with the addition of fiber. Fibers helped supply
hydroxyl functionality necessary to promote urethane formation with isocyanate. Cellulose
materials were seen to act as a source of hydroxyl functionality, modifying the properties of
foam and stiffening the PU foam.

2.6.8 Incorporation of hemp fiber in polyurethane


Hemp fiber has also been studied for addition into PU foam. The hemp filler consists of groups
that are theorized to be able to help with PU synthesis and properties. However, the difference
in hydrophilicity between the fiber and matrix is a barrier for successful fabrication of hemp-
based composites. As a result, compatibilizers and surface modification techniques such as
alkali treatment etc. have reportedly been used to increase matrix-fiber compatibility [31].

25
Another research [32], showed improvements in the surface of PU when hemp fiber was
treated verses untreated fibers. The untreated fiber resulted in a rough structure for PU,
whereas MDI-treated hemp fiber was smooth. The surface change was mainly a result of
urethane bonding between hydroxyl groups of the fiber surface and the isocyanate of the MDI.

2.6.9 Potential of flax fiber as natural filler


Flax fiber is a member of the Pinaceae family, and an important crop for several different food
sources in the world. They are also used widely in the textile industry. In comparison to other
fibers, flax fibers are readily available and because of their effective physical properties such as
moderate strength and low density [33], they have attracted a lot of attention to being used in
applications such as polymer composites, building materials, and absorbent materials [34].

2.6.10 Potential of kenaf fiber as natural filler


[35] carried out research into the use of kenaf based composites. An internal mixer was used at
a temperature of 190oC along with a hot plate. The results showed that fibers within the range
of 125-300 micrometers resulted in the best mechanical properties. Another study on kenaf
fiber showed that adding 30% of kenaf fiber to the PU composite helped exhibit good tensile
strength, and the modulus increased with increase in fiber content. The strain, impact strength,
abrasion resistance, and thermal stability of the composite decreased with increase of fiber
loading [35].

26
2.7 Conclusion
Research into the use of PU foams has shown a lot of promise and possibility to be used in
many different engineering applications. In addition to conventional PU foam, the addition of
fillers has also helped improve its properties while adding cost-performance benefits to it.
Different forms of testing were used to compare the properties of conventional foam with
filler-based foam. The results included mechanical testing in addition to characterization.
Mechanical tests helped to analyze local and global mechanical strength of the material
through experiments whereas characterization helped to analyze PU foam by providing
analyses on the chemical bonding and microstructure of foamed composites. The foam industry
continues to expand in present times due to their vast properties and improvements in terms of
cost benefit. As non-renewable resources and microplastics continue to harm living organisms,
the world progresses towards the development of high performance, high comfort, and eco-
friendly products; to meet the increasingly necessary and critical environmental regulations,
that are needed for a sustainable future.

27
A Comparative Study on the Mechanical Properties of Different Natural
Fiber Reinforced Free-rise Polyurethane foam Composites

Abstract
The goal of this study was to improve tailored mechanical properties of foam by incorporating
different filler types. Flexible polyurethane foam thermosets with varying wt.% of natural fibers
were tested and characterized according to their mechanical, morphological, and thermal
properties. The fillers used included lignin, chitin, chitosan, hazelnut shells, and polysaccharide,
and the results were investigated to compare the effects of fillers on the foam properties. A
morphological analysis showed good overall dispersion of chitin and hazelnut fillers, and
consequent improvements in tensile and tear strength of polyurethane foam. Additionally,
these fillers helped increase resiliency while reducing foam hardness. Polysaccharide showed
similar improvements in tensile results, while the tear strength decreased. However, lignin and
chitosan reduced foam mechanical properties which suggested low compatibility of these fillers
(in pre-treated form) with polyurethane foam. Based on the results, these fillers can be
incorporated into polyurethane foam to fabricate more sustainable and eco-friendly thermoset
polyurethane foams for industrials applications; such as the production of mattresses, car seats,
insulating material in construction or textile products, as well as in the cosmetic industry.

28
3.1 Introduction
Application of polyurethane (PU) foam has dominated industrial markets and in turn has
increased global regulations to promote the use of eco-friendly materials for PU production
[36]. Although microplastics and synthetic fibers, such as multi-walled carbon nanotubes
(MWCNT) and glass fiber have helped improve mechanical and thermal properties of PU foam
[37], [38], [39] these fibers pose harmful impacts to the environment and living organisms upon
extended exposure and use [40]. The focus of this study is to use natural fibers including lignin,
chitin, chitosan, hazelnut shells, and polysaccharide to enhance PU foam properties. These
natural fibers provide a cost benefit compared to synthetic fibers [41], while also promoting the
foam’s sustainable properties. Moreover, they consist of hydroxyl groups which may form
bonds with -NCO group of isocyanates upon interaction during the foaming process [42]. In
addition to the integrity fillers provide to foam cells, interfacial interactions can help strengthen
bonds between the fiber and PU, which gives these natural fibers added compatibility with PU
chains [42]. Alternatively, fillers such as hazelnut have hydrophobic properties which can
improve foam hydrolytic stability. Aside from the desired properties natural fillers possess, they
are also easily accessible in the natural environment. Consequently, natural fillers can help
provide cost benefits to industries that incorporate them into their formulation.
Lignin is the second most abundant naturally occurring complex organic material on earth and it
is an integral component of the plant cell wall. It is extracted by pulp and paper industries,
cellulosic biorefineries etc. in order to make high-value products such as syngas, carbon fiber,
phenolic compounds, and hydrocarbon products [43], [44] , [45], [46]. Polysaccharide is also an
organic filler obtained from food waste. Due to its green properties such as biocompatibility,
biodegradability, bio adhesivity, and nontoxicity, these fillers are used in biomedical,
pharmaceutical, and numerous other applications [47]. Similarly, chitin is the most abundant
biopolymer, present abundantly as α-Chitin [46] in fungal and yeast cell walls, krill, lobster, and,
shrimp, insect cuticle, and crab tendons/ shells [48]. In industrial processing, chitin is extracted
through acid and alkaline treatments in order to dissolve calcium carbonate and proteins
through chitin-based waste resources. Chitosan is obtained once the high-quality chitin is
treated with partial deacetylation under alkaline conditions. Both chitin and chitosan are widely

29
used for biomedical and pharmaceutical applications as they are biocompatible and
biodegradable in the body [49]. Additionally, these fillers are non-toxic, non-allege, antifungal,
and antibacterial amongst other green properties which make them excellent fillers for PU
foam [49], [50], [51], [52].
Hazelnut shells are obtained as byproduct from agricultural sectors. The world production of
hazelnut had achieved 488110 metric tons (kernel basis) in 2015 and the production has since
increased by almost 50% in the past decade [53]. Hazelnut shell is the major byproduct of
hazelnut industry production and represents more than 50% of the total nut weight [54]. The
shell itself is composed of different naturally occurring molecular chains. It is composed of 30%
hemicelluloses, 27% celluloses, and 43% lignin [55] which are biocompatible molecular chains,
that can be treated for use in applications such as water purification (due to their superior
adsorption properties) [56]. The hazelnut shell can be treated for applications such as in the
formulation of useful chemicals including methanol [57], reducing sugar, and furfural [58], [59].
The following figure shows the general structure of each filler used for this study:

(a)

(b)

30
(c)

(d)

(e)
Figure 3.1: Chemical structure of (a) chitin [11], (b) chitosan [11], (c) hazelnut [22], (d) lignin [23], (e) polysaccharide [24]

Past studies have researched on the use of natural fillers in PU foams. [60] researched on the
use of rapeseed oil, ultra-fine cellulose, and ground walnut shells in rigid PU. The results
showed that adding rapeseed oil helped to reduce reactivity of PU foams by reducing the
maximum temperature of the reaction mixture during the foaming process, while adding
walnut shell improved the mechanical properties. The addition of cellulose resulted in higher

31
foam resilience, and in turn improved the comfort factor of foam. Moreover, adding natural
products in the form of bio polyol to foam helped increase the foam average cell size while
maintaining the foam’s chemical structure, which resulted in lower foam density. However, the
introduction of micro cellulose into flexible PU foam system increased the apparent density of
the obtained materials. Additionally, the hardness of cellulose based foams increased while the
resiliency % decreased.
Another study, [61] focused on the preparation of natural polyol from castor oil by alcoholises
and triethanolamine. The results were analyzed using infrared spectroscopy and the analyses of
functional groups showed that the hydroxyl content increased significantly after the alcoholises
reaction. Wood flour was incorporated with this natural filler and the PU foam properties were
measured according to physical, thermal, and mechanical properties. The results showed that
using natural polyol increased curing time, which helped attain uniform PU shape. Additionally,
the mechanical properties decreased slightly while the thermal stability of foam increased.
Similarly, [62] worked on the use of silkworm cocoons as filler for PU foam and their results
showed that adding these fillers increased the density of unreinforced foam by 30%-167% of
neat PU and improved both the absolute and specific compressive properties.
Lastly, [63] researched on the development of a series of rigid PU foam composites with
eggshell as natural filler. The PU was manufactured by adding Rokopol RF551 polyol, Rokopol
G500 polyol, catalyst, surfactant, and water, to Ongronat TR 4040. Composition of eggshell in
the foam was varied according to the weight percentage of 5%, 10%, 15%, 20%, and 25%. The
results showed significant impact of eggshell on rigid PU foam properties. Eggshell helped
improve foam thermal stability, apparent density & compressive strength, water absorption,
friability, dimensional stability in the aqueous media, and resistance towards bacteria adhesion.
These studies show the potential natural fillers have for the future of eco-friendly foam
fabrication in industries. Nevertheless, current studies on natural fillers lack thorough testing of
mechanical properties, which is essential in order for industries to analyze foam-based
products. This paper focuses on an intrinsic comparison of different fillers based on a wide
range of mechanical properties, morphology, and thermal behaviour. By maintaining a
controlled foam formulation and comparing the effect each filler has on pure PU foam, detailed

32
insight is extracted. A comprehensive analysis of the tensile and tear strength, hardness,
resiliency, as well as hydrolytic stability amongst other properties, along with a morphological
and thermal analysis on the effect each filler has on PU foam will help open doors to a more
sustainable future for the environment and living organisms. Moreover, there is very little
research work on the application of lignin, polysaccharide, and hazelnut on PU foams, whereas
chitin and chitosan results have not been compared to these fillers in a controlled environment,
in detail. This paper targets this analyses by providing a thorough comparison of these fillers.
Consequently, the aim of this study is to open doors for industries on a large scale, for the
application of natural fiber-based polymer synthesis. The results go through a broad range of
mechanical properties to discuss the benefits and drawbacks of one filler versus another based
on properties such as the comfort factor of foam, foam elasticity, and hydrolytic stability
amongst other key parameters, which form the basis for sustainable industrial foam
applications.

33
3.2 Experimental
3.2.1 Materials selection
Raw materials used in this research include natural fillers (namely lignin, polysaccharide,
hazelnut, chitin, and chitosan), polyol, additives, and isocyanate. Lignin was obtained through
Dotmar, hazelnut from Nut Nursery, polysaccharide from Dupont, and chitin and chitosan from
Tidal Vision.
The polymer used for this research was free-rise flexible PU foam. It was fabricated using polyol
supplied by a collaborative company for this research. Mondur CD isocyanate was used as
additive B (additive A being polyol) for the PU addition reaction. Other reagents, including the
surfactant, catalyst, and blowing agent were also added to the foam, obtained from Sigma
Aldrich. Due to company confidentiality agreements, exact reagent compositions are not
mentioned.

3.2.2 Preparation and analyses of natural fillers used for polyurethane composites
Each natural filler was mechanically grinded using a willy mill lab grinder. The powder obtained
through grinding was sieved using a 200µm sieve in order to achieve fine powder. This ensured
that a high amount of filler surface area came into contact with the matrix, which helped
maintain effective filler matrix interaction. The filler was prepared in weight compositions of
1%, 2.5%, and 5% in order to study the effect of each filler’s composition on PU as well as to
study the threshold amount before the filler became oversaturated in the foam and resulted in
poor foam properties. These different weight amounts were identified based on initial
experimentation. The effects of adding fillers were identified by adding excess filler of higher
than 10% into foam and reducing the amount of filler added incrementally. The threshold value
which gave effective foaming was found to be 5% filler. Neat PU was also prepared alongside
the filler-based PU foams; to compare the effects of filler on PU foam.

34
3.2.3 Polyurethane composite synthesis
The PU foam was fabricated through a process of addition reactions. First, the reagents,
consisting of catalyst, surfactant, compatibilizer, and blowing agent were prepared in their
respective amounts and added into the polyol. The filler amount to be added (either 1%, 2.5%,
or 5% wt.) was also measured and added into the polyol. The polyol was then mixed using a
high shear mixer, which helped ensure that the filler dispersed evenly throughout the matrix
without agglomeration and to ensure good reagent mixing. Once the filler dispersed effectively,
the viscous nature of polyol helped prevent re-agglomeration. The polyol was then heated to
reduce viscous effects of polyol which helped disperse it uniformly when poured into the mold
during the reaction. After heating, the polyol beaker was taken out of the oven and isocyanate
was added while mixing at high shear rates. After approximately 10 seconds of mixing, the
material was poured into a mold for the foaming reaction to take place. Once poured, foam
was left to expand and cure for 24 hours before testing. The following schematic below shows
this process:

Figure 3.2: Steps for Polyurethane foam Composites

3.2.4 Characterization of polyurethane foam composites


3.2.4.1 Procedure for mechanical analysis of polyurethane foam composites
Mechanical analyses of foams were based on six different types of testing. These included
tensile testing to measure the maximum stress the foam was able to withstand before
breaking, and elongation % which measured the maximum elastic capacity of foam without

35
breaking. Tensile testing was done according to ASTM D3574-17 testing method. Type IV sized
dog bones were used for the test and the test was performed using a Microtester 5848
(Instron) machine. Five dog bone samples were tested at a rate of 500 mm/min and tensile
stress along with ultimate elongation % results were extracted. Additionally, foam density was
also measured using ASTM D3574-17 using three foam samples to obtain an average density.
The sample volume was measured using a caliper followed by the mass of each sample. Density
was obtained using mass per cubic cm. Hardness of the sample was measured using an Asker
apparatus.
Another important test that was carried out was tear testing for the material, which measured
the tear propagation resistance of foam. This was also performed using ASTM D3574-17 testing
method for foam samples with dimensions of approximately 25mmx40mmx152.4mm. Five
samples were taken and the Microtester 5848 (Instron) machine was used for the test.
Rebound resilience testing was used to measure foam resiliency and carried out based on ASTM
D3574-17. A ball was dropped from a height of ~500mm and the height it bounced to after
hitting the foam block was measured. Three different readings were taken to ensure accurate
resiliency readings.
Compression set testing was carried out using SATRA TM64. Three foam samples in the shape
of circular disks of diameter 29-30mm were cut using a drill press and circular die cutter. The
samples were compressed at a pressure of approximately 0.69 MPa at room temperature for 24
hours. Following the test, the samples underwent 1 hour of recovery without any stress
induced on them, following which the sample thickness difference was measured.
Lastly, foam samples were tested for their hydrolytic stability to measure how the mechanical
properties retain tensile strength after coming into contact with 100% humidity for seven days.
SATRA TM328 standard was followed for hydrolysis testing. This test involved placing dog bone
samples in the oven at 70oC for 6 days in a sealed chamber containing distilled water. The
accelerated process was used to measure tensile properties of the PU foam and how they were
affected over time.

36
3.2.4.2 Procedure for morphological analysis of polyurethane
Cell size of each PU foam composites was measured using scanning electron microscopy (SEM).
This was done using the JSM-IT100 machine by analyzing the nitrogen fractured surface of
foam. The surface was sputter coated with gold prior to SEM imaging in order to obtain a
conductive surface for analysis. Fragments of each foam sample were selected randomly and
approximately 150-200 cells were analyzed within the structure using the ImageJ software.

3.2.4.3 Procedure for thermal analysis of polyurethane foam composites


Thermal analysis was performed with the help of differential scanning calorimetry (DSC)
by using the DSC Q2000 (TA instruments) under an oxygen environment. A heat-cool cycle was
used to analyze the crystallization temperature and melting temperature of the composites.
The starting temperature was chosen to be at room temperature, and the temperature was
increased up to 200°C at the rate of 10°C/min. Once the heating cycle was complete, the
sample was cooled to below -50oC at -10°C/min.

3.3 Results and Discussion


3.3.1 Morphology of polyurethane foam composites
3.3.1.1 Morphological characterization of polyurethane composites using Scanning
Electron Microscopy (SEM)
A morphological study of the composites was done to help understand the quality of each PU
foam composite. Additionally, it gave insight into some of the important parameters of PU foam
including tensile strength and elongation %. The surface of each foam was imaged to identify
filler effects and to see how the cell shape and size varied with filler wt. %. Furthermore, the
SEM images helped identify cell rupture and poor foaming.
Results presented in table 1 show that the cell size decreased as filler content increased. As
discussed in the mechanical analyses section, this was due to moisture content present within
each filler [64] which acted as a blowing agent and promoted cell growth. Nevertheless, the
images also show that as filler size increased, it impacted the cell structure, causing rupture and

37
poor foaming by acting as a hindrance or impurity within the foaming reaction. This showed
that adding high filler content reduced tensile strength and ultimate elongation.

38
Pure PU 1% Chitin 2.5% Chitin 5% Chitin

Foam void/ crack


Foam void/ crack

1% Chitosan 2.5% Chitosan 5% Chitosan 1% Hazelnut

2.5% Hazelnut 5% Hazelnut 1% Lignin 2.5% Lignin

Foam void/ crack Foam void/ crack

Foam void/ crack

5% Lignin 1% Polysaccharide 2.5% Polysaccharide 5% Polysaccharide


Foam rupture

Foam rupture
Foam void/ crack
Foam void/ crack

Figure 3.3: Scanning Electron Microscope (SEM) images for polyurethane composites with 1%, 2.5%, and 5% fiber loading

39
Table 3.1: Average cell size ( m) of polyurethane foam composites

Thermoset Average cell size (µm) Thermoset Average cell size (µm)
Pure PU 263.20 PU with 2.5% hazelnut 295.87
PU with 1% chitin 249.80 PU with 5% hazelnut 175.75
PU with 2.5% chitin 194.12 PU with 1% lignin 204.24
PU with 5% chitin 203.21 PU with 2.5% lignin 221.32
PU with 1% chitosan 257.91 PU with 5% lignin 218.22
PU with 2.5% chitosan 251.22 PU with 1% polysaccharide 282.78
PU with 5% chitosan 171.78 PU with 2.5% polysaccharide 321.99
PU with 1% hazelnut 243.55 PU with 5% polysaccharide 216.66

3.3.2 Mechanical properties of polyurethane foam composites


3.3.2.1 Significance of mechanical testing from an industrial standpoint
The properties of PU foam composites were extracted at their breaking point based on tensile
strength and ultimate elongation. Additionally, the different polymer blends were further
compared according to their tear strength, density, rebound resilience, compression set, and
hardness. The results (plotted against PU wt.% composition) are shown in Fig. 4 and Fig. 5.
Tensile strength of PU foam is an essential parameter for industrial applications. Tensile
strength determines the maximum axial force an object can withstand before failure/ break.
Having low PU tensile strength leads to compromise of product quality; thus, it was important
to ensure that adding any filler to PU foam did not significantly reduce this parameter. Similarly,
Fig. 4b shows ultimate elongation of each filler compared to pure PU. Ultimate elongation
measures the maximum elasticity of PU foam before rupture. This property is related to foam
elasticity and is important for applications such as use in the textile industry for mattress
production or for shoe soles where foam is constantly stretched and stressed with high forces.
Low ultimate elongation indicates low stretch or bending capacity of foam which can cause
early product failure and low quality.
Another key aspect for industries is to study the impact moisture can have on foam-based
product. Hydrolytic testing helps to analyse foam stability and properties, to ensure that foam
can maintain its properties after high exposure to humid conditions. If PU foam has poor

40
hydrolytic stability, it raises concerns and limitations for use in applications such as in
construction, automotive industries, clothing etc. where foam is exposed to high levels of
moisture. Low hydrophobicity causes a gradual decline of foam tensile properties, translating to
low product quality and lifespan. One of the important aspects of this study was to see whether
PU foam hydrophobicity improves with the use of each filler. Results presented in Fig. 4c,4d
show that all PU foam composites have a reduction in tensile properties. This is expected as the
foam had low density and was exposed to high thermal stress (70oC) and vapor pressure for 7
days during aging. The constant stress disrupted polymer chains and caused them to weaken,
resulting in lower mechanical strength. When comparing aging results to those of pure PU
however, the composites sustained their tensile strength while the ultimate elongation % was
seen to increase; due to reduction in brittleness induced by aging.
The tear strength of the material is shown in Fig. 5a. Tear strength measures foam break
strength and tear propagation as force is applied across it. At an industrial scale, having low
tear strength can lead to increased chances of tear propagation in foam through the crack tip.
For instance, in the automotive sector, it is important to maintain high tear resistance to
improve the quality and comfort of car seats. Similarly, for packaging industries, it is essential
that foam can withstand tearing forces in order to provide good insulation protection. Other
applications such as use in mattresses also requires good tear properties to maintain a high life
span and comfort level. Fig. 5b shows the results for rebound resilience measurements of PU
foam. These results measure elasticity of foam composites. Lastly, hardness of each PU
composite was measured as shown in Fig. 4d. Hardness is an important aspect when
considering applications where product comfort and stiffness are concerned [65].

3.3.2.2 Discussion of mechanical testing results


Analyzing tensile strength results (presented in Fig. 4a), an increase in tensile strength of
approximately 35% and 34% was measured, for 1% chitin and 1% hazelnut respectively. This
shows that chitin and hazelnut helped promote strong interactions between the matrix and
filler in the axial direction, compared to other fillers. As the amount of filler was increased,
there was a reduction in tensile strength which showed that PU foam had a filler limiting value
for each composite and adding too much filler to foam caused the properties to decrease. The

41
exception to this trend was lignin, for which the mechanical properties increased at 5 wt.%. The
main reason for this was that lignin filler helped bind foam cells together but interfered in
effective urethane linkage which reduced the foaming rate. Consequently, the cells were closely
packed upon foaming resulting in an increase in foam plastic region per cubic centimeter. This
caused higher tensile strength for 5% lignin. Nevertheless, this tensile strength had
comparatively high standard deviations compared to other PU foams due to high stress
concentrations present within 5% lignin-based PU foam, which resulted in non-uniform stress
distributions causing high deviations in mechanical properties. An important aspect that led to
fluctuations in results between each filler can be attributed to polar groups present within
them and their hydrophobic properties; due to difference in the structure of each filler. The
results indicated that fillers such as chitin and hazelnut had functionalized polar groups present
at the surface which formed interactions with the polyol matrix. However, fillers including
chitosan, lignin, and polysaccharide did not have such polar groups that were able to readily
interact with the matrix. Instead, these fillers had an affinity to agglomerate together during
foaming instead of interacting with the matrix [66]. This was particularly evident for 5% lignin,
which had visibly agglomerated fibers across the foam volume. Nevertheless, polar groups of
such fillers can be treated with chemical solvents in order to improve matrix filler interaction
and should be researched for future studies.
Analyzing Fig. 4b, ultimate elongation % was seen to increase when chitin was added. The
percent increase for 1% chitin, 1% polysaccharide, and 1% hazelnut was 39%, 21.7%, and 9.4%
respectively. This increase in elongation is attributed to the increase in viscosity of the polymer
matrix upon addition of 1% filler. Increase in viscosity helped promote foaming and reduce
foam brittleness [67]. Moreover, reduction in density (shown in Fig. 5c) for filler-based PU
further explains the less brittle nature of these PU foam composites. Density reduction is due to
the moisture present within fillers at the time of foaming, which acts as blowing agent causing
the growth of a higher number of cells [68]. This moisture resulted in an increase in the open
cell content of foam resulting in lower foam density. Increase in strain is something which has
been reported in previous research studies as well. [69] measured the properties of PU foam
upon addition of organoclay and modifying the hard/ soft segments of PU foam. The results

42
indicated an increase in the elongation % upon small additions of the filler, but the strain at
break reduced as the amount of filler in the PU foam was increased. Similarly, [70] reported an
increase in elongation of approximately 30% upon addition of clay-based fillers into PU. The
increase in value is also through effective matrix and filler interactions during the reaction. The
surface polar groups present within chitin and polysaccharide formed effective interactions
with the N=C=O group and polyol. As the amount of filler increased however, the elongation%
decreased. This was expected because higher levels of filler promote filler agglomeration,
leading to high stress concentrations. These stress concentrations caused foam to cure
unevenly, forming voids, cell rupture, and foam cracks (shown in Fig. 3). Uneven stress
distribution is also the cause of large deviations in 1% hazelnut-based PU.
Comparing pure PU tear strength to its composites (Fig. 4a), the strength of the material was
subjective to the filler used. For chitin, the strength of PU remained constant at 1% and 2.5%
chitin loading while it increased by ~20-30% for 5% chitin. For hazelnut, 1%, 2.5%, and 5% filler
helped improve foam tear strength although 5% hazelnut showed high deviation, due to poor
filler dispersion and non-uniform stress distributions throughout the filler. Chitosan also
showed good tear strength properties, however there was a decrease in tear strength as higher
composition of filler was added. Lastly, lignin and polysaccharide showed low tear properties
upon addition of filler. The high tear strength shown by hazelnut was something reported by
other studies as well [71], [72], [73]. The main reason for this increase was due to good matrix
filler interactions and fracture toughness of the filler which helped bind foam cells together to
withstand tear forces.
Rebound resilience % results of Fig. 4b showed that resiliency of PU composites was maintained
close to that of pure PU at approximately 38% energy transfer. The remaining energy was
transferred through foam cells in the form of vibration and heat energy. 1% chitin and 1%
hazelnut showed improvements by approximately 8% and 22% respectively. These
improvements can be justified by the fact that fillers help dissipate energy more effectively
throughout PU foam upon impact, thus enabling more efficient transfer of energy back to the
ball [68], [74]. On the other hand, any decrease in resiliency may be attributed to foam cell
rupture upon addition of excess filler due to stress buildup within the foam. Excess filler acted

43
as impurity resulting in poor matrix-filler interactions and cell coalescence/ cell rupture [40].
[75] investigated the comfort factor of foams with different resiliency values. They reported
that increasing foam resiliency helped improve its comfort factor; an important factor for
automotive and textile industries using PU. Consequently, it can be deduced that 1% chitin and
1% hazelnut helped improve foam comfort.
For PU composite density (Fig. 4c), it was found that adding filler into foam typically reduced
foam density. The reduction % was approximately 26.5% of pure PU density. This decrease was
due to the moisture content present within the natural filler. Moisture present within the filler
acted as blowing agent and promoted cell growth [68]. Moreover, it is hypothesized that filler
addition resulted in an increase in the number of open cells as well as chemical/ physical
displacements in foam linkage, which resulted in reduction of density. This reduction is a
desirable property for industrial applications, particularly automotive industry, where using low
density, high strength material helps improve fuel efficiency. Lastly, foam hardness (Fig. 5d)
shows that for all PU composites besides 5% lignin, hardness was lower compared to pure PU
foam. This reduction is attributed to the low density of PU composites. Nevertheless, 5% lignin
was an outlier to this trend which can be linked to high curing time and low foaming rate, which
resulted in a densely packed foam structure as shown in the SEM images in Fig. 5. This densely
packed structure resulted in shorter bond lengths and increased foam firmness.

44
1 800
1% Filler 1% Filler
2.5% Filler 2.5% Filler

Ultimate Elongation (%)


Tensile Strength (MPa)

0.8 5% Filler 5% Filler


600
0.6
400
0.4
200
0.2

0 0
PU

e
n

t
n

t
PU

e
n
nu

nu
rid
iti

sa

ni

rid
ni
iti

sa
Lig
Ch

Lig
re

zle
ito

Ch

zle
re

ito
ha

ha
Pu

Pu
Ch

Ha

Ch

Ha
ith

ac
ith

ith

ith

ac
lys

lys
w

w
ith

ith

ith

ith
Po
PU

Po
PU

PU

PU
w

w
PU

ith

PU

PU

ith

PU
w

w
PU

PU
(a) (b)
0.8 800
Ultimate Elongation % after aging
Tensile Strength (MPa) after aging

1% filler 1% filler
2.50% filler 2.50% filler
0.6 600 5% filler
5% filler

0.4 400

0.2 200

0 0

t
PU

e
n
PU

e
n

t
n

nu
nu

rid
ni
rid

iti

sa
iti

sa

ni

Lig
Ch
Lig

zl e
Ch

re

ito
re

zl e
ito

ha
ha

Pu
Pu

Ch

Ha
Ch

Ha

ith

ith

ac
ith

ac
ith

lys
lys

w
w

ith
ith

ith
ith

Po
Po

PU

PU
PU

PU

w
w

w
w

PU

ith
PU

PU
ith

PU

w
w

PU
PU

(c) (d)

Figure 3.4: Mechanical results for polyurethane foam composites displaying (a) Tensile strength of each composite for 1%, 2.5%,
and 5% fiber loading, (b) Ultimate elongation % of each composite for 1%, 2.5%, and 5% fiber loading, (c) Tensile strength of PU
composites after aging for 1%, 2.5%, and 5% fiber loading (d) Ultimate elongation % of each composite after aging for 1%, 2.5%,
and 5% fiber loading

45
3 60
1% Filler 1% Filler
2.5% Filler 2.50% Filler

Rebound Resilience %
5% Filler
Tear Strength (N/mm)

5% Filler
2 40

1 20

0 0
PU

e
n

t
n

PU

e
nu

t
n
rid
iti

sa

ni

nu
rid
iti

sa

ni
Lig
Ch
re

zle
ito

ha

Lig
Ch
re

zle
ito

ha
Pu

Ch

Ha

Pu
ith

ac
ith

Ch

Ha
ith

ac
ith
lys
w

w
ith

lys
ith

w
ith

ith
Po
PU

PU
w

Po
w

PU

PU
w

w
PU

ith

PU

PU

ith

PU
w

w
PU

PU
a b
0.2 40
1% Filler
1% Filler
0.16 2.5% Filler
30 2.5% Filler
Hardness (Asker)
5% Filler
Density (g/cm3)

5% Filler
0.12
20
0.08
10
0.04

0 0
n

e
n

t
PU

nu
t
PU

rid
e
n

iti

sa

ni
nu
rid
ni
iti

sa

Lig
Ch

zl e
ito
re

ha
Lig
Ch

zl e
re

ito

ha

Pu

Ch

Ha
ith

ac
ith
Pu

Ch

Ha
ith

ith

ac

lys
w

w
ith

ith
lys
w

w
ith

ith

Po
PU

PU
w

w
Po
PU

PU
w

PU

ith

PU
PU

ith

PU

w
w

PU
PU

c d

Figure 3.5: Mechanical results for polyurethane foam composites displaying (a) Tear strength of each composite for 1% fiber,
2.5%, and 5% fiber loadings, (b) Ball drop resiliency % of each composite for 1%, 2.5%, and 5% fiber loading, (c) Density of PU
composites after aging for 1%, 2.5%, and 5% fiber loading (d) Hardness of each composite after aging for 1%, 2.5%, and 5% fiber
loading

46
3.3.3 Thermal analysis using Differential Scanning Calorimetry (DSC)
3.3.3.1 Significance of thermal testing from an industrial standpoint
Due to the inner material structure of thermosets, most bonds within PU foam were crosslinked
with each other. Consequently, DSC results showed that there was no cold crystallization or
melting temperature for PU and its composites. However, the graphs in Fig. 6 indicate that
there was a glass transition temperature for the PU foams. Glass transition temperature (Tg) is
an important physical property for industrial sectors using thermoset polymers, when
considering end-use applications. Tg is the temperature at which a material transitions from a
brittle, glassy state to the rubbery state. For thermoset polymers below glass transition
temperature, chemical crosslinking prevents particles from flowing and there is not enough
versatility for particles to unwind, resulting in inflexible and glassy polymer. Knowing the Tg
value is important for industries as it helps to understand PU limitations such as durability in
varying climates, or to determine glass transition temperature for specific applications such as
insulation, packaging etc.

3.3.3.2 Discussion of thermal analysis for polyurethane foam composites


Table 2 shows glass transition temperature results obtained through DSC runs. The DSC trace
gives a value of approximately -27.68°C glass transition temperature for pure PU. It was
observed that the glass transition temperature for the foam composites were between -27°C to
-32°C. The average glass transition temperature of all thermoset polymers was -29.66°C, which
was anticipated when comparing to the control formulation. Accordingly, all testing samples
possessed similar glass transition temperatures due to high composition of the matrix present
within the fillers. This was not unexpected as fillers have been reported as having little effect to
lower the glass transition temperature (Tg) of polymers [76].

47
Table 3.2: Glass transition temperature of PU foam composites

Thermosets Glass transition temperature (Tg) °C Weight(g)


Pure PU -27.28 0.0065
1% chitin -30.28 0.0065
2.5% chitin -28.44 0.0065
5% chitin -31.64 0.0065
1% chitosan -29.56 0.0065
2.5% chitosan -29.99 0.0065
5% chitosan -29.81 0.0065
1% lignin -30.10 0.0065
2.5% lignin -29.72 0.0065
5% lignin -26.29 0.0065
1% hazelnut -29.20 0.0065
2.5% hazelnut -29.16 0.0065
5% hazelnut -28.32 0.0065
1% polysaccharide -29.30 0.0065
2.5% polysaccharide -31.99 0.0065
5% polysaccharide -31.24 0.0065

48
0.4 0.3
Pure PU Pure PU
1%Chitin 0.3 1%Chitosan 0.2
2.5%Chitin 0.2 2.5%Chitosan 0.1

Heat Flow (W/g)


Heat Flow (W/g)
5%Chitin 0.1 5%Chitosan
0
0 -60 -40 -20 -0.1 0 20
-60 -40 -20 -0.1 0 20
-0.2
-0.2
-0.3
-0.3
-0.4 -0.4
-0.5 Glass transition temperature -0.5
Glass transition temperature -0.6 -0.6
Temperature (oC) Temperature (oC)
a b

0.3 0.3
Pure PU Pure PU
1% Lignin 0.2 1% Polysaccharide 0.2
2.5% Lignin 0.1 2.5% Polysaccharide0.1
Heat Flow (W/g)

5% Lignin

Heat Flow (W/g)


5% Polysaccharide
0 0
-60 -40 -20 -0.1 0 20 -60 -40 -20 -0.1 0 20
-0.2 -0.2
-0.3 -0.3
-0.4 -0.4
Glass transition temperature -0.5 -0.5
Glass transition temperature
-0.6 -0.6
Temperature (oC) Temperature (oC)
c d

Pure PU
0.3
1% Hazelnut 0.2
2.5% Hazelnut0.1
Heat Flow (W/g)

5% Hazelnut
0
-60 -40 -20 -0.1 0 20
-0.2
-0.3
-0.4
Glass transition temperature -0.5
-0.6
Temperature (oC)
e
Figure 3.6: DSC heat-cool cycle for (a) Chitin (b) Chitosan (c) Lignin (d) Polysaccharide (e) Hazelnut at 1%, 2.5%, and 5% fiber
loadings

49
3.4 Conclusions
In this study, mechanical, morphological, and thermal properties of PU foam composites with
1%, 2.5%, and 5% of chitin, chitosan, lignin, polysaccharide, and hazelnut were investigated.
Results show that natural fillers have potential to replace synthetic ones for use in the industrial
sector. 1% chitin, polysaccharide, and hazelnut showed significant improvements in tensile
strength and elongation. These results were attributed to effective matrix-filler interactions and
binding of foam cells. Similarly, 1% hazelnut showed increase in resiliency while foam hydrolytic
stability and tear strength also showed improvements. Lignin and chitosan however showed
lower mechanical properties in terms of tensile and elongation % values, due to low matrix-
filler interactions and high agglomeration. Due to the ease of filler processing and fabrication of
PU composites, this process can easily be adopted at an industrial scale to produce more eco-
friendly foam products; leading to a more sustainable future. Lastly, this research study helps
provide valuable insight on filler behaviour, which industries can use to tailor bio-based foams
towards high performance applications.

50
Natural fibers as reinforcement for closed-molded polyurethane foam
plaques: evaluating the mechanical, morphological, and thermal
properties
Abstract
Natural fillers including cellulose, chitin, hazelnut, and eggshell are some of the most abundant
bio-resources available for performance-cost-benefit applications. For this study, the
aforementioned fibers were added to pure polyurethane foam and the properties were
investigated based on mechanical (tensile, split tear, elongation, resilience, compression set,
and hardness), thermal (thermogravimetric and differential scanning calorimetry), and
morphological analyses. These natural fibers were added in wt. % of 1%, 2.5%, and 5% using
high shear mixing rates. The closed molded polyurethane foam was prepared using an addition
reaction of polyol and isocyanate. The resulting foam composites embedded with natural fibers
had uniform color distribution, indicating good fiber dispersion. 1% fiber loading showed
improvements in tensile and resiliency, whereas split tear properties decreased. Moreover,
morphological results indicated uniform foaming and dispersion of fillers. This study reinforces
the possibility of a more sustainable future by replacing synthetic fillers with natural fillers in
industrial applications; including in the automotive, textile, construction, and other leading
consumer based industries.

51
4.1 Introduction
In present times, novel research on polyurethane (PU) foam products has vastly increased their
application in industrial sectors. As a result of their valuable characteristics (including light
weight and comfort), development of PU foam has increased significantly over the years; with
the help of additives and fillers. Adding fillers into polymers helps provide added strength and
rigidity to the pure polymer [77]. The improvement in mechanical and thermal properties is
mainly a result of interfacial interactions between the fibers and matrix. Many different types
of fillers have been incorporated into PU foam including carbon fiber, PET, carbon nanotube,
tire rubber etc. [78], [79], [80], [81]. Although these fillers are highly effective in enhancing
polymer properties, they pose harmful effects to living organisms. Carbon fibers and nanotubes
are known to be highly persistent in the environment causing health problems for living
organisms by producing toxic response upon reaching the lungs in significant quantity [82].
Moreover, CNTs are known to have low biodegradability and water solubility [83]. Additionally,
they highly energy intensive and costly life cycle [84], [85]. Furthermore, it has been reported
through cradle to gate level comparisons for life cycle assessment that PET causes major
damage in terms of human health and ecosystem quality [86]. As a result, although these fillers
enhance the properties of polyurethane foam, environmental and sustainability concerns have
led to an increased need to use more eco-friendly fillers [87], [88], [89].
Significant research interest has been focused on the use of natural fibers in polyurethane
foam. For instance, [90] added sisal fibers to polyurethane as reinforcement material. Sisal fiber
is one of the widest used natural fibers as it is easily cultivated. It is extracted from the leaves of
the Agave Sisalana plant [91]. These fibers are present as a bundle of sub-fibers reinforced with
spirally oriented cellulose, in a hemicellulose and lignin matrix. Their appealing characteristics
such as low cost and density, in addition to high specific strength and modulus, renewability,
and sustainability make them enticing fillers for testing. Similar to sisal fibers, chitin, cellulose,
eggshell, and hazelnut shells are also abundantly available through natural means. Chitin can be
extracted from exoskeleton or arthropods while cellulose is sourced from wood, straw,
bagasse, bacteria, and plant cell wall [92], [93]. Cellulose is the most abundant natural polymer
and chitin is the second most abundant natural polymer. Both possess similar polysaccharide

52
structures with slight differences in hydroxyl and acetyl amine functional groups [94], [95], [96].
Hazelnut shell is also widely available as waste resource, produced annually at more than
600,000 tons [97], [98] [99] and the global consumption of hazelnut is expected to increase as
hazelnuts are considered part of a healthy diet [97]. Eggshell weighs approximately 10% of the
total hen egg mass and it is a very significant solid waste produced from food processing and
manufacturing plants [100], [101]. In China, about 4,000,000 tons of eggshell waste is
generated, and this amount is estimated to grow [102]. Traditionally, majority of this eggshell
waste has been disposed into landfills without any pretreatment, but in accordance with the
environmental odor from biodegradations [103] this is not a desirable practice. Accordingly, a
great deal of effort is being conducted for the application of eggshell as value-added products
[101], [104], [105]. The use of these natural fibers is vital as it helps improve the overall
sustainability of processes. Synthetic fibers such as glass, carbon, aramid, Kevlar, etc. are known
to possess serious drawbacks such as high cost, high density, and poor recycling and degradable
properties [77], [106], [107], [108].
Polyurethane foam is used in several industrial applications due to enticing properties such as
low density and high strength. Viscoelasticity of this foam is also an essential advantage as it
helps to maintain rigidity while also enabling the dissipation of energy through frictional losses;
by behaving as a viscous fluid [109]. In present times, the use of polyurethanes is one of the
most important class of polymers that helps provide ongoing improvements in the quality of
human life [110]. In 2017, the worldwide consumption of PU was estimated at 60.5 billion USD
and predicted to be over 79 billion USD by 2021 [111]. Moreover, statistics show that almost
67% of global PU consumption is in the form of foams [112]. Another advantage that PU foams
bring is the wide range of properties that can be achieved by using small modifications of the
formulation, such as increasing the polyol functionality without changing the molecular weight
leads to an increase in foam hardness and reduction in elongation %, tensile, and tear strength
[113].
Currently, comparative research on the use of eggshells, hazelnut shells, cellulose, and chitin to
enhance polyurethane foam properties in a closed mold is very limited. To the best of our
findings, no formal investigation has been performed to test the effect these natural fillers have

53
on a controlled PU formulation. For this study, each filler was added into foam at weight % of
1%, 2.5%, and 5% through high shear mixing. The resulting matrix filler slurry was uniformly
dispersed (evident through uniform color distribution), and viscous nature of polyol prevented
the filler fibers from agglomerating together. The composites that were fabricated through this
PU addition reaction were tested according to their mechanical, thermal, and morphological
properties. Additionally, they were characterized to measure enhancements in mechanical
strength compared to the controlled foam formulation. Enhancements in mechanical strength
with the help of such fillers provides promising replacements for the use of synthetic fibers;
which will inevitably improve sustainable industrial processes and open more job opportunities
in the research and development of eco-friendly composite processing.

54
4.2 Experimental
4.2.1 Materials selection
Raw materials used in this research study included natural fillers (hazelnut, cellulose, chitin, and
eggshell), polyol, additives, and isocyanate. Hazelnut was obtained from Nut Nursery, while the
other fillers were sourced from Sigma Aldrich.
Closed-mold flexible PU foam was used for this study. It was fabricated using polyol supplied by
a collaborative company working on this research. Mondur CD isocyanate was used as additive
B (additive A being polyol) for the PU addition reaction. Other reagents, including surfactant,
catalyst, and blowing agent were also added to the foam, obtained from Sigma Aldrich. Due to
company confidentiality agreements, specific reagent compositions were not mentioned.

4.2.2 Preparation of natural filler for addition in polyurethane


Each natural filler was mechanically grinded using a willy mill lab grinder. The powder obtained
through grinding was sieved using a 100µm sieve in order to attain fine powder. Nevertheless,
fiber outliers ranging from 100-300µm also existed within the powder in very small amounts.
Small filler size ensured that a high amount of filler surface area came into contact with the
matrix, which helped maintain effective filler matrix interaction. The filler was prepared in
weight compositions of 1%, 2.5%, and 5% in order to study the effect of each filler’s
composition on PU and to study the saturation amount, after which the foaming reaction
resulted in poor foam properties. Neat PU was also prepared alongside the filler-based PU
foams. This was the control formulation and used to compare the effects of filler on PU foam.

(a)

55
(b)

(c)

(d)
Figure 4.1: Chemical structure of (a) hazelnut [38] (b) cellulose [39], (c) chitin [40], (d) eggshell [41]

56
In short, the results showed that natural fibers added had effective interactions with the matrix.
All fillers consisted of hydroxyl groups which were hypothesized to have reacted with the
excess isocyanate, and in turn, strong interfacial interactions were formed throughout the
foam. The figures below show these hypothetical reactions which helped enhance the
mechanical properties.

Figure 4.2: Reactions of natural fillers involved in polyurethane foaming

57
4.2.3 Steps to fabricate polyurethane composites
PU foam was fabricated through a process of addition reactions. For preparation, the reagents,
consisting of catalyst, surfactant, compatibilizer, and blowing agent were prepared in their
respective wt.% and mixed into the polyol. The filler amount to be added (either 1%, 2.5%, or
5% wt.) was also measured and added into the polyol. The polyol was then mixed using a high
shear mixer, which helped ensure that the filler dispersed evenly throughout the matrix
without agglomeration and to ensure good reagent mixing. Once the filler dispersed effectively,
the viscous nature of polyol helped prevent re-agglomeration. The polyol was then heated to
reduce viscous effects of polyol which helped disperse it uniformly when poured into the mold
during the reaction. After heating, the polyol beaker was taken out of the oven and isocyanate
was added while mixing at high shear rates. After approximately 10 seconds of mixing, the
material was poured into a mold for the foaming reaction to take place. Once poured, foam
was left to expand and cure for 24 hours before testing. The following schematic below shows
this process:

Figure 4.3: Steps to fabricate polyurethane foam composite

4.2.4 Characterization of rigid polyurethane foams


4.2.4.1 Procedure for mechanical testing of polyurethane foam composites
Mechanical testing was based on a wide range of stress loading properties. These included
tensile testing which measures the maximum stress of the foam before failure, and elongation
% which measured the maximum elastic potential of foam before breaking. Tensile testing was
done according to ASTM D3574-17 tensile testing. Type III dog bones were used for the test
using the eXpert 7600 Admet tension testing machine. Five dog bone samples were tested at a
rate of 500 mm/min and tensile stress along with ultimate elongation % results were extracted.

58
Additionally, foam density was also measured using ASTM D3574-17 using three foam samples
to obtain an average density. The sample volume was measured using a caliper followed by the
mass of each sample. Density was obtained using the measured mass and volume. Hardness of
the sample was measured using an Asker apparatus.
Furthermore, split tear testing was also carried out, which measured the tear propagation
resistance of foam. This was also performed using the SATRA TM65 testing method. Five
samples were taken and the eXpert 7600 Admet tension testing machine was used for the test.
Rebound resilience testing was used to measure foam resiliency % and carried out using the
ASTM D3574-17 standard. A ball was dropped from a height of 500mm and the recovered
height after hitting the foam block was measured. Three trials were carried out to ensure
accurate resiliency readings.
Compression set testing was carried out using SATRA TM64. Three foam samples in the shape
of circular disks of 30mm diameter were cut using a drill press and circular die cutter. The
samples were compressed at a pressure of 0.69 MPa at room temperature for 24 hours.
Following the test, the samples were left to recover for 1 hour without any stress induced on
them, after which sample thickness was measured. Permanent reduction in thickness was
measured by using the difference in thickness.
Lastly, foam samples were tested for their hydrolytic stability (aging) to measure how the
mechanical properties retain tensile strength after coming into contact with 100% humidity for
seven days. SATRA TM328 standard was followed for hydrolytic stability testing. This test
involved placing dog bone samples in the oven at 70oC for 6 days in a sealed chamber
containing distilled water. The accelerated process was used to measure tensile properties of
PU foam and how the foam was affected over time.

4.2.4.2 Morphological analysis of polyurethane foam composites


Cell size of each PU foam composites was measured using scanning electron microscopy (SEM).
The JSM-IT100 machine was used by analyzing the nitrogen fractured surface of foam. The
surface was sputter coated with gold prior to SEM imaging in order to obtain a conductive

59
surface for analysis. Fragments of each foam sample were selected randomly and
approximately 150-200 cells were analyzed within the structure using the ImageJ software.

4.2.4.3 Procedure for thermal analysis of polyurethane foam composites


Thermal analysis was performed with the help of differential scanning calorimetry (DSC) by
using the DSC Q2000 (TA instruments) under an oxygen environment. A heat-cool cycle was
used to analyze the crystallization temperature and melting temperature of the composites.
The experiment was initiated at room temperature, and was increased up to 200°C at the rate
of 10°C/min. Once the heating cycle was complete, the sample was cooled to below -50oC at -
10°C/min.
Thermogravimetric analysis (TGA) was also carried out. This was done by placing the composite
in a nitrogen chamber and heating the sample at a rate of 10OC/min to 800oC. The profile was
recorded based on the degradation temperature of hard and soft segments.

4.2.4.4 Procedure to measure foam porosity


Ultrapycnometer 1000 was used to measure foam porosity. A sample of each composite was
placed in a gaseous chamber and 10-20 gas runs were carried out. The results measured the
percentage of closed cells present within the composite by passing a gas through the
composite’s open cells and pores.

4.3 Results and Discussion


4.3.1 Mechanical testing of polyurethane foam composites
4.3.1.1 Tensile testing results analyses of polyurethane foam composites
Comparing the results, it was observed that adding 1 wt.% filler to the control formulation of
PU foam helped increase tensile strength. The average increase in tensile strength was
measured at 15-20% for hazelnut, cellulose, and eggshell whereas chitin showed an 8%
improvement. Increasing the amount of filler to 2.5 wt.% had a negative impact on cellulose
and chitin fillers, decreasing their tensile properties by 5% and 10% respectively. On the other

60
hand, hazelnut and eggshell maintained good tensile strength, with a 13% increase in strength.
However, upon increasing the wt.% to 5%, there was a noticeable decrease in the properties,
with all composite strengths ranging very close to the control formulation. Nevertheless, the
results indicated that although adding excess filler reduced the properties, they were still
comparable to the control formulation. The highest drop in strength was observed for chitin
where the strength reduced by 5% with a standard deviation of ±0.13.
An important aspect for PU foam properties is the impact of moisture. Figure 3 shows the
tensile strength of PU foam after aging. The results showed that the properties for filler-based
foams were not impacted critically due to moisture exposure at high temperature. It is evident
from the graph that 1% filler loading maintained improvement in tensile strength. Particularly,
hazelnut and eggshell showed higher improvements of 14% and 21% whereas cellulose showed
5% improvements while there was a slight decrease in the strength of chitin fibers. As the
amount of filler increased, there was a further decrease in tensile strength of cellulose and
chitin by 7% and 19% respectively. However, hazelnut and eggshell maintained higher strength,
showing an increase of 12%. As the amount was further increased to 5% fiber loading, it was
observed that cellulose maintained 10% improvements (although the strength was seen to
decrease for 2.5% cellulose, this was accounted for by the large deviation in results), hazelnut
and eggshell based PU composites maintained slight improvements, whereas cellulose showed
10% improvements and chitin showed a reduction in the strength.
As a result of the random nature of polyurethane reactions, there are several reasons which
explain the effect of fillers to improve foam properties. To maintain low deviation due to
formulation differences, 3-5 dog bone samples were tested for each formulation. As a result,
the improvements are attributed to three fundamental reasons. Firstly, the filler particles
present within foam helped restrict mobility of polymer chains during tensile loading [118].
Moreover, adding filler to foam helped increase the resistive surface area which in turn
improved the tensile loading capacity of PU foam [119]. Lastly, [120] studied the effects of
bonding on polyurethane, and the improvements indicated that at low amounts, the filler
interacted effectively with the matrix forming high matrix-filler reinforcements to resist loading
forces [121]. This explains the improvement shown in figure 2 and figure 3 below. Nevertheless,

61
it was shown that increasing the filler to 2.5% and 5% reduced tensile strength. This was caused
by the impact excess filler has on foam structure. Adding filler increased the possibility for more
void content within foam and prevented the formation of effective matrix-filler interactions
[118], [122], [122]. Furthermore, it acted as an impurity by providing minimal strengthening
effect to foam, which resulted in the reduction of tensile strength in PU [123].

1% Filler 2.5% Filler 5% Filler


1.8 1.8 1.8
Tensile Strength (MPa)

Tensile Strength (MPa)

Tensile Strength (MPa)


1.6 1.6 1.6
1.4 1.4 1.4
1.2 1.2 1.2
1.0 1.0 1.0
0.8 0.8 0.8
2. %Ch …
lo

ll
C t

Eg in
Ha rol

5% se
Ha rol

ll
Ce ut

Eg in
5% u

he
1% zel l

ll
Ce nut
1% lose

Eg tin

he
2. ellu
5% it
Ha tro

2. zeln

5% hit
he

5% zeln

lo
5% nt

5% ont
gs
1% Chi

gs
llu
2. Co
1% on

C
gs
llu

C
C

(a) (b) (c)


Figure 4.4: Tensile strength of polyurethane foam composites before aging for (a) 1% filler, (b) 2.5% filler, (c) 5% filler

1% Filler 2.5% Filler 5% Filler


Conditioned Tensile Strength (MPa)

Conditioned Tensile Strength (MPa)

Conditioned Tensile Strength (MPa)

1.5 1.5 1.5


1.3 1.3 1.3
1.1 1.1 1.1
0.9 0.9 0.9
0.7 0.7 0.7
0.5 0.5 0.5
ll
Ce ut

Eg in
2. aze l

5% e

Ha rol

5% ose

Eg in
H ro

ll
Ce ut
1% ose

Eg in
Ha rol

ll
Ce ut

he
2. ulos

he
5% Chit
he

5% ln

5% Chit
5% zeln
1% hit

5% nt
1% zeln

5% ont
gs
1% ont

gs
l

llu
gs

2. Co
llu

ll
C

C
C

(c)
2.

(a) (b)
Figure 4.5: Tensile strength of polyurethane foam composites after aging for (a) 1% filler, (b) 2.5% filler, (c) 5% filler

4.3.1.2 Ultimate elongation results analyses of polyurethane foam composites


Ultimate elongation results showed that for 1% loading of PU foam, elongation was sustained
at the control formulation level. Slight reductions of 1-5% were observed for hazelnut,
cellulose, and chitin, whereas eggshell showed improvements of approximately 4.6%. For 2.5%
filler loading, the eggshell elongation was maintained at approximately 1.2%. However,

62
hazelnut reduced by 7% whereas cellulose and chitin reduced by 27%. As the amount of filler
was further increased to 5%, there was a visible decrease in elongation properties with only
eggshell having comparable properties to the control formulation, with a slight reduction of 2%.
A critical observation that was made was the impact of moisture on elongation. Although the
properties decreased as expected, fillers were seen to perform better at sustaining foam
elongation compared to the control formulation. 1% filler showed good resistance to moisture
with 2-7% in hazelnut, cellulose, and eggshell, whereas chitin elongation reduced by 14%.
Similar trends were seen for 2.5% filler for hazelnut and eggshell with 1-5% increase in
elongation compared to the control formulation. However, cellulose and chitin had high
reduction of approximately 23%. Lastly, looking at the results for 5% filler, the results were
similar to 2.5% filler results. Hazelnut and eggshell withheld elongation properties, however,
there was a reduction in cellulose and chitin results.
Although the increase in ultimate elongation was minimal, effective matrix-filler reinforcements
and bonding within PU foam [120] helped attain improvements in eggshell. Furthermore, past
research showed that surfactant used in polymer synthesis has tendencies of forming an
interface between the natural filler and polymer resin, which reduced crystallinity and
increased plasticity of the matrix [123]. Nevertheless, reductions in elongation were significant,
especially before the aging process. The main cause of these reductions was the rigid matrix
filler interactions which resulted in a decrease of polyurethane bonding. Although at low
amounts of filler, matrix-filler interactions did not impact elongation significantly, high filler
content caused an increase in the rigid interface between matrix and filler. Consequently, this
hindered the elastic properties of foam and resulted in lower ultimate elongation [124].

63
1% Filler 2.5% Filler 5% Filler
600.0 600.0 600.0
Ultimate Elongation (%)

Ultimate Elongation (%)

Ultimate Elongation (%)


500.0 500.0 500.0

400.0 400.0 400.0

300.0 300.0 300.0

200.0 200.0 200.0

ll
Eg in
2. aze l

5% e
Ce ut

Ha rol

5% ose

Eg in
ll
Ce nut
1% ose
Ha rol

ll
Ce ut

Eg in

H ro

he
2. ulos

he
he

5% Chit

5% hit
5% ln
1% Chit
1% zeln

5% nt

5% ont
1% ont

5% zel

l
gs

gs
l

llu
gs
llu

C
2. Co

ll

C
C

2.
(a) (b) (c)
Figure 4.6: Ultimate elongation of polyurethane foam composites before aging for (a) 1% filler, (b) 2.5% filler, (c) 5% filler

1% Filler 2.5 % Filler 5% Filler


450.00 450.00
Conditioned Ultimate Elongation %

450.00

Conditioned Ultimate Elongation %


ConditionedUltimate Elongation %

350.00 350.00 350.00

250.00 250.00
250.00

150.00 150.00
150.00
2. aze l

5% e

ll
Ce ut

Eg in

Ha rol

ll
Ce nut
5% ose

Eg tin
H ro

he
he
2. ulos
1% azel l

ll
Ce nut
1% lose

Eg tin

5% Chit
5% ln
H tro

5% nt

5% Chi
5% ont
he

5% zel

l
gs

gs
llu
1% Chi

2. Co

ll
1% on

gs
llu

C
C

2.

(a) (b) (c)


Figure 4.7: Ultimate elongation of polyurethane foam composites after aging for (a) 1% filler, (b) 2.5% filler, (c) 5% filler

4.3.1.3 Young’s modulus results analyses of PU foam composites


The results showed that before aging, adding filler resulted in an increase in the foam’s
modulus. While hazelnut, cellulose, and chitin showed an increase of approximately 25%,
eggshell only showed 10% increase for all fiber loadings. After aging, the modulus values
decreased to approximately 7.5%. This was expected since water and high temperature
weakened the rigidity of the structure, giving it a rubberier integration.
Young’s modulus was used to measure the elasticity of the material. Low modulus value
indicated that lower force was needed to stretch the foam. Understanding the value of young’s
modulus is fundamental in industrial applications. Several studies relating to finite element

64
analysis, aerospace and composite applications, as well as higher performance thin films [125],
[126], [127] reinforce the significance of young’s modulus for industrial applications and high
performance products. While some applications require the material to be more elastic to
increase comfort level, other applications need rigid foam for their applications. The young’s
modulus helped distinguish between these critical application parameters.

1% Filler 2.5% Filler 5% Filler

0.36 0.36 0.36


Modulus (MPa)

Modulus (MPa)
Modulus (MPa)

0.28 0.28 0.28

0.20 0.20
0.20

5% zel l

ll
5% lose

Eg tin
Ce nut
Ha tro
Eg in
2. az ol

ll
Ce ut

he
he
2. ulos
1% ose
Ha rol

ll
Ce nut

Eg in

5% it
H r

5% Chi
5% eln
he

5% nt

5% on

gs
2. %Ch
1% Chit

llu
gs
1% ont

2. Co
1% zel

gs

ll

C
llu

5
C

(a) (b) (c)


Figure 4.8: Young’s modulus of polyurethane foam composites before aging for (a) 1% filler, (b) 2.5% filler, (c) 5% filler

1% Filler 2.5% Filler 5% Filler


Conditioned Modulus (MPa)

Conditioned Modulus (MPa)

Conditioned Modulus (MPa)

0.36 0.36 0.36

0.28 0.28 0.28

0.20 0.20
0.20
5% zel l

5% ose

Eg tin
ll
Ce nut
Ha tro
Eg in
2. aze l

he
ll
Ce ut
H ro
Ha rol

ll
Ce ut

1% ose

Eg in

he
2. ulos

5% hi
5% it
5% ln
he

l
5% on

gs
5% nt

llu
1% hit

2. %Ch
1% zeln

C
gs
1% ont

gs

2. Co

C
llu

ll
C
C

(a) (b) (c)


Figure 4.9: Young’s modulus of polyurethane foam composites after aging for (a) 1% filler, (b) 2.5% filler, (c) 5% filler

65
4.3.1.4 Ball drop resiliency % of polyurethane foam composites
Resiliency results showed that adding 1% filler to foam helped improve the foams resiliency by
10%. The results were similar for all fiber loadings and increasing the amount of fiber showed
trends of improvements in resiliency. Each filler showed slight improvements and there was no
significant difference between the improvements shown by one filler compared to another.
These improvements were due to a more well-connected network within the polyurethane
structure, since fillers have been shown to help dissipate energy more effectively throughout
PU foam upon impact [128] [129]. As the ball impacted the foam, energy was transferred more
effectively through the crystalline structure formed by matrix-filler interactions, as compared to
urethane bonds. As a result, more energy was transferred back into the ball as compared to
absorption by the foam. The results show that even though fillers resulted in void formation
and higher porosity, the matrix filler reinforcements were more effective, and thus, they
resulted in higher foam resilience.
High foam resiliency is critical for designing foam products as a result of the valuable properties
it adds to foam, including comfort and cost-effectiveness [130]. Many applications such as
automotive, cosmetic, and textile amongst other applications depend on high foam resiliency in
order to manufacture high quality products [131] [132]. [133] studied high resiliency in foams
and evaluated that high molecular weight polyol with high reactivity precursors and low
variability in formulation helped to achieve high resilience foams. By making the blown
polyurethane rubberier, the resiliency of foam was improved [133].

66
1% Filler 2.5% Filler 5% Filler

42.0 42.0 42.0

Resiliency %

Resiliency %
Resiliency %

38.0 38.0
38.0
34.0 34.0
34.0
30.0 30.0
30.0

Eg in
2. aze l

ll
Ce ut
H ro

Ha rol

ll
5% se

Eg in
Ce ut
he
2. ulos
5% it

he
5% ln
Ha rol

ll
Ce ut

1% se

5% it
Eg in

5% nt

5% zeln

lo
2. %Ch

5% ont
he

gs

Ch

gs
1% hit
1% zeln

lo

llu
2. Co
1% ont

ll
gs
llu

C
C

5
C

(a) (b) (c)


Figure 4.10: Ball drop resiliency % of polyurethane foam composites for (a) 1% filler, (b) 2.5% filler, (c) 5% filler

4.3.1.5 Tear strength of polyurethane foam composites


Tear strength of the composites was seen to decrease. Hazelnut based PU had higher reduction
in strength compared to the other composites, with a 16% reduction for all fiber loadings.
Eggshell had a similar amount of reduction (approximately 14%), followed by cellulose which
had 13% reduction, and chitin with only slight reductions of 2-5%.
The results showed that although tensile strength of the material improved, it resulted in the
reduction of tear loading strength. Studies have shown that elastomers reinforced with fillers
can stiffen and strengthen the elastomer, but also result in anisotropy of strength and
subsequent splitting apart of the sample with more ease than the control formulation [134],
[135], [136], [137]. Unlike tensile strength of a material, the tear strength is not controlled by a
single high stress concentration. Depending on the radius of the hole proceeding the crack tip,
the tear strength can change. If the hole is bigger than the crack tip, higher tear strength is
required and vice versa [138]. Although the addition of filler usually helps in increasing surface
roughness of PU, if the fiber size is too small compared to natural tear roughness, it can result
in little influence on the tear strength of the material [139]. The reduction in tear strength
showed that this was a possible cause of reduction for the PU composites in figure 9. As
roughness did not play a significant role and due to the anisotropic strength distribution within
the elastomer, it was possible that the crystalline nature of filler-matrix interactions led to a
smoother surface and consequent reduction in tear strength of the polymer.

67
1% Filler 2.5% Filler 5% Filler
2.0 2.0 2.0

Split Tear (N/mm)


Split Tear (N/mm)
Split Tear (N/mm)

1.8 1.8 1.8

1.6 1.6 1.6

1.4 1.4 1.4

1.2 1.2 1.2

Ha rol

ll
5% se

Eg in
Ce ut
ll
Ce ut

Eg in
2. aze l

e
1% ose
1% zel l

ll
Ce nut

Eg tin

H ro
Ha tro

he
he
2. ulos
he

5% hit
5% it

5% zeln

lo
5% ln

5% ont
5% nt
1% hi

2. %Ch

gs
gs
l
1% on

llu
gs
llu

C
C

2. Co

ll

C
C

5
(a) (b) (c)
Figure 4.11: Split tear results measuring the tear loading strength of polyurethane foam composites for (a) 1% filler, (b) 2.5%
filler, (c) 5% filler

4.3.1.6 Compression set % of polyurethane foam composites


Compression set results showed that there was no particular trend of increasing or decreasing
in the permanent deformation of foam. While some composites showed approximately 4%
compression set, other composites showed just 1% compression. Nevertheless, the critical
observation that was noted was that fillers did not worsen the permanent deformation
properties of foam. High compression set is an indication of higher chain slipping, resulting in
permanent deformation [140]. The results showed that although filler loading did not prevent
chain slipping, the polymer and matrix-filler interactions helped inhibit chain slipping from
taking place.
Knowing the compression set of foam is essential in order to ensure a long product life cycle. If
the compression set % is high, it indicates that the foam material is degrading and losing its
performance. When an elastomer is deformed using high pressure, it may not recover
completely, but instead has permanent deformation [141], [142]. [143] studied the effects of
temperature on compression set of elastomers, and the results showed minimal compression
set between 20oC-50oC. As the temperature reduced below 20oC or increased above 50oC, the
compression set % increased at an exponential rate to 100% compression, at -60oC and 160oC.
Studies show that elastomers are widely incorporated into applications where knowing their
compressive strength is very significant. From use in shoe soles to structural dampers, as well as

68
impact-based equipment, the use of low compression set polyurethane foam is in high demand
[144], making it a vital parameter for high quality applications.

1% Filler 2.5% Filler 5% Filler

4.0

Compression set %
Compression set %

4.0

Compression set %
4.0
3.0 3.0 3.0
2.0 2.0 2.0
1.0 1.0 1.0
0.0 0.0 0.0

Eg tin
2. Haz ol

5% se

ll
Ce nut

5% lose
5% zel l

ll
Ce nut

Eg tin
Ha rol

ll
Ce nut
1% ose

Eg tin

Ha tro
he

he
he

tr

2. lulo
5% Chi

5% Chi
1% hi
1% ont

5% el
5% n

gs

5% on

gs
1% zel

llu
gs
llu

2. Co
C

C
C

2.
(a) (b) (c)
Figure 4.12: Compression set% of polyurethane foam composites measuring permanent deformation in thickness of (a) 1% filler,
(b) 2.5% filler, (c) 5% filler

4.3.1.7 Foam hardness of polyurethane foam composites


Producing high quality foam products is one of the most important aspects for industries,
particularly in applications where cushioning and comfort are of critical importance. Certain
qualifications have to be met, including economic production, safety standards, durability, etc.
The results for hardness indicated that as more filler was added to foam, the overall hardness
was reduced. This was a result of reductions in polyurethane bonds due to filler interactions,
which resulted in softer bonds compared to urethane linkage. Compared to the control
formulation, there was an overall decrease of approximately 10% on average. However,
hazelnut was shown to increase the resiliency slightly by 1%, for 1% fiber loading. As the fiber
amount was increased to 2.5%, the trends looked similar, with hazelnut showing similar
hardness compared to the control formulation, while other composites produced softer foam.
For 5% fiber loading however, there was a noticeable increase in hardness. The reason for this
could be due to saturation of the filler. At low amounts, hardness was dominated by urethane
linkage, and the filler reduced the amount of urethane bonds causing reduction in hardness.
However, filler was evidenced to dominate hardness when 5% filler was added.
Hard or soft foam material is manufactured based on the nature of the application in question.
For applications in the textile or automotive industry, where the foam’s cushioning capabilities

69
are important, studies have shown maximum durability, physical strength, and comfort are
obtained through high density, high resilience, and soft foams. Consequently, fillers provide an
outlet to achieve low foam hardness while maintaining density, elongation, and good tensile
properties.

1% Filler 2.5% Filler 5% Filler


40.0 40.0
40.0

Hardness (Asker)
Hardness (Asker)
36.0 36.0
Hardness (Asker)

36.0
32.0 32.0 32.0

28.0 28.0 28.0

24.0 24.0 24.0

5% zel l

ll
Ce nut
5% lose

Eg tin
2. Haz ol

5% se

ll
Ce nut

Ha tro
Eg in

he
Ha rol

ll
Ce nut
1% ose

Eg tin

he
r

5% Chit
he

2. ulo

5% Chi
5% nt

5% on

gs
5% el

llu
gs
1% hi
1% ont
1% zel

gs

2. Co
llu

ll
C

C
C

2.
(a) (b) (c)
Figure 4.13: Hardness of polyurethane foam composites for (a) 1% filler, (b) 2.5% filler, (c) 5% filler

4.3.2 Thermal analyses of polyurethane foam composites


Degradation of polyurethane and its composites is shown in the TGA curve below (figure 4.12).
Decomposition was seen to occur in two steps. The first step involved degradation of hard
segments including urethane bonds and filler. This took place at a temperature of
approximately 150oC-340oC. The second stage involved degradation of soft segments such as
ester groups, occurring at approximately 420oC to 600oC [145], [146], [147],. The results were
as expected, with the composites displaying similar degradation trends as that of the control
formulation. Slight differences in thermal degradation were observed which were believed to
be caused by runtime errors and formulation discrepancies.
DSC results are shown in figure 4.13. The glass transition temperature was the point at which
polyurethane transitioned from glassy to a rubbery state. The figure showed that the glass
transition temperature of the composites was approximately -38.4oC and the value remained
unchanged with the addition of fillers. This was expected as the filler did not influence the foam
structure during the reaction.

70
300

Control
200
1% Hazelnut
2.5% Hazelnut
100 5% Hazelnut
Glass transition temperature: -38.4oC 1% Cellulose
Heat Flow (W/g)

0 2.5% Cellulose
-60 -50 -40 -30 -20 -10 0 5% Cellulose
Temperature (oC) 1% Chitin
-100
2.5% Chitin
5% Chitin
-200 1% Eggshell
2.5% Eggshell
-300 5% Eggshell

-400

-500

Figure 4.14: Differential Scanning Calorimetry (DSC) results used to measure the glass transition temperature of polyurethane
foam composites

100
90
Control
80
1% Hazelnut
70 2.5% Hazelnut
5% Hazelnut
60 1% Cellulose
Weight %

2.5% Cellulose
50 5% Cellulose
1% Chitin
40 2.5% Chitin
5% Chitin
30
1% Eggshell
20 2.5% Eggshell
5% Eggshell
10
0
0 100 200 300 400 500 600 700 800 900
Temperature (oC)

Figure 4.15: Thermogravimetric Analysis (TGA) results measuring the thermal degradation of polyurethane foam composites

71
4.3.3 Morphological analyses of polyurethane foam composites
Results showed that porosity of cells decreased slightly upon the addition of 1% filler (by
approximately 1%-2%). As more fiber was loaded, a slight increase in hazelnut and eggshell
porosity was seen. Although formulation discrepancies had an impact on higher or lower cell
porosity, an important aspect was that the addition of fillers did not influence foam porosity.
Although fillers may impact the porosity and increase it [118], however studies have shown that
fillers can also help bind the cells, reinforcing its structure and preventing cell coalescence
[148].
Closed-mold controlled foaming reaction resulted in the formation of spherical cells and less
collapse throughout its foam morphology. Free-rise foam has been reported to form irregular
and uneven cell structure due to unstable foam expansion rate due to CO2 gas [149]. In closed-
mold conditions, the cellular structure became more homogeneous due to a gradual increase in
cell size as a result of the pressurized system inside the mold during the foaming reaction [149],
[150]. The SEM images produced show how each composite looks. Aside from slight variations
in formulation, it was noted from the graphical result and linear regression model that the
average cell size remained unchanged while varying filler composition and formulation. It
showed that despite the impact of filler on urethane interactions, there was no impact on the
average cell size of polyurethane. Moreover, the fiber images extracted below show how each
fiber interacted with the cells, piercing through the cell wall and interconnecting with several
cells.
Measuring the porosity alongside the SEM images was critical in understanding foam
morphology and the effect of fillers. The uniform cell structure and porosity indicated the
compatibility nature of the organic fillers used, in improving mechanical results for PU foam for
high performance industrial applications.

72
Cell Size (µm) Porosity %

100
150
200
250
94
96
98
100
1% Con
H tr 1% Con
1% aze ol Ha trol
Ce lnu 1% ze
llu t Ce lnut
llu

(a)
1% los

(a)
1% lose
1% Ch e

1% Filler
1% Ch
1% Filler

Eg itin
gs Eg itin
he gs
ll he
ll

Cell Size (µm) Porosity %

100
150
200
250
94
96
98
100

2. C 2. Co
5% on 5% n
2. Ha trol t
5% ze 2. Haz rol
5% e
Ce lnu Ce lnut
t
(b)

73
2. llulo
5% s 2. lulo
5% se
5% filler
2. e 2.

(b)
5% Ch 5% Ch
Eg itin
2.5% Filler
Eg itin
2.5% Filler

gs gs
he he
ll ll

Porosity %
Cell Size (µm)
94
95
96
97
98
99
100

100
150
200
250

5% Con 5% on
H tr H tro
5% aze l
(c)

5% aze ol
Ce lnu Ce lnut
llu t llu
5% los 5% lose
e

(c)
5% Chi
5% Filler

5% Ch
5% Filler

i
Eg tin Eg tin
gs gs
he

Figure 4.17: Average cell size of polyurethane foam composites measured using Scanning Electron Microscope (SEM)
he ll
ll
Figure 4.16: Ultrapycnometer results measuring the porosity% of polyurethane foam composites(a) 1% filler, (b) 2.5% filler, (c)
Control: 185.1μm 1% Hazelnut: 181.34 μm 1% Cellulose: 161.3 μm 1% Chitin: 221.84 μm

Foam void/ crack

1% Eggshell: 160.04 μm 2.5% Hazelnut: 159.61 μm 2.5% Cellulose: 168.32 μm 2.5% Chitin: 191.19 μm

Foam void/ crack

Foam void/ crack

2.5% Eggshell: 161.60 μm 5% Hazelnut: 194.31 μm 5% Cellulose: 152.38 μm 5% Chitin: 215.71 μm

Foam void/ crack


Foam void/ crack

5% Eggshell: 170.49 μm

Foam void/ crack

Figure 4.18: Scanning Electron Microscope (SEM) of polyurethane foam composites

74
Eggshell fiber Hazelnut fiber Chitin fiber

Figure 4.19: Scanning Electron Microscope (SEM) images of eggshell, hazelnut, and chitin fiber

4.4 Conclusions
In this study, mechanical, thermal, and morphological properties of PU composites with 1%,
2.5%, and 5% of hazelnut, cellulose, chitin, and eggshell were investigated. Results showed that
natural fillers had good compatibility with PU foam and are viable sources of replacement to
synthetic fibers in the near future. 1% fiber loading showed significant improvements in tensile
strength, whereas the strength decreased upon excess loading. Low filler content also produced
good elongation properties. Moreover, higher strength was recorded post hydrolysis for PU
composites as compared to the control formulation, and resiliency % also increased.
Nevertheless, the results showed slight reductions in split tear strength. The thermal and
morphological analyses carried out indicated a uniform structure of PU composites and
effective foaming. Due to the simple reaction mechanism and filler processing, the closed-mold
fabrication process can be easily scaled up to produce more eco-friendly foam products. Lastly,
this research helped provide valuable insight on the nature of organic fillers that were tested,
which will help industries in manufacturing bio-based foams for a more sustainable tomorrow.

75
Influence of reagents and additives on the mechanical properties of
closed-molded polyurethane foams
Abstract
This study aims to achieve an understanding of the effects of reagents on polyurethane foam plaques. In
present times, the use of polyurethane foam products has dominated industries, and there is a vital
need to develop a fundamental understanding of how different additives and material compositions
impact foam properties. Herein, a control formulation of polyurethane (PU) foam with known
mechanical properties was compared to PU fabricated with slight variations in the formulation. The
fabricated foam was tested for its mechanical properties and morphology. The analyses helped provide
more understanding to the effects of formulation changes. The results showed good tensile and tear
properties for glycerol and silica. However, the resiliency dropped significantly showing low comfort and
durability. Nevertheless, this study helped indicate that using a lower purity polyol than what was
currently used gave similar mechanical properties to the control formulation. Moreover, the results also
indicated how slight additions of foreign reagents/ additives significantly impacted foam properties,
making it essential that foam formulations be prepared using very precise measurements; in order to
reduce the variability in fabricated foam results.

5.1 Introduction
As the world progresses towards more sustainable production, the need to develop eco-friendly
techniques of polyurethane fabrication have also increased. In recent times, a great deal of research
interest has been gained for the use of renewable resources to produce industrial products [151].
Significant research and development work has been focused on polyurethane foams due to their
numerous advantages and impact on human lives [152]. One of the main advantages of polyurethane
foam is the wide range of polyurethane grades that can be formulated. Different types of polyurethanes
can be produced by using the same major raw materials (e.g. polyol and isocyanate) while varying the
amount of moisture, blowing agent and other reagents [153]. However, a major drawback to
polyurethane foam is that relatively large variations in properties can occur with similar formulations,
which cannot be attributed to errors or machine malfunctions [153].

76
Globally, PU usage has been estimated at approximately 60.5 billion USD in 2017, and the number is
expected to surpass 79 billion USD by 2021 [154]. The versatility of PU foams arises from their wide
range of properties that are achievable through small changes in formulations. For instance, increasing
polyol functionality without changing the molecular weight increases foam hardness while reducing
tensile strength, tear strength, and elongation [155]. Similarly, varying the amount of blowing agent can
impact foam expansion and cell density. Two types of blowing agents are mainly used for foam
fabrication. These include physical blowing agents which expand the polymer by vaporization, and
chemical blowing agents which expand the polymer by the CO2 produced. [156] researched on the use
of blowing agents and their effect on polyurethane foaming. The results showed that using chemical
blowing agents resulted in shorter maximum rise heights and high temperature of the reaction mixture.
However, the use of physical blowing agents had no significant impact on the dielectric polarization and
temperature of the reaction mixture.

Reagents such as surfactants, cross-linkers, and catalysts also have a significant impact on polyurethane
foam properties. Surfactants lower the surface tension while emulsifying incompatible ingredients
present within the reactants. In addition to this, they also have a significant impact on cell nucleation
and stability, which effects cell size and air permeability [157]. [158] researched on the most effective
amount of surfactant to improve foam properties. The results showed that the data gathered was not
sufficient to extrapolate the work to other foam systems, and consequently, more extensive studies are
needed to further knowledge of surfactants. [151] researched on increasing crosslinker content in the
PU foam and showed that increasing the amount of crosslinker increased the number of hard segments
present within the foam. Consequently, the foam’s tensile properties were seen to improve while the
glass transition temperature had a relatively low value. Catalysts have also been observed to impact the
foam’s reaction temperature and expansion rate. Research shows that the combination of reactants
results in a very sensitive foaming reaction, where precision and effective formulation are key to getting
good polyurethane foam.
Polyurethane foams have a very diverse set of properties that can be determined by manipulating the
stoichiometric balance of its reactant components [159]. Manipulating the factors impacts their
chemical, physical, and thermal properties and so, a good understanding of the effects of reagents is
vital for foam fabrication. In previous decades, petroleum-based polyols and CFCs dominated the
polyurethane fabrication sector [160] [161]. As global regulations have grown stricter, the need to

77
research on eco-friendly foam fabrication processes has become vital. For this purpose, this study aims
to provide a stress loading analyses of PU foams upon slight changes/ additions in formulation. To the
best of our knowledge, there is no comparative mechanical study of the effect bio-friendly reactants
have on a controlled polyurethane foam. The foams produced for this study are tested according to
their mechanical properties, including tensile, elongation, modulus, and tear resistance. Successful
implementation of this study is vital to large scale foam production; where understanding the subtleties
of varying the formulation can be critical in developing high quality foam products.

5.2 Experimental
5.2.1 Materials selection
Raw materials used in this research include natural fillers (namely lignin, polysaccharide,
hazelnut, chitin, and chitosan), polyol, additives, and isocyanate. Lignin was obtained through
Dotmar, hazelnut from Nut Nursery, polysaccharide from Dupont, and chitin and chitosan from
Tidal Vision.
The polymer used for this research was free-rise flexible PU foam. It was fabricated using polyol
supplied by a collaborative company for this research. Mondur CD isocyanate was used as
additive B (additive A being polyol) for the PU addition reaction. Other reagents, including the
surfactant, catalyst, and blowing agent were also added to the foam, obtained from Sigma
Aldrich. Due to company confidentiality agreements, exact reagent compositions are not
mentioned.

5.2.2 Preparation and analyses of natural filler for polyurethane fabrication


Each natural filler was mechanically grinded using a willy mill lab grinder. The powder obtained
through grinding was sieved using a 200µm sieve in order to achieve fine powder. This ensured
that a high amount of filler surface area came into contact with the matrix, which helped
maintain effective filler matrix interaction. The filler was prepared in weight compositions of
1%, 2.5%, and 5% in order to study the effect of each filler’s composition on PU as well as to
study the threshold amount before the filler became oversaturated in the foam and resulted in

78
poor foam properties. Neat PU was also prepared alongside the filler-based PU foams; to
compare the effects of filler on PU foam.

5.2.3 Procedure for polyurethane foam composite synthesis


PU foam was fabricated through an addition reaction. First, the reagents, consisting of catalyst,
surfactant, compatibilizer, and blowing agent were prepared in their respective amounts and
added into the polyol. The polyol was then heated to reduce its viscous effects and isocyanate
was added while mixing at high shear rates. After approximately 10 seconds of mixing, the
material was poured into a mold for the foaming reaction to take place. Once poured, foam
was left to expand and cure for 24 hours before testing. For this study, silica, chain extender,
and glycerol were added at weight% of 2-5% and the results were normalized according to their
density. The following schematic below shows this process:

Figure 5.1: Polyurethane foam synthesis reaction

5.2.2 Characterization of rigid polyurethane foam composites


5.2.2.1 Procedure for mechanical analysis of polyurethane foam composites
Mechanical analyses of foams were based on three stress loading analyses. These included
tensile testing to measure the maximum stress the foam was able to withstand before
breaking, and elongation % which measured the maximum elastic capacity of foam without

79
breaking. Tensile testing was done according to ASTM D3574-17 testing method. Type IV sized
dog bones were used for the test and the test was performed using a Microtester 5848
(Instron) machine and five dog bone samples were tested at a rate of 500 mm/min. Tensile
stress along and ultimate elongation % results were extracted. Foam modulus was calculated
using the stress and strain data that was obtained, in order to gain a better understanding of
foam elasticity.
Lastly, tear testing was carried out, which measured the tear propagation resistance of foam.
This was also performed using ASTM D3574-17 testing method and five samples were tested
using the Microtester 5848 (Instron) machine.
The following reactions show how silica, glycerol, and chain extender influenced the
polyurethane matrix; providing it with more rigidity and increasing the mechanical properties:

80
Figure 5.2: Reactions involved for additives within polyurethane

5.3 Results and Discussion


Tensile strength results showed that adding silica and glycerol maintained or slightly improved
the tensile strength of polyurethane. Moreover, using low purity polyol also showed that it
maintained good foam tensile strength. This result showed cost-saving potential as it highlights
the possibility of fabricating less pure polyurethane at a cheaper rate to form polyurethane
polymer with identical mechanical properties. However, it was shown that adding high chain
extender to foam resulted in a decrease in the tensile strength.
The main reason for this decrease in strength was due to the chain extender causing a
reduction in the matrix viscosity. The reduction resulted in higher expansion rates and lower
average bond-density to resist tensile loading [162]. Silica was added to PU as it is reported to
function as a multi-functional cross-linker as well as a reinforcing filler that can influence the
mechanical properties of foam significantly [163]. The results conformed to this study of silica
particles. Adding silica helped improve the tensile strength due to the strengthening effects it
provided to PU. Lastly, glycerol was added in low amounts to PU to understand how it impacts
the foam’s tensile properties. Since glycerol itself is a renewable polyol, it has functionality to

81
react with the isocyanate and produce interactions [14]. The results conformed with this theory
and helped improve the tensile properties slightly, by 4%.
Figure 3 shows the results for ultimate elongation of polyurethane with adjusted formulations.
The results showed a 38% decrease in silica elongation whereas glycerol showed improvements
of approximately 35%. Adding low purity polyol and high chain extender helped maintain
ultimate elongation property.
It is hypothesized that the decrease in ultimate elongation % using silica was due to an increase
in the rigid segments present within foam [163]. The increase in rigid segments reduced foam’s
ductility and as a result, it was seen to reduce elongation significantly. On the other hand, the
significant improvements in glycerol showed effective interactions between isocyanate and
hydroxyl groups of glycerol, which enhanced the foam’s elasticity to much higher than the
control formulation [164]. Finally, chain extender was also seen to withhold ultimate elongation
of foam. This was a result of improved viscosity and foam expansion rates as a result of the
chain extender, which helped maintain foam elongation properties even though the tensile
strength decreased. Nevertheless, it was important to note that the ultimate elongation was
not improved by adding chain extender.
The young’s modulus was extracted using the tensile strength and strain value obtained
through the tensile test. The results showed that adding glycerol resulted in an increase in
rigidity (shown by the high Young’s modulus compared to the control). This was expected and
seen in past research studies as well, for instance [163], where silica particles evidently
increased the rigidity of foam and improved the mechanical properties. Low purity polyol and
high chain extender maintained the same young’s modulus as control. Chain extender
maintained the modulus value because even though tensile strength reduced, the ultimate
elongation value slightly increased which resulted in a similar modulus compared to control.
Lastly, glycerol-based PU reduced the modulus, indicating more elasticity of the foam. This is
explained by the different chemical structure of glycerol-based PU, which helped reduce
elasticity.
Figure 5 shows tear propagation resistance of foam. The results indicate that adding silica,
glycerol, and chain extender improved the foams tear strength significantly. Each of the foams

82
showed improvements in normalized tear strength by more than 50% whereas low purity
polyol gave slightly lower tear resistance results which are likely to be attributed to formulation
discrepancies rather than bond strength, as the tensile strength and elongation were seen to be
similar to the control formulation.
These improvement were caused by effective bonding and interconnection of polyurethane
foam through silica, glycerol, and chain extender [163], [162] [164].

5.0 2500.0

Normalized Ultimate Elongation (%/gcm-3)


Normalized Tensile Strength (MPa/gcm-3)

2000.0
4.0
1500.0

1000.0
3.0
500.0

2.0 0.0
Control PU with PU with PU with PU with Control PU with PU with PU with PU with
silica glycerol low high silica glycerol low high
purity chain purity chain
polyol extender polyol extender
(a) (b)

83
14.0

Normalized Tear Strength (N/mmgcm-3)


12.0
0.8
Young's Modulus (MPa)

10.0

8.0
0.4
6.0

4.0

0.0 2.0
Control PU with PU with PU with PU with Control PU with PU with PU with PU with
silica glycerol low high silica glycerol low purity high chain
purity chain polyol extender
polyol extender (d)
(c)
40

37
Rebound Resiliency%

34

31

28
Control PU with PU with PU with PU with
silica glycerol low purityhigh chain
polyol extender
(e)
Figure 5.3: Mechanical properties of polyurethane foam plaques (a) Normalized tensile strength, (b) Normalized
ultimate elongation, (c) Young’s modulus, (d) Normalized tear strength, (e) Ball drop resilience %

Ball drop resiliency of foam was measured, and the results indicated that there was a drop in
resilience of each foam. Although silica and chain extender only reduce resilience by
approximately 7-9%, glycerol-based PU reduced the resiliency % by approximately 27%. This
shows that adding glycerol significantly impacted the force imparted throughout the foam

84
structure. The rigidity had a negative impact and absorbed more impact rather than
transferring back the impacted force to the ball. This result was critical as it showed that
although glycerol showed high improvements in tear strength as well as tensile and elongation,
resiliency dropped significantly which can greatly reduce the durability and comfort of glycerol-
based PU.

5.4 Conclusions
The results showed meaningful observations for tensile and tear loading properties of PU foam.
Glycerol based PU showed improvements in tensile, elongation, and tear properties.
Nonetheless, the ball drop resiliency decreased significantly due to poor force transfer back to
the ball. Similarly, silica also showed improvements in the tensile and tear properties. However,
the ultimate elongation reduced significantly, and resilience showed a slight decrease.
Moreover, although chain extender improved the tear strength of foam, tensile strength and
resilience reduced. Nevertheless, there were two significant discoveries that were made from
this study. Firstly, the results showed that using old polyol produced similar results to those of
control formulation. This showed the potential adoption of lower purity polyol in order to
reduce costs. Secondly, the results indicate how sensitive foam reactions can be, and how the
addition of very small quantities of additives can have significant impacts. This exemplifies the
importance of understanding foam chemistry at a very deep level and ensuring precise
formulation preparations in order to achieve high quality foam.

85
6.1 Conclusions and Future Work
6.1.1 Concluding remarks
This work was aimed at enhancing the properties of polyurethane (PU) foam by adding natural
fillers. Furthermore, the focus of this project was to understand how different types of natural
fillers can influence the mechanical, thermal, hydrolytic, and morphological properties of foam.
Each composite PU prepared was characterized according to a diverse range of mechanical
properties, from tensile strength to the foam’s elongation, tear strength, and resiliency%. To
fully understand the nature of thermoset polyurethane reactions, polymer chemistry was
altered, and mechanical impact was measured. Moreover, a diverse range of reinforcing natural
fillers, from chitin and chitosan, to plant-based fillers such as cellulose and polysaccharide were
studied to develop an understanding of how they impact foam properties.

Firstly, free-rise foam buns were targeted with addition of natural fibers. Chitin, chitosan, lignin,
polysaccharide, and hazelnut were incorporated into PU foam. The free-rise foaming technique
helped understand the compatibility of fillers with PU matrix before carrying out more
extensive experiments on closed-molded plaques. The second study focused on closed-molded
foaming with addition of hazelnut, cellulose, chitin, and eggshell fillers. The composites were
characterized according to their mechanical, thermal, and morphological properties to compare
enhancements in foam composites that were tested. Lastly, the foam formulation was targeted
for enhancements by adding different reagents and additives. These plaques were mechanically
tested to understand the effect of altering foam formulation on stress loading and resiliency
results.

From the first study, 1% chitin, polysaccharide, and hazelnut showed improvements in tensile
strength and elongation due to effective matrix filler interactions while lignin and chitosan
showed a decrease in mechanical properties. Moreover, 1% hazelnut showed an increase in
foam resilience, relating to good durability and comfort. Chapter 4 results showed similar
findings for closed-molded plaques. While 1% fiber loading showed significant improvements in

86
tensile strength, elongation, and resiliency, the split tear propagation saw a decrease in
strength. Lastly, chapter 5 discussed the effects of adding/ altering the formulation on PU foam.
The results showed good tensile and tear properties for silica and glycerol addition whereas the
resiliency % dropped significantly. Consequently, the results indicate the cost saving potential
of natural fibers while maintaining/ enhancing mechanical properties. The study in chapter 5
showed the sensitive nature of PU foaming reactions and the need to fabricate foam using
extremely precise formulation techniques; in order to mitigate errors and produce high quality
products. Ultimately, the use of natural fillers has shown great potential as a substitute to
microplastics. Moreover, the use of natural resources is seen to produce good mechanical
properties which reinforces the notion to replace non-renewable resources with renewable
ones and, as a result, conserve the natural environment.

6.1.2 Fundamental scientific contribution


The results have shown the potential for hazelnut shell, cellulose, chitin, and eggshell for use in
closed molded polyurethane foams for large scale applications. Aside from the significant
increase in tensile strength, critical contributions of natural fibers to industries is reinforced by
the 10% increase in resilience, particularly the positive influence of adding hazelnut shell and
eggshell by-products. Compared to control formulation, consistent increase in results were
shown for all-natural fiber-based composites (shown in figure 4.9). The results are critical for
scale up applications as higher resilience shows improvements in durability and comfort of the
foam fabricated.
Moreover, the results show potential for maintaining a detailed design of experiment
approach; by studying the effects of each factor in greater depth through high factorial designs.
The figures below highlight the design of experiment approach taken for experimentation
performed in these studies using an averaged 95% confidence interval and shows how each
factor effected foam mechanical properties. The results show potential for reverse engineering
foam formulations and obtain tailored mechanical properties for polyurethane foam:

87
1 1
Tensile Strength (MPa)
1% low high low

Ultimate Elongation %
1% low high
high high

5% low low high 5% low high


0 0
Natural Glycerol Silica Chain Natural Glycerol Silica Chain
filler Extender filler Extender

(a) (b)

1 1

Tear Strength (N/mm)


1% low low low low high high high
Resilience %

high high

0% low low low


high high
0 0
Natural Glycerol Silica Chain Natural Glycerol Silica Chain
filler Extender filler Extender

(c) (d)
Figure 6.1: Effects of additives on (a) tensile strength, (b) ultimate elongation %, (c) Resilience %, (d) tear strength

6.1.3 Future work


There are several aspects which should be studied as a continuation and future work for this
project.
Firstly, each natural filler comprises of unique functional groups which have property
enhancement capabilities. However, research has shown that some natural fillers have high
intermingling of chains and need treatment to functionalize the bonds. Consequently, a deep
understanding of each filler needs to be carried out in order to enhance its properties and
further improve matrix-filler interactions. Moreover, understanding the nature of matrix filler
interactions is also vital to understand strength improvements and filler behaviour with the
matrix. This understanding will help control the foam formulation to generate high
performance PU for scale up industrial applications.

88
Secondly, filler orientation needs to be studied more thoroughly. Past research has shown that
the filler loading direction can significantly impact the tensile strength of a material. Therefore,
controlling the direction of fiber loading for closed-molded polyurethane plaques will set a
good standpoint in understanding fiber reinforcement effects at a deeper level.
Moreover, based on the findings of this study, we see that adding natural fillers as well as other
additives including glycerol and silica, has helped improve the foam’s mechanical properties.
Using the test results discussed from this study, and by carrying out more experiments on the
effects of fillers and additives, it is of interest to reverse engineer and develop a model to
predict the behaviour of polyurethane upon addition of fillers and different additives. It is also
important to identify how the mechanical properties can be tailored in order to achieve the
desired foam properties, for instance, improving the resilience of foam by adding natural filler
(shown in chapter 4) while also increasing the tear strength with the help of silica (shown in
chapter 5). This will be achieved by analyzing these additives using a thorough design of
experiments approach to understand the effects of each natural filler and additive.
Lastly, polyurethane foams have been known to persist in the environment and accumulate in
landfills. Due to their long lifespan, they can cause harm to living organisms, especially if they
have microplastic additions within them. Degradation of polyurethane is a vital topic that needs
more research and development. By incorporating fillers to enhance degradation and carrying
out degradation tests such as soil burial experiments, the results will help revolutionize the
polymer industry and bring forth biodegradable foams that conserve the environment.

89
References

[1] "Polyurethane Global Market Size Forecast," July 2019. [Online]. Available:
https://www.grandviewresearch.com/press-release/global-polyurethane-pu-market.

[2] K. H. Badri, "Biobased Polurethane from Palm Kernel Oil-Based Polyol," Intech, 2012.

[3] J. G. M. S. C. M. B. T. Rachid Dris, "Synthetic fibers in atmospheric fallout: A source of


microplastics in the environment?," Marine Pollution Bulletin, vol. 104, no. 1-2, pp. 290-
293, 2016.

[4] A. L.-S. D. C.-A. L. I.-R. M. S. Sanchez-Adsuar, "Influence of the Nature and the Content of
Carbon Fiber on Properties of Thermoplastic Polyurethane–Carbon Fiber Composites,"
Journal of Applied Polymer Science, vol. 90, no. 10, pp. 2676-2683, 2003.

[5] H. L. H. D. S. R. P. B. Dahlke, "Natural Fiber Reinforced Foams Based on Renewable


Resources for Automotive Interior Applications," Journal of Cellular Plastics, vol. 34, no.
4, pp. 361-379, 1998.

[6] S. A. Ernie Suzana Ali, "Bionanocomposite hybrid polyurethane foam reinforced with
empty fruit bunch and nanoclay," Composites: Part B, vol. 43, no. 7, pp. 2813-2816,
2012.

[7] Y. G. D. N. W. C. P. H. S. Z. L. V. D. R. Yan Hong Hu, "Rigid Polyurethane Foam Prepared


from a Rape Seed Oil Based Polyol," Journal of Applied Polymer Science, vol. 84, no. 3,
pp. 591-597, 2001.

[8] C. P. N. R. ST Lee, Polymeric foams: Science and Technology, Taylor and Francis Group ,
2007.

[9] S. M. M. Khosrow Khodabakhshi, "Polyurethane Polymers: Composites and


Nanocomposites, Chapter 8.," in Composites and Nanocomposites of PU Polymers Filled
with Natural Fibers and their Nanofibers, Elsevier, 2017.

90
[10] I. A. A. D. Nurain Johar, "Extraction, preparation and characterization of cellulose fibres
and nanocrystals from rice husk," Industrial Crops and Products, vol. 37, no. 1, pp. 93-99,
2012.

[11] C. B. P. a. N. R. Shau-Tarng Lee, Polymeric Foams: Science and Technology, Taylor &
Francis, 2006.

[12] N. a. S. J. Rivier, Foams and Emulsions, Kluwer, Dordrecht, The Netherlands, 1999.

[13] A. H. a. R. M. T. Midgley, "Heat Transfer". US Patent 1,833,847, 31 July 1968.

[14] A. H. a. R. M. T. Midgley, "Manufacture of Aliphatic Fluoro Compounds". U.S. Patent


1930129, 1933.

[15] J. Wei, "The Third Paradigm of Chemical Engineering: Molecular Product Engineering,"
New Jersey Institute of Technology, 2001.

[16] F. Z. Eram Sharmin, "Polyurethane: An Introduction," in Polyurethane, InTech, 2012, pp.


4-16.

[17] M. E. ,. R. A. ,. S. C. &. B. B. Myriam Desroches, "From Vegetable Oils to Polyurethanes:


Synthetic Routes to Polyols and Main Industrial Products," Polymer Reviews, 2012.

[18] K. R. D.K. Chattopadhyay, "Structural engineering of polyurethane coatings for high


performance applications," Progress in Polymer Science, vol. 32, no. 3, pp. 352-418,
2007.

[19] M. P. S., Polymer Chemistry an Introduction, Oxford : 3rd Edn. New York: Oxford
University Press, 1999.

[20] M. G. L. Ashby, Cellular Solids, Pergamon, Oxford, London, 1988.

[21] K. t. N. D.W. Van Krevelin, Properties of Polymers, New York: Elsevier, 1990.

[22] D. V. Krevelin, Properties of Polymers, New York: 3rd edition Elsevier, 1990.

[23] K. Kolossow, Plastics Extrusion Technology, Hanser, Munich: Fourth Edition, 1988.

[24] M. I. A. I. R. N. E. Marcovich, "Microfoams based on castor oil polyurethanes and


vegetable fibers," Journal of Applied Polymer Science, vol. 105, no. 5, pp. 2791-2800,
2007.

91
[25] M. K. L. L. S. A. B. M. K. K. B. J. R. Milena Zieleniewska, "Development and applicational
evaluation of the rigid polyurethane foam composites with egg shell waste," Polymer
Degradation and Stability, vol. 132, pp. 78-86, 2016.

[26] M. Z. K. P. P. C. J. R. Anna Bryśkiewicz, "Modification of flexible polyurethane foams by


the addition of natural origin fillers," Polymer Degradation and Stability, vol. 132, pp. 32-
40, 2016.

[27] T. A. D. F. C. W. Suqin Tana, "Rigid polyurethane foams from a soybean oil-based Polyol,"
Polymer, vol. 52, no. 13, pp. 2840-2846, 2011.

[28] W. Z. A. C. Andrzej K Bledzki, "Natural-fibre-reinforced polyurethane microfoams,"


Composites Science and Technology, vol. 61, no. 16, pp. 2405-2411, 2001.

[29] M. S. M. T. K. Ayan Chakraborty, "Cellulose microfibrils: A novel method of preparation


using high shear refining and cryocrushing," Holzforschung, vol. 59, no. 1, pp. 102-107,
2005.

[30] I. J. Z. P. A. Guo, "Rigid polyurethane foams based on soybean oil," Journal of Applied
Polymer Science, vol. 77, no. 2, pp. 467-473, 2000.

[31] W. Z. A. C. Andrzej K.Bledzki, "Natural-fibre-reinforced polyurethane microfoams,"


Composites Science and Technology, vol. 61, no. 16, pp. 2405-2411, 2001.

[32] O.-J. K. B. C. C. J. W. C. &. J.-S. P. Nari Lee, "Characterization of castor


oil/polycaprolactone polyurethane biocomposites reinforced with hemp fibers," Fibers
and Polymers, vol. 10, pp. 154-160, 2009.

[33] R. W. X. S. Sun, Bio-Based Polymers and Composites, Academic Press, 2005.

[34] S. L. Bo Smeder, "Market oriented identification of important properties in developing


flax fibres for technical uses," Industrial Crops and Products, vol. 5, no. 2, pp. 149-162,
1996.

[35] S. M. S. K. A. E. S. Z. Y.A. El-Shekeil, "Development of a new kenaf Bast fibre reinforced


thermoplastic polyurethane composite," Bioresources, vol. 6, no. 4, pp. 4662-4672, 2011.

92
[36] S. Selke, Biodegradation and packaging. A literature review., Leatherhead, Surrey: PIRA
Information Services, 1990.

[37] T. M. G. P. B. J. E. Caroline McClory, "Thermosetting Polyurethane Multiwalled Carbon


Nanotube Composites," Journal of Applied Polymer Science, vol. 105, no. 3, pp. 1003-
1011, 5 August 2007.

[38] H. C. P. H. M. J. &. B. K. K. S. H. Kim, "Glass fiber reinforced rigid polyurethane foams,"


Journal of Materials Science, vol. 45, pp. 2675-2680, 2 February 2010.

[39] L. K. S, "Preparation and Tensile Strength Evaluation of Synthetic Fibers Sandwiched with
Foam Structures," International Research Journal of Engineering and Technology (IRJET),
vol. 5, no. 9, pp. 1647-1651, September 2018.

[40] C. B. P. a. N. R. Shau-Tarng Lee, Polymeric Foams: Science and Technology, Bocan Raton,
FL: CRS Press, 2007, p. 220.

[41] M. A. M. M. A. B. A. G. G. G. M. E. E. G. Bogoeva-Gaceva, "Natural fiber eco-composites,"


Polymer composites, vol. 28, no. 1, pp. 98-107, February 2007.

[42] C. Kau, "Damage processes in reinforced reaction injection molded polyurethanes,"


Journal of Reinforced Plastics and Composites, vol. 8, pp. 18-39, January 1989.

[43] N. K. a. S. B. Akriti Agrawal, "Derivatives And Applications Of Lignin : An Insight," The


Scitech Journal, vol. 1, no. 7, pp. 30-36, July 2014.

[44] G. T. B. M. J. B. R. C. F. C. M. F. D. B. H. D. R. A. D. P. G. M. K. P. L. A. K. N. J. N. S. T. J. T. G.
A. Arthur J. Ragauskas, "Lignin Valorization: Improving Lignin Processing in the
Biorefinery," Science, vol. 344, no. 6185, May 2014.

[45] M. J. S. A. A. H. Ľudmila Hodásová, "Lignin, potential products and their market value,"
Wood research, vol. 60, no. 6, pp. 973-986, December 2015.

[46] G. A. S. D.S.Bajwa, "A concise review of currernt lignin production, applications, products
and their environmental impact," Industrial Crops & Products, vol. 139, 1 November
2019.

93
[47] C. K. Harsh Yadav, "Natural polysaccharides: Structural features and properties,"
Polysaccharide Carriers for Drug Delivery, pp. 1-17, 21 June 2019.

[48] P. J. C. H. Bartnicki-Garcia S, "An electron microscope and electron diffraction study of


the effect of calcofluor and congo red on the biosynthesis of chitin in vitro.," Ach
Biochem Biophys, vol. 310, no. 1, pp. 6-15, April 1994.

[49] R. M. Younes I, "Chitin and Chitosan Preparation from Marine Sources. Structure,
Properties and Applications," Mar Drugs, vol. 13, no. 3, pp. 1133-1174, 2 March 2015.

[50] M. Rinaudo, "Main properties and current applications of some polysaccharides as


biomaterials.," Polymer International, vol. 57, no. 3, pp. 397-430, 9 October 2008.

[51] M. Rinaudo, "Physical properties of chitosan and derivatives in sol and gel states,"
Chitosan-Based Systems for Biopharmaceuticals: Delivery, Targeting and Polymer
Therapeutics, pp. 23-43, 17 February 2012.

[52] M. Rinaudo, Materials based on chitin and chitosan. In Bio-Based Plastics: Materials and
Applications, S. Kabasci, Ed., John Wiley & Sons, Ltd, 2014, pp. Chapter 2, 63-80.

[53] M. L. K. M. E. L. D. I. M. A. BoYuan, "Extraction, identification, and quantification of


antioxidant phenolics fromhazelnut (Corylus avellanaL.) shells," Food Chemistry, vol. 244,
pp. 7-15, 2018.

[54] B. A. Atila Çaǧlar, "Isothermal co-pyrolysis of hazelnut shell and ultra-high molecular
weight polyethylene: The effect of temperature and composition on the amount of
pyrolysis products," Journal of Analytical and Applied Pyrolysis, vol. 86, no. 2, pp. 304-
309, 2009.

[55] A. Demirbaş, "Properties of charcoal derived from hazelnut shell and the production of
briquettes using pyrolytic oil," Energy, vol. 24, no. 2, pp. 141-150, 1999.

[56] M. H. V. K. G. V. G. K. K. K. Y. Olivera S, "Potential applications of cellulose and chotosan


nanoparticles/ composites in wastewater treatment: A review," Carbohydrate Polymers,
vol. 153, pp. 600-618, 20 November 2016.

94
[57] D. Güllü and A. Demirbaş, "Biomass to methanol via pyrolysis process," Energy
Conversion and Management, vol. 42, no. 11, pp. 1349-1356, Jult 2001.

[58] A. Demirbas, "Furfural production from fruit shells by acid-catalyzed hydrolysis," Energy
Sources, Part A: Recovery, Utilization, and Environmental Effects , vol. 28, no. 2, pp. 157-
165, 15 August 2006.

[59] S. C. D. Uzuner, "Hydrolysis of hazelnut shells as a carbon source for bioprocessing


applications and fermentation," International Journal of Food Processing, vol. 10, no. 4,
pp. 799-808, 2014.

[60] M. K. E. M. L. S. M. Z. J. R. J. F. A. R. Aleksander Prociak, "Biobased polyurethane foams


modified with natural," Polimery, vol. 60, no. 09, pp. 592-599, 2015.

[61] G. A. D. M. I. A. A. N. M. M. A. Mosiewicki, "Polyurethane Foams Obtained from Castor


Oil-based Polyol and FIller with Wood Flour," Journal of COMPOSITE MATERIALS, vol. 43,
no. 25, pp. 3057-3072, 2009.

[62] F. V. D. P. Darshil U. Shah, "Silk cocoons as natural macro-balloonfillers in


novelpolyurethane-based syntactic foams," Polymer, vol. 56, pp. 93-101, 2015.

[63] M. K. L. L. S. A. B. M. K. K. B. J. R. Milena Zieleniewska, "Development and applicational


evaluation of the rigid polyurethane foam composites with egg shell waste," Polymer
Degradation and Stability, vol. 132, pp. 78-86, 2016.

[64] Y. L. F. Hsieh, "Water-Blown Flexible Polyurethane Foam Extended with Biomass


Materials," Journal of Applied Polymer Science, vol. 65, no. 4, pp. 695-703, 1997.

[65] W.-S. C. J.-N. L. Cha-Cheol Park, "Effects of hardness and thickness of polyurethane foam
midsoles on bending properties of the footwear," Fibers and Polymers, vol. 8, no. 2, pp.
192-197, 2007.

[66] M.-K. S. Soo-Jin Park, "Interface Science and Composites," Interface Science and
Interface, vol. 18, 2011.

[67] K. D. S. K. Akihiro Abe, Biopolymers - Lignin, Proteins, Bioactive, Nanocomposites,


Springer, 2010.

95
[68] F. H. Y. Lin, "Water-blown flexible polyurethane foam extended with biomass materials,"
Journal of Applied Polymer Science , vol. 65, no. 4, pp. 695-703, 1998.

[69] M. S. K. J. Y. L. C. J. Jin, "A study on viscoelasticity of polyurethane–organoclay


nanocomposites," Journal of Applied Polymer Science, vol. 99, no. 6, pp. 3677-3683,
2006.

[70] P. H. K. J. A. M. R. T. P. C. R. K. a. D. M. Bradley Finnigan, "Effect of the Average Soft-


Segment Length on theMorphology and Properties of Segmented
PolyurethaneNanocomposites," Journal of Applied Polymer Science, vol. 102, no. 1, pp.
128-139, 2005.

[71] M. Rinaudo, "Chitin and chitosan: Properties and applications," Progress in Polymer
Science, vol. 31, no. 7, pp. 603-632, 2006.

[72] W. E. Showalter, "Preparation of polyurethane foams containing an aluminum silicate


filler". United States Patent US3227666A, 1962.

[73] H. R. M. C. G. .̧ S. L. M. S. L. Q. F. R. L. L. G. G. S. Magnovaldo Carvalho Lopes, "High


performance polyurethane composites with isocyanate-functionalized carbon
nanotubes: Improvements in tear strength and scratch hardness," Journal of Applied
Polymer Science, vol. 134, no. 2, 2016.

[74] W. Z. V. K. Z. S. P. V. D. Ivan J. Javni, "Effect of Nano-and Micro-Silica Fillers on


Polyurethane Foam Properties," Journal of Cellular Plastics, vol. 38, 2002.

[75] J. H. H. James W. Hartings, "Fatigue Investigation of Urethane Seat Pads," Journal of


Cellular Plastics, vol. 14, no. 2, pp. 81-86, 1978.

[76] K. R. W. W. Erwin Baur, Chemical Resistance of Commodity Thermoplastics, 2016.

[77] B. K. a. I. M.A., "Natural Fiber as a substitute to Synthetic Fiber in Polymer Composites: A


Review," Research Journal of Engineering Sciences, vol. 2, no. 3, pp. 46-53, 2013.

[78] S. N. S. E. K. C. W. M. G. Harikrishnan, "Nanodispersions of carbon nanofibers for


polyurethane foaming," Polymer, vol. 51, no. 15, pp. 3349-3353, 2010.

96
[79] S. H. P. S. C. A. Darlan de Mello, "The effect of post-consumer PET particles on the
performance of flexible polyurethane foams," Polymer Testing, vol. 28, no. 7, pp. 702-
708, 2009.

[80] S. H. K. J. H. P. B. K. K. S. M. Kang, "Carbon nanotube reinforced shape memory


polyurethane foam," Polymer Bulletin, vol. 70, pp. 885-893, 2013.

[81] A. H. K. F. M. D. &. M. S. Łukasz Piszczyk, "Effect of ground tire rubber on structural,


mechanical and thermal properties of flexible polyurethane foams," Iranian Polymer
Journal, vol. 24, no. 1, pp. 75-84, 2015.

[82] P. W. A. K. K. S. a. C. S. Aasgeir Helland, "Reviewing the Environmental and Human


Health Knowledge Base," Environmental Health Perspectives, vol. 115, no. 8, pp. 1125-
1131, 2007.

[83] J. J. M. R. H. R. Lam C-W, "Pulmonary toxicity of single-wall carbon nanotubes in mice 7


and 90 days after intratracheal instillation.," Toxicol Science, vol. 77, no. 1, pp. 126-134,
2004.

[84] D. K. B. A. S. a. S. M. R. Arvidsson, "Prospective life cycle assessment of graphene


production," Environmental Science Technology, vol. 48, no. 8, pp. 4529-4536, 2014.

[85] B. R. B. a. L. J. L. V. Khanna, "Carbon nanofiber production: life cycle energy consumptoin


and environmental impact," Journal Industrial Ecology, vol. 12, no. 3, pp. 394-410, 2008.

[86] F. G. a. V. Piemonte, "Life Cycle Assessment ofPolylactic Acid and


PolyethyleneTerephthalate Bottles forDrinking Water," Department of Chemical
Engineering Materials and Environment, vol. 30, no. 3, pp. 459-468, 2010.

[87] N. C. K. J. Libo Yan, "Effect of triggering and polyurethane foam-filler on axial crushing of
natural flax/ epoxy composite tubes," Materials & Design, vol. 56, pp. 528-541, 2014.

[88] M. Z. K. P. P. C. J. R. Anna Bryskiewicz, "Modification of flexible polyurethane foams by


the addition of natural origin fillers," Polymer Degradation and Stability, vol. 132, pp. 32-
40, 2016.

97
[89] S. A. Ernie Suzana Ali, "Bionanocomposite hybrid polyurethane foam reinforced with
empty fruit bunch and nanoclay," Composites Part B: Engineering, vol. 43, no. 7, pp.
2813-2816, 2012.

[90] F. O. C. P. H. A. K. M. B. I.O. Bakare, "Mechanical and thermal properties of sisal fiber-


reinforced rubber seed oil-based polyurethane composites," Materials & Design, vol. 31,
no. 9, pp. 4274-4280, 2010.

[91] R. L.-P. I. .̃ a. M. Saioa Garbizu, "Surface Modification of Sisal Fibers: Effects on the
Mechanical and Thermal Properties of Their Epoxy Composites," Polymer Composites,
vol. 26, no. 2, pp. 121-127, 2005.

[92] M. A.-S. S. M. F.A. Al Sagheera, "Extraction and characterization of chitin and chitosan
from marine sources in Arabian Gulf," Carbohydrate Polymers, vol. 77, no. 2, pp. 410-
419, 2009.

[93] J. S. S. H. I. G. M. G. M. E. G. a. J.-F. R. William J. Orts, "Application of Cellulose


Microfibrils in Polymer," Journal of Polymers and the Environment, vol. 13, pp. 301-306,
2005.

[94] J. W. A. (. G. a. H. E. N. Muhammad A. S. Anwer, "Chitin nano-whiskers (CNWs) as a bio-


based biodegradable reinforcement for epoxy: evaluation of the impact of CNWs on the
morphological, fracture, mechanical, dynamic mechanical, and thermal characteristics of
DGEBA epoxy resin," Royal Society of Chemistry, vol. 9, no. 20, pp. 11063-11076, 2019.

[95] E. K. a. L. Y. Lim, "Implantable applications of chitin," Biomaterials, vol. 24, no. 13, pp.
2339-2349, 2003.

[96] A. N. a. V. O. M. Mincea, "Preparation, modification, and applications of chitin


nanowhiskers: a review," Rev. Adv. Materl. Sci, vol. 30, no. 3, pp. 225-242, 2012.

[97] D. C. Sibel Uzuner, "Hydrolysis of Hazelnut Shells as a Carbon Source for Bioprocessing
Applications and Fermentation," International Journal of Food Engineering, vol. 10, no. 4,
pp. 799-808, 2014.

[98] A. Y. S. NE., "Effects of pretreatment methods for hazelnut shell hydrolysate


fermentation with Pichia Stipitis to ethanol," Bioresour Technol, pp. 8664-70, 2010.

98
[99] D. A., "Utilization of lignin degradation products from hazelnut shell via supercritical fluid
extraction.," Energy Source, pp. 891-897, 2002.

[100] W. Stadelman, Eggs and egg products. In: Francis, F.J. (Ed.), Encyclopedia of Food Science
and Technology, Wiley, 1999, pp. 593-599.

[101] C. X. B. L. Ziku Wei, "Application of waste eggshell as low-cost solid catalyst for biodiesel
production," Bioresource Technology, vol. 100, no. 11, pp. 2883-2885, 2009.

[102] M. o. A. (. o. t. P. R. o. China, "Yearbook ofAgriculture Statistics in China," Beijing, 2006,


pp. 3-5.

[103] J. Y. H. H. C. L. K. L. a. C. C. W.T. Tsai, "Development and characterization of mesoporosity


in eggshell ground by planetary ball milling," Micropor. Mesopor. Mater., pp. 379-386,
2008.

[104] N. C. V. L. C. W. S. K. Achanai Buasri, "Application of Eggshell Wastes as a


Heterogeneous," Sustainable Energy, vol. 1, no. 2, pp. 7-13, 2013.

[105] A. Madhavi Gaonkar, "Application of Eggshell as Fertilizer and Calcium Supplement


Tablet," International Journal of Innovative Research in Science, Engineering and
Technology, vol. 4, no. 8, 2015.

[106] C. T. Coutinho F.M.B., "Performance of polypropylene-wood fiber composites," Polymer


Testing, vol. 21, no. 3, pp. 581-587, 1999.

[107] S. D. a. J. J.P., "Natural Fiber Polymer Composites: A Review," Advanced Polymer


Technology, vol. 18, no. 4, pp. 351-363, 1999.

[108] U. A. a. H. D. Bettini S.H.P., "Effect of Process Parameters and Composition on


Mechanical, Thermal and Morphological Properties of Polypropylene/Sawdust
Composites," Journal of Applied Polymer Science, vol. 108, no. 4, pp. 2233-2241, 2008.

[109] R. D. Lanzhu Zhang, "Measurement and identification of dynamic properties of flexible


polyurethane foam," Journal of Vibration and Control, vol. 28, no. 4, pp. 517-526, 2010.

[110] M. Ionescu, "Chemistry and Technology of Polyols for Polyurethanes," Rapra Technology
Limited, Shawbury, 2005.

99
[111] S. R. Department, "Market value forecast of polyurethane worldwide from 2016 to
2021," November 2016. [Online]. Available:
https://www.statista.com/statistics/720449/global-polyurethane-market-size-forecast/.

[112] P. Insight, "Polyurethane Production, Pricing and Market Demand," Plastics Insight, 2018.

[113] K. Ashida, "Polyurethane and Related Foams Chemistry and Technology," Taylor &
Francis Group: Boca Raton, FL, 2007.

[114] M. Z. M. R. K. F. C.L. Wei, "Tensile performance improvement of low nanoparticles filled-


polypropylene composites," Composites Science and Technology, vol. 62, no. 10-11, pp.
1327-1340, 2002.

[115] H. D. Rahul R.Maharsia, "Enhancing tensile strength and toughness in syntactic foams
through nanoclay reinforcement," Materials Science and Engineering: A, Vols. 454-455,
pp. 416-422, 2007.

[116] J. Z. L. D. H. W. J. L. W. H. a. Y. H. Fangzing Li, "Study of the Synthesis of High Elongation


Polyurethane," Eur. Polym. J., vol. 34, no. 1, pp. 59-66, 1998.

[117] O. K. N. L. Nattapon Kaisangsri, "Biodegradable foam tray from cassava starch blended
with natural fiber and chitosan," Industrial Crops and Products, vol. 37, no. 1, p. 542–
546, 2012.

[118] C. B. P. a. J. J. B. L. M. Matuana, "Cell morphology and property relationships of


microcellular foamed pvc/wood-fiber composites," Polymer Engineering & Science, vol.
38, no. 11, pp. 1862-1872, 1996.

[119] M. D. P. C. K. O. Y. Z. U. W. E. G. Jir-ShyrChen, "Study of the interlayer expansion


mechanism and thermal–mechanical properties of surface-initiated epoxy
nanocomposites," Polymer, vol. 43, no. 18, pp. 4895-4904, 2002.

[120] Z. Z. J. H. F. a. W. J. M. Zoran S. Petrovic, "Thermal Degradation of Segmented


Polyurethanes," Department of Polymer Science and Engineering, pp. 177-181, 1976.

100
[121] Y. G. M. S. Z. W. J. N. Ding, "Experimental and theoretical study of Young modulus in
micromachined polysilicon films," Tsinghua Science and Technology, vol. 7, no. 3, pp.
270-275, 2002.

[122] A. G. A. G. S. T. a. P. C. E. Prasoon Joshi, "Improvement of the elastic modulus of


micromachined structures using carbon nanotubes," Proceedings of SPIE - The
International Society for Optical Engineering, vol. 875, pp. 27-32, 2006.

[123] P. E. H.-R. S. K. L. a. Y.-L. M. P. E. Wei Zheng, "Impact of Nanotechnology on Future Civil


Engineering Practice and Its Reflection in Current Civil Engineering Education," Journal of
Professional Issues in Engineering Education and Practice, vol. 137, no. 3, 2011.

[124] Y. L. F. Hsieh, "Water-blown flexible polyurethane foam extended with biomass


materials," Journal of Applied Polymer Science , 1996.

[125] W. Z. V. K. a. Z. S. P. I. Javni, "Effect of Nano- and Micro-Silica Fillers on Polyurethane


Foam Properties," Journal of Cellular Plastics, Arlington, 2002.

[126] M. A. Koshute, "Influence of PU Foam Thickness, Density, and Hardness on Seat Cushion
Durability," SAE International , 1993.

[127] T.-T. L. Y.-C. C. C.-H. H. a. C.-W. L. Jia-Horng Lin, "Resilience Property of Spacer Fabric
Compounded PU Foam Composites," Advanced Materials Research, vol. 910, pp. 218-
221, 2014.

[128] L.-m. S. J. H. a. J. S. Yuding Zhu, "Study on Resilience of Mattress Foam and Body Pressure
Distribution Characteristic of Mattress," Advanced Materials Research, Vols. 415-417, pp.
281-284, 2012.

[129] K. B. C. a. F. J. Preston, "A Novel Approach to High Resilience Foam," 1979.

[130] A. N. G. a. H. J. Kim, "Tear Strength of Stretched Rubber," Rubber Chemistry and


Technology, vol. 51, no. 1, pp. 35-44, 1978.

[131] W. F. Busse, "Tear Resistance and Structure of Rubber," Industrial Engineering Chemistry,
vol. 26, no. 11, pp. 1194-1199, 1934.

101
[132] J. H. Fielding, "Impact Resilience in Testing Channel Black," Industrial Engineering
Chemistry, vol. 29, no. 8, pp. 880-885, 1943.

[133] R. C. a. P. Thirion, "Proc. Internatl. Rubber Conf.," American Chemical Society,


Wachington D.C., 1959.

[134] D. T. G. R. H. D. Gibala, "Cure and Mechanical Behaviour of Rubber Compounds


Containing Ground Vulcanizates: Part III. Tensile and Tear Strength," Polymer Science
Department, vol. 72, no. 2, pp. 357-360, 1999.

[135] A. N. G. a. J. R. W. P. Dreyfuss, "Tear Strength and Tensile Strength of Model Filled


Elastomers," Institute of Polymer Science, vol. 54, no. 5, pp. 1003-1010, 1981.

[136] M. Y. J. C. Z. J. Y. M. Jun Zhang, "Synthesis and properties of polyurethane elastomers


based on renewable castor oil polyols," Journal of Applied Polymer Science, vol. 136, no.
14, 2018.

[137] W. W. Buc Slay, "Stress relaxation of elastomer compounds," Sealing Technology, vol.
2011, no. 2, pp. 9-12, 2011.

[138] R. M. Yu Hua Yi, "Study of the Main Influence Factors on Compression Set of PTMG
Based Polyurethane Elastomers," Advanced Materials Research, pp. 67-70, 2013.

[139] H. J. J. a. H. H. Bertram, "The Compression Set Behaviour of Nitrile Rubber," Rubber


Chemistry and Technology, vol. 46, no. 1, pp. 305-330, 1973.

[140] C. D. M. S. C. Slater, "Compression set of thermoplastic polyurethane under different


thermal-mechanical-moisture conditions," Polymer Degradation and Stability, vol. 96,
no. 12, pp. 2139-2144, 2011.

[141] Z. Z. J. H. F. a. W. J. M. Zoran S. Petrovic, "Thermal Degradation of Segmented


Polyurethanes," Journal of Polymer Science, vol. 51, no. 6, pp. 1087-1095, 1994.

[142] E. A. S. S. C. N. Y. P. M. G. O. C. Graziella Trovati, "Characterization of Polyurethane


Resins by FTIR, TGA,and XRD," Journal of Applied Polymer Science, vol. 115, no. 1, pp.
263-268, 2010.

102
[143] S. C. Neto, "Physical chemistry characterization of a polyur-ethane derived from castor
oil used for bone implants," 1997.

[144] L. M. M. C. B. P. J. J. Balatinecz, "Cell Morphology and Property Relationships of


Microcellular Foamed PVC/Wood-Fiber Composites," Polymer Engineering and Science,
vol. 38, no. 11, pp. 1862-1872, 2004.

[145] F. Z. T. S. C. a. D. N. U. L. Hazmi Hilmi, "Mechanical properties of palm oil based bio-


polyurethane foam of free rise and various densities," Advanced Materials for
Sustainability and Growth, vol. 1901, no. 1, 2017.

[146] G. B. a. R. M. M. Avalle, "Characterization of polymeric structural foams under


compressive impact loading by means of energy-absorption diagram," International
Journal of Impact Engineering, vol. 25, no. 5, pp. 455-472, 2001.

[147] S. Oprea, "Synthesis and Properties of Polyurethane Elastomers with Castor Oil as
Crosslinker," Journal of the American Oil Chemists Society, vol. 87, no. 3, pp. 313-320,
2009.

[148] M. Ionescu, Chemistry and Technology of Polyols for Polyurethanes, Smithers Rapra
Technology, 2005.

[149] C. d. H. R. Schiffauer, "Flexible Polyurethane Slabstock Foam: The Influence of


Formulation, Climatic Conditions and Storage Conditions on Foam Properties," Journal of
Cellular Plastics, vol. 19, no. 1, pp. 61-64, 1983.

[150] "Polyurethane Global Market Size Forecast 2021.," 1 November 2016. [Online].
Available: https://www.statista.com/statistics/720449/global-polyurethane-market-size-
forecast/. [Accessed 2016].

[151] K. Ashida, Polyurethane and Related Foams Chemistry and Technology, Taylor & Francis
Group, 2006.

[152] A. P. S. M. K. Z. Maria Kurańska, "The influence of blowing agents type on foaming


process and properties of rigid polyurethane foams," Polimery, Cracow, Poland, 2017.

103
[153] S.-T. L. N.S. Ramesh, "Polymeric Foams: Mechanisms and Materials," CRC Press, New
York, 2004.

[154] D. H. C. A. D. U. D. L. Schimidt, "The Effect of Surfactant Properties on a Rigid Foam


System," Journal of Cellular Plastics, 1984.

[155] K. X. Narine SS, "Physical properties of polyurethane plastic sheets produced from
polyols from canola oil biomacromolecules," Biomacromolecules, vol. 8, no. 7, pp. 2203-
2209, 2007.

[156] M. H. J. Peter Kjeldsen, "Release of CFC-11 from disposal of Polyurethane Foam Waste,"
Environmental Sci. Technol., vol. 35, no. 14, pp. 3055-3063, 2001.

[157] P. K. J. V. W. V. C. Rittmeyer, "Decomposition of organohalogen compounds in municipal


solid waste incineration plants. Part II: Co-combustion of CFC containing polyurethane
foams," Chemosphere, vol. 28, no. 8, pp. 1455-1465, 1994.

[158] E. L.-G. M. A. R.-P. M. A. Heura Ventura, "Effect of chain extender and water-quenching
on the properties of poly(3-hydroxybutyrate-co-4-hydroxybutyrate) foams for its
production by extrusion foaming," European Polymer Journal, vol. 85, pp. 14-25, 2016.

[159] M. K. S. K. H. P. H. J. B. K. S.M. Kang, "Polyurethane foam/silica chemical hybrids for


shape memory effects," J. Mater. Res, vol. 27, no. 22, pp. 2837-2843, 2012.

[160] B. S. C. S. F. R. S. A. F. a. A. B.-T. Nuno V Gama, "Effect of unrefined crude glycerol


composition on the properties of polyurethane foams," Journal of Cellular Plastics, vol.
54, no. 3, pp. 633-649, 2018.

[161] S. U. Cekmecelioglu, "Hydrolysis of hazelnut shells as a carbon source for bioprocessing


applications and fermentation," International Journal of Food Engineering, pp. 799-808,
2014.

[162] W. Y. N. M. J. S. B. O. A. H. Mohamad Nurul Azman, "Influence of Hydrophobicity of


Acetylated Nanocellulose on the Mechanical Performance of Nitrile Butadiene Rubber
(NBR) Composites," Fibers and Polymers, vol. 19, no. 2, p. 383, 2018.

104
[163] P. B.-G. S, "An electron microscope and electron diffraction study of the effect of
calcoflour and congo red on the biosynthesis of chitin in vitro," Ach Biochem Biophys,
vol. 310, pp. 6-15, 1995.

[164] O. S. I. Fayomi, Artist, Chemical Structure of Eggshell Powder. [Art]. Covenant University
Ota Ogun State, 2018.

105

You might also like