You are on page 1of 116

THE USE OF AGRICULTURAL WASTE AS SUGAR

SUBSTITUTE IN THE SYNTHESIS OF HYDROTALCITE


CATALYST FOR THE PRODUCTION OF BIODIESEL

MOHAMED AFIQ BIN MOHAMED MOFFIT

THESIS SUBMITTED IN FULFILLMENT OF THE REQUIREMENTS FOR


THE DEGREE OF MASTER OF ENGINEERING TECHNOLOGY
(CHEMICAL ENGINEERING)

UNIVERSITI KUALA LUMPUR MALAYSIAN INSTITUTE OF CHEMICAL


AND BIOENGINEERING TECHNOLOGY

2020

i
APPROVAL PAGE

This thesis was submitted to the Senate of Universiti Kuala Lumpur and has
been accepted as fulfilment of the requirement for the degree of Master of
Engineering Technology (Chemical Engineering). The members of the
Supervisory committee were as follow:

___________________________
Name of Main Supervisor
Dr. Mohd Razealy Anuar
UniKL MICET

___________________________
Name of Co – Supervisor 1
Assoc. Prof. Dr. Kelly Yong Tau Len
UniKL MICET

ii
DECLARATION

I declare that the work in this thesis was carried out in accordance with the
regulations of Universiti Kuala Lumpur. It is original and is the result of my own
work, unless otherwise indicated or acknowledge as reference work. This topic
has not been submitted to any academic institution or non-academic institution
for any other degree or qualification.

In an event that this thesis is to be found to violate the conditions above, I


voluntarily waive the right of conferment of my degree and be subjected to the
disciplinary rules and regulations of Universiti Kuala Lumpur

Name : Mohamed Afiq Bin Mohamed Moffit

Student ID : 55360117005

Programme : Master of Engineering Technology (Chemical


Engineering)

Thesis Title : The Use of Agricultural Waste as Sugar


Substitute in The Synthesis of Hydrotalcite
Catalyst for The Production of Biodiesel
Place : Universiti Kuala Lumpur Malaysian Institute of
Chemical and Bioengineering Technology

Signature of candidate: ……………………………………………..

Date : ……………………………………………..

iii
STATEMENT OF COPYRIGHT

The copyright of this thesis rests with the author. No quotation from its should
be published without his prior written consent and information derived from it
should be acknowledged.

By signing this form, I acknowledge that I have read and understand the UniKL
Intellectual Property Right and Commercialization policy

Affirmed by Mohamed Afiq Bin Mohamed Moffit

………………………
Signature

………………………
Date

iv
ACKNOWLEDGEMENT

This thesis exists with generous help and support of many individuals. I would
like to extend my sincere thanks to all of them.

Foremost, I want to offer this endeavour to Almighty GOD for without the
wisdom he bestowed upon me, the strength, peace of my mind and good
health, this research would not have been possible.

I dedicate my thesis to my family especially special acknowledgement to my


parents, Mohamed Moffit Othman and Normah Sayuti for giving me an
opportunity and encouragement to do the thesis. To my brother and sister,
Faisal and Sabrina for keeping me in cherish and advising me during
hardships.

Special thanks to my supervisors, Dr. Mohd Razealy Anuar and Assoc. Prof.
Dr. Kelly Yong Tau Len for giving me lots of guidance, advices, critics and
valuable comments for the thesis. To the technicians especially Mrs. Nur
Ahzlinda Ahmad and Mr. Mohd Sukri Rahmat that occasionally gave me
advices and supervision during experiments.

And I would like to express my gratitude to my best friends and partner, Zuliza
Mokhtar, Aminuddin Mohamed Alwee and Mohd Firdaus Zubbir for keeping
my courage and morale at its peak. My fellow master students should also be
recognized for their supports. I extend my sincere appreciation to all my
colleagues and others who have provide support, courage, and assistance.
Their tips and views are useful indeed.

v
TABLE OF CONTENT

Approval Page II
Declaration III
Statement of Copyright IV
Acknowledgement V
List of Tables IX
List of Figures X
List of Abbreviations XII
List of Appendices XIII
Abstract XIV
Abstrak XV

CHAPTER ONE: INTRODUCTION 1


1.1 Biodiesel Demand 1
1.2 Biodiesel Production Prospect 3
1.3 Agricultural Waste to Value Added Product 5
1.4 Problem Statement 6
1.5 Objectives 7
1.6 Scope of Study 8

CHAPTER TWO: LITERATURE REVIEW 10


2.1 Biodiesel Feedstock 10
2.2 Biodiesel from Waste Cooking Oil 12
2.3 Biodiesel Production 13
2.4 Catalyst for Transesterification Reaction 14
2.4.1 Homogeneous Catalyst 14
2.4.2 Heterogeneous Catalyst 17
2.5 Hydrotalcite 21
2.5.1 Hydrotalcite Synthesis Route 22
2.5.2 Combustion Method 24

vi
2.6 Agricultural Wastes as Fuel for Hydrotalcite Synthesis 25
2.7 Reaction Kinetics 28
2.8 Concluding Remarks 33

CHAPTER THREE: METHODOLOGY 34


3.1 Materials and Equipment 34
3.2 Overall Experimental Flowchart 36
3.3 Preparation of Hydrotalcite 37
3.4 Fuel Calorific Value Determination 38
3.5 Average Molecular Weight Value Determination of WCO 39
3.5.1 Saponification Value Determination 39
3.5.2 Acid Value Determination 40
3.6 Transesterification Reaction of Biodiesel 40
3.7 Kinetic Studies of Transesterification Reaction 43
3.8 Catalyst Analysis 44
3.8.1 XRD Analysis of Hydrotalcite 44
3.8.2 BET Analysis of Hydrotalcite 44
3.8.3 FTIR Analysis of Hydrotalcite 44
3.8.4 TGA Analysis of Hydrotalcite 45
3.9 FAME Analysis 45
3.9.1 Gas Chromatography Analysis of FAME 45
3.9.2 FTIR Analysis of FAME 45

CHAPTER FOUR: RESULT & DISCUSSION 47


4.1 Catalyst Characterization 47
4.1.1 XRD Analysis of HT 47
4.1.2 BET Analysis of HT 50
4.1.3 FTIR Analysis of HT 52
4.1.4 TGA Analysis of HT 54
4.2 Catalyst Performance 56
4.2.1 Effect of Fuel Type 57
4.2.2 Effect of Calcination Temperature 58
4.2.3 Reusability of HT 60

vii
4.3 FTIR Studies on Biodiesel 62
4.4 Kinetic Studies of Biodiesel Synthesis 64
4.4.1 Effect of Reaction Temperature and Time on 64
Transesterification Reaction
4.4.2 Transesterification Reaction Kinetics 66

CHAPTER FIVE: CONCLUSION 69


5.1 Conclusion 69
5.2 Recommendation 70

REFERENCES 71
APPENDICES 91

viii
LIST OF TABLES

2.1 List of homogeneous acid and base catalyst used in 15


transesterification reaction
2.2 List of heterogeneous catalyst used in transesterification 19
reaction
2.3 List of potential waste material used as fuel for synthesization of 27
hydrotalcite in comparison with current fuel
2.4 Activation energy values for various catalysts 29
3.1 List of all equipment used in this research 35
3.2 List of all HT samples in this research 38
4.1 Calorific value of tested fuel 49
4.2 Surface characteristics of HT 50
4.3 List of activation energy values for synthesized HTs 66

ix
LIST OF FIGURES

1.1 Demand of biodiesel in Malaysia 2


2.1 General chemical equation for transesterification reaction of 13
triglyceride
2.2 Undesirable side reaction of transesterification; saponification 16
2.3 Typical structure of Hydrotalcite 21
3.1 Overall experimental flowchart 36
3.2 Transesterification reaction of biodiesel flowchart 41
3.3 Experimental setup for transesterification reaction 42
3.4 Transesterification reaction of biodiesel process steps for 43
kinetic studies
4.1 XRD patterns of (a) HT-SS 650, (b) HT-RS 650, (c) HT-RH 47
650, (d) HT-CS 550, (e) HT-CS 650, (f) HT-CS 750 and (g)
HT-CS 850
4.2 FTIR spectra of (a) HT-SS 650, (b) HT-RS 650, (c) HT-RH 52
650 and (d) HT-CS 650
4.3 (a) TGA profile of HT-SS 650 54
4.3 (b) TGA profile of HT-RS 650 54
4.3 (c) TGA profile of HT-RH 650 55
4.3 (d) TGA profile of HT-CS 650 55
4.4 Catalytic activity of HT with different type of fuel 57
4.5 Catalytic activity of HT- CS with different calcination 59
temperature
4.6 Reusability of catalytic activity of catalyst at 650°C for 5-hour 61
reaction
4.7 FTIR Spectra of (a) WCO (b) Biodiesel produced using HT- 63
SS 650, (c) HT-CS 650, (d) HT-RH 650 and (e) HT-RS 650
4.8 (a) Catalytic activity of HT-SS 650 under 5-hour reaction time, 15:1 64
methanol to oil molar ratio, 8% catalyst loading with various
temperature reaction.
4.8 (b) Catalytic activity of HT-CS 650 under 5-hour reaction time, 64
15:1 methanol to oil molar ratio, 8% catalyst loading with
various temperature reaction.

x
4.9 Arrhenius plot ln k versus 1/T for transesterification reaction 66
for various HT catalyst (HT-SS 650, HT-RH 650 and HT-CS
650)

xi
LIST OF ABBREVIATION

A Frequency factor T temperature


BET Brunauer, Emmet, Teller t time
CS Coconut Shell TG Triglyceride
°C Degree Celsius TGA Thermogravimetric Analysis
Ea Activation Energy WCO Waste Cooking Oil
FAME Fatty Acid Methyl Ester wt% Weight percentage
FFA Free Fatty Acid XRD X-Ray Diffraction
FTIR Fourier Transform Infra-Red
G Glycerol
GC Gas Chromatography
g gram
hr hour
HT Hydrotalcite
IUPAC International Union of Pure and Applied Chemistry
k constant
kg kilogram
LDH Layered Double Hydroxide
ln Natural logarithm
m meter
MJ Megajoule
mg milligram
R Gas constant
-ra Rate of disappearance
RH Rice Husk
ROH Alcohol
RS Rice Straw
SS Saccharose

xii
LIST OF APPENDICES

Appendix A Example of Bomb Calorimeter Data 92


Appendix B Example of Bomb Calorimeter Calculation 93
Appendix C Example of Acid Value and Saponification Value 94
Appendix D Example of Average Molecular Weight Oil Determination 95
of WCO
Appendix E Example of HT Preparation 95
Appendix F Example of GC Data 96
Appendix G Example of FAME Yield Calculation 97
Appendix H Example of Kinetic Study Calculation 98
Appendix I -ln(1-FAME) versus t graph 99
Appendix J Activation Energy Calculation 100

xiii
ABSTRACT

Hydrotalcite was synthesized via combustion method using coconut shell, rice husk,
rice straw and saccharose as fuel to produce biodiesel from waste cooking oil was
studied. The performance of each hydrotalcite was investigated with respect to the
biodiesel yield. Two factors were examined on the catalyst which are the effect of fuel
type and the effect of calcination temperature. The data from synthesis of hydrotalcite
via combustion method using saccharose as fuel was used as reference. The
hydrotalcite was calcined at various temperatures, i.e. 550 °C, 650 °C and 750 °C
respectively. As per XRD result, it was found that the synthesized hydrotalcite was
able to maintain its layered double hydroxide structure up to 650 °C calcination
temperature. Increase of further calcination temperature can result of structure
collapse, hence it will not become hydrotalcite. In BET analysis, it was found
hydrotalcite synthesized using coconut shell (HT-CS 650) has the highest surface
area development as comparable to reference hydrotalcite (HT-SS 650) which are
115.558 m2/g and 28.326 m2/g respectively. The high surface area of HTCS-650 was
enhanced with the findings on FTIR and TGA analysis where it depicts to have highest
content of carbonate and hydroxide. For the performance of hydrotalcite towards
catalytic activity, it was found that all hydrotalcite calcined at 650 °C has the highest
performance towards biodiesel yield. HTCS-650 has the best performance in the
catalytic activity with 92.86% which is higher than hydrotalcite using saccharose (HT-
SS 650) as fuel (73.88%). The HT-CS 650 can be reused up to 3 cycles with minimum
reduction of 9.09% biodiesel yield. Various transesterification reaction temperature
i.e. 55 °C, 60 °C, 65 °C and 70 °C respectively were conducted for kinetic study. It
was found that the best reaction temperature for transesterification for all hydrotalcites
at 65 °C reaction temperature. The reaction followed pseudo-first order kinetics. HT-
CS 650 has the lowest activation energy (Ea) which is 246.68 kJ/mol. For reference
HT-SS 650, the activation energy is 295.64 kJ/mol. The study shows that the
substitution towards agricultural waste from saccharose improve the structure of
hydrotalcite. It also improves the performance of hydrotalcite towards production of
biodiesel. With the use of agricultural waste for synthesis of hydrotalcite could improve
the performance in production of biodiesel, it can also reduce the agricultural waste
which as a result have the potential to save the environment.

xiv
ABSTRAK

Hidrotalsit telah dihasilkan melalui kaedah pembakaran dengan menggunakan


tempurung kelapa, sekam padi, jerami padi dan gula sebagai bahan bakar bertujuan
untuk menghasilkan biodiesel dari minyak masak sisa telah dipelajari. Prestasi setiap
hidrotalsit telah disiasat berkaitan dengan penghasilan biodiesel. Dua faktor penting
yang mempengaruhi penghasilan pemangkin tersebut iaitu jenis bahan bakar dan
kesan suhu kalsin pemangkin telah diperiksa. Data hidrotalsit yang menggunakan
gula sebagai bahan api telah digunakan sebagai rujukan. Dalam kajian ini, hidrotalsit
telah dikalsin pada pelbagai suhu iaitu 550 °C, 650 °C dan 750 °C. Berdasarkan kajian
XRD, ia telah didapati bahawa struktur hidrotalsit yang telah dihasilkan hanya mampu
bertahan struktur lapisan dwi-hidroksida sehingga pada suhu 650 °C waktu dikalsin.
Jika suhu kalsin ditinggkatkan lebih tinggi, ia boleh menyebabkan struktur hidrotalsit
rapuh, dan gagal menjadi hidrotalsit. Dalam kajian BET, ia telah didapati bahawa
hidrotalsit yang dihasilkan menggunakan tempurung kelapa (HT-CS 650) mempunyai
kawasan permukaan yang besar jika dibandingkan dengan hidrotalsit rujukan (HT-SS
650) iaitu 115.558 m2/g dan 28.326 m2/g. Ia telah diperkukuhkan lagi dengan hasil
kajian FTIR dan TGA dimana ia telah menunjukkan bahawa hidrotalsit tersebut
mempunyai kandungan karbonat dan hidroksida yang tinggi. Dari segi keberkesanan
hidrotalsit dalam aktiviti katalitik, ia telah didapati bahawa kesemua hidrotalsit yang
telah dikalsin pada suhu 650 °C mempunyai tahap pencapaian yang tinggi terhadap
penghasilan biodiesel. HT-CS 650 mempunyai tahap pencapaian yang tertinggi
dalam aktiviti katalitik dengan dapat menghasilkan 92.86% biodiesel berbanding
dengan HT-SS 650 (73.88%). HT-CS 650 juga boleh dipakai semula sehingga 3 kali
penggunaan dengan pengurangan hasil biodiesel sebanyak 9.09%. Pelbagai suhu
reaksi iaitu 55 °C, 60 °C, 65 °C dan 70 °C telah dikaji untuk bertujuan mengkaji kinetik
reaksi transesterifikasi. Kesemua hidrotalsit dapat menghasilkan biodiesel dengan
peratus tinggi dengan menggunakan suhu reaksi 65 °C. Reaksi transesterifikasi ini
juga mematuhi urutan pseudo-pertama kinetik. HT-CS 650 mempunyai tenaga
pengaktifan (Ea) yang paling rendah iaitu 246.68 kJ/mol. Manakala HT-SS 650
didapati mempunyai tenaga pengaktifan 295.64 kJ/mol. Kajian ini telah menunjukkan
bahawa dengan menggunakan sisa pertanian dapat menambah baik kepada struktur
hidrotalsit. Ia juga dapat menambah baik tahap pencapaian hidrotalsit dalam
penghasilan biodiesel. Dengan menggunakan sisa pertanian dalam penghasilan
hidrotalsit dapat menambah baik aktiviti katalitik, ia juga mampu mengurangkan sisa
pertanian, malah hasilnya berpotensi menyelamatkan alam sekitar.

xv
CHAPTER 1

INTRODUCTION

1.1 Biodiesel Demand

Biodiesel has become popular in one of alternative renewable energy. The


popularity of biodiesel is due to its chemical structure which quite resembles to the
alternative of distillate diesel which having carbon chain ranging from 14 to 16 (Wang
et al, 2014). In addition, the fuel is considered to be green alternative fuel and clean
energy resource due to its low emission sulphur, carbon monoxide and nitrogen
dioxide (Jha & Das, 2017; Saxena, 2019). Therefore, some countries such as USA,
Brazil, Europe, China, Malaysia and others have initiated various policies and
programs to support biodiesel production as an ideal alternative energy especially in
transportation sector (Da Silva César et al, 2019; Geng et al, 2019; Silveira et al,
2019; Zailani et al, 2019). By achieving the biodiesel policies and programs, it is
expected that the environmental impact such as acid rain, increase of CO2 emission,
increase of greenhouse pollution, air pollution and climate change can be reduced as
much as 78% less CO2 (Tyson, 2001; Zailani et al, 2019).

Large demand of biodiesel has been created with the initiation from the
governments, fossil fuel energy crisis, and environmental awareness from community
(Kataria et al, 2019; Zailani et al, 2019). Zailani et al (2019) has outlined four factors
which leads to the demand of biodiesel which are (i) government support, (ii) customer
demands, (iii) environmental commercial benefits and (iv) competitive pressure.
However, these four factors are lacking in Malaysia. Unlike in EU, the main driving
force of biodiesel market to be successful is due to initiative and tax exemption from
their government the where 0 tax rate were imposed to either pure biodiesel or
biodiesel blend since 1990 up till 2008 (Gärtner & Reindhart, 2005, Sinico, 2008).
With that generous incentive from the German government, the production of
biodiesel in Germany has led in a rapid expansion in 18 years without any tax imposed
to the biofuel which leads to one of most three important biofuels followed by ethanol

1
and pure vegetable oil (Rauch & Thöne, 2012). In current situation in Germany, the
German’s national car consist of diesel-powered engine to allow more options for the
user to use either diesel or biodiesel. Therefore, the Malaysian government, policy
makers and managers will have to take a stand with an interest in the use of biodiesel.
With mentioned implications, the biodiesel will slowly gain attention in Malaysia and
growing demands in the future. Therefore, in future, it is hoping that there will be diesel
engine vehicle for Malaysian national car so that biodiesel can be applicable for
everyone as now most of Malaysian national cars are petrol engine vehicle.

In 2006, the government of Malaysia released the national biofuel policy with
the idea to reduce the dependency of fossil fuel by planning to introduce the use of
sustainable energy such as biodiesel B5, B7, B10 and B20 as part of National Biofuel
Policy (Wahab, 2019). In 2008, The government of Malaysia began to involve in the
implementation of biodiesel B5 to the transportation sector. With the implementation,
the demand of biodiesel from Malaysian consumer began to grow. Figure 1.1 shows
the demand of biodiesel from 2010 to 2018.

1400
1245
1200
Biodiesel (Million Liters)

1000 917

800 730
660
611
600 507

400 271
204
200 130

0
2010 2011 2012 2013 2014 2015 2016 2017 2018
Years
Figure 1.1: Demand of biodiesel in Malaysia (Wahab, 2019)

From the figure, it can be seen that the demand of biodiesel is increasing
annually. It shows that the biodiesel will have great prospect in line with the initiative
of the Malaysian government to reduce the dependency of fossil fuel. It also increases
the industrial economy of Malaysia where 750 million liters of biodiesel has been

2
exported throughout other countries such as Europe and China (Wahab, 2019). From
the literature reviewed and Figure above, it is predicted that the demand of biodiesel
will be positive and will sustain in few years in future.

1.2 Biodiesel Production Prospect

Since the development of diesel engine in the mid-nineties, neat or pure


vegetable oil has a potential to be used directly by diesel engine without modification
(Fontaras et al 2011). However, when it is closely analyzed, neat vegetable oil has
high viscosity due to its triglyceride chain which leads to lower performance compared
to fossil diesel when in used in the engine (Dabi & Saha, 2019). In addition, the use
of neat or pure vegetable oil directly into the diesel engine could cause engine wear.
Therefore, the neat or pure vegetable oil requires a modification to lowering down the
viscosity property and the heating value of the vegetable oil which the process is
called transesterification reaction (Sahoo et al, 2019). After the modification via
transesterification, the oil, which is biodiesel has gain attractive as an alternate fuel
due to its compatibility, renewability, higher combustion efficiency and others
(Mardhiah et al, 2017)

Biodiesel was traditionally produced via transesterification with the use of


sodium hydroxide or potassium hydroxide as homogeneous catalyst is one of the
most favorable process due to high yield of biodiesel. However, as the biodiesel
production process becomes mature and community starts to concern about
environmental issues, the biodiesel production process needs become more
environmentally friendly in line with the promotion of environmental-friendly fuel,
biodiesel. Therefore, the use of sodium hydroxide or potassium hydroxide could cause
negative environmental issues (Onoji et al, 2017; Reyero et al, 2013). Hence,
heterogeneous catalyst was developed in order to overcome the homogeneous
catalyst issues (Salimi & Housseini, 2019).

Palm oil has been used as a source of renewable feedstock to produce


biodiesel in Malaysia as Malaysia is one of biggest palm oil producer in the world.
Palm oil also is the most suitable and sustainable due to the feedstock is readily
available where the production of palm oil is continuous throughout the year (Zahan

3
& Kano, 2018). However, the use of palm oil as a feedstock to produce biodiesel has
been a controversial since palm oil is also one of food source in Malaysia which could
leads to food versus fuel agenda (Lam et al, 2009; Ramli et al, 2017). In addition, the
use of palm oil in the production of biodiesel could leads to rapid expansion oil palm
plantation (Johari et al, 2015). Consequently, the use of palm oil for production of
biodiesel has a negative impact in a long term to the environment predominantly
deforestation, loss of biodiversity, air pollution and water pollution due to no proper
waste management in the palm oil mill (Lam et al, 2009). In terms of economy viability,
the increase price of palm oil also has resulted in the delay of the full implementation
of biodiesel B5 in covering the entire Malaysia which took almost 6 years until 2014
(Wahab, 2019). Therefore, the use of waste cooking oil is considered to be the
possible solution in overcoming the food issues in the production of biodiesel as it will
not inflict a food vs fuel debate as no food source were used in the production of
biodiesel (Ramli et al, 2017). The use of waste cooking oil in the production of
biodiesel also offers the initiative of converting waste into useful energy. Furthermore,
the use of the use of waste cooking for production of biodiesel will further solve the
environmental problems that arises from the oil improper disposal (Hanisah et al,
2013; Kabir et al, 2014; Ramli & Farooq, 2015; Ramli et al, 2017).

A promising technology which could overcome the problems associated with


homogeneous catalyst in the production of biodiesel is the use of heterogeneous
catalyst (Mardhiah et al, 2017). Transesterification reaction is the most common
method to be used in production of biodiesel. The reaction involves between cooking
oil or vegetable oil with alcohol and the presence of catalyst to produce biodiesel and
glycerol. The use of homogeneous catalyst in previous technology such as potassium
hydroxide or sodium hydroxide is due to cost effective and readily available
(Thangaraj et al, 2019). However, the use of heterogeneous catalyst in production of
biodiesel via transesterification reaction provide environmentally advantageous in
which the production of biodiesel is simplified where no water washing and purification
of products required (Marinkovi et al, 2016). With the growing demand of biodiesel,
sustainable feedstock from waste and the promising technology of production of
biodiesel mentioned above, the prospect of the production of biodiesel is expected to
be bright.

4
1.3 Agricultural Waste to Value Added Product

In Malaysia, agricultural waste is one of wastes that is not fully utilized such
as rice husk, rice straw, coconut shell, banana stem, palm oil shell and others (Khalaf
et al, 2019). Researchers have been reported that the agricultural waste was left after
harvest and disintegrated in open burning (Praveena et al, 2019; Rosmiza et al, 2014).
The open burning from agricultural waste could result in the increase of particulate
matter pollution, formation of heavy haze, environmental health hazards and
ecological damage to the affected areas (Praveena et al, 2019). Therefore, the use of
agricultural waste in the production of biodiesel should be initiated as it can promote
waste-to-energy.

Coconut is one of most major crops to be grown in Malaysia. It is the fourth


important crop to be planted in Malaysia after oil palm, rubber and paddy (DOSM,
2018). According to Food and Agriculture Organization of United Nation (2020),
Malaysia has produced 519,153 tonnes in 2018. According to Banzon (1980) each
1.2 kg of coconut could produce 0.18 kg coconut shell. Therefore, with 519,153 tonnes
of coconut generated in 2018, it eventually produces a preponderance amount of
agricultural waste approximate of 93 thousand tonnes of coconut shell were
generated in that year. For a long time in Malaysia, the method of disposal for coconut
shell has caused negative environmental issue such as open burning or dumping into
river system which could cause air and water pollution (Jaafar et al, 2019). Coconut
shell contains several types of carbohydrates. The coconut shell contains 19.8%
cellulose, 68.7% holocellulose and 30.1% lignin (Daud & Ali, 2004) in which resulted
of having a high calorific value of 5500 kcal/kg (Banzon, 1980).

According to Food and Agriculture Organization of United Nation (2019),


Malaysia has a production of 2.718 million tons of paddy in 2018, in which alternatively
produced very high amount of biomass product. Each ton of rice paddy harvested will
generate 0.22 ton of rice husk and 0.29 ton of rice straw (Pode et al, 2015). Therefore,
it can be summed up that Malaysia has produced 598.18 thousand tons of rice husk
and 788.51 thousand tons of waste rice straw in that year. It shows that the rice waste
is abundance and a good raw material to be used as a substitute in producing low-
cost catalyst. Since the source of fuel used in the synthesis of catalyst can affect the

5
yield in the production of biodiesel shown from previous work, it is recommended to
use the waste as source of fuel in the synthesis of hydrotalcite.

With reasons of the agricultural waste is always readily available, abundance,


renewable, and virtually free (Khalaf et al, 2019), various recent studies were
conducted to utilize the agricultural waste. The most common methods of utilization
of agricultural waste are convert the waste into fertilizer or activated carbon as
adsorbents or filtration materials for heavy metal, dye, or organic compound removal
(Babel & Kurniawan, 2004; Praveena et al, 2019). Researcher also took step further
in the use of agricultural waste in asphalt mixtures in reducing the traffic noise
coefficient (Norhafizah et al, 2016). It can be concluded that agricultural waste has a
potential to be used which can result in value added product.

1.4 Problem Statement

The current catalyst used in transesterification of waste cooking oil into


biodiesel is the heterogeneous catalyst reaction. In this study, hydrotalcite, a solid
catalyst is used in the transesterification reaction due to having advantage on high
resistance towards free fatty acid and the water content in waste cooking oil (Atadashi
et al, 2012; Di Serio et al, 2007; Zeng et al, 2009). It is expected to reduce or eliminate
the effect of free fatty acid and water content on low-cost feedstock towards
production of biodiesel to obtain high yield of biodiesel.

In this study, combustion method was chosen as the method of preparing the
hydrotalcite. Combustion method was chosen due to due to easy steps and requires
less time to prepare. The combustion method also offers tunable properties by the
intensity of the combustion reaction through varying the type of fuels (Lazarova et al,
2019). In the current method of hydrotalcite synthesis via combustion method, sugar
or saccharose is used as a fuel for constructing the layered structure for better
biodiesel yield as proved by various researchers (Anuar & Abdullah, 2016; Coelho et
al 2017). However, in this method, using sugar or saccharose is considered limitation
or disadvantage in the production of biodiesel. The reason of the disadvantage of
using sugar or saccharose in the synthesis of hydrotalcite for the production of
biodiesel is that it is not suitable as sugar is a food product. It might also cause a

6
social impact to the society such as food vs fuel debate. It also could potentially
increase the price of sugar as sugar has now become another demand to produce
catalyst to produce energy instead was supplied to the country as food. Therefore, a
novel initiation of using waste product as a sugar substitute for the synthesis of
catalyst to produce biodiesel is proposed. Therefore, the novelty in this study is the
substitution to agricultural waste from saccharose in the preparation of hydrotalcite
via combustion method.

As proposed for sugar substitute in the synthesis of hydrotalcite, three waste


candidates are chosen. These candidates are coconut shell, rice straw and rice husk.
In Malaysia, the mentioned wastes are abundance and available throughout the year.
The data from Food and Agriculture Organization (2019) were used to estimate the
amount of the waste candidates produced annually. Accordingly, there are about 93
thousand tonnes of coconut shell, 841.55 thousand tonnes of waste rice straw and
638.42 thousand tonnes of rice husk produced in 2018.

As a result of using or substituting the element, which is the fuel for the
initiation of combustion during calcination steps in the preparation of hydrotalcite, the
structure of hydrotalcite may change or alter. Several researchers found out that the
best calcination temperature for synthesizing hydrotalcite catalyst was in range of 550
°C to 750 °C, therefore, the calcination step will be kept constant within the range.
Due to the possibility of property change on the synthesized hydrotalcite towards the
controlled hydrotalcite, therefore it is essential to analyse the physical characteristics
of the synthesized hydrotalcite. The result of change in property of the hydrotalcite, it
will be subjected to the catalytic performance of the synthesized hydrotalcites towards
the biodiesel production, followed by kinetics of the synthesized hydrotalcites in the
transesterification reaction of biodiesel production.

1.5 Objectives

1) To synthesize hydrotalcite via combustion method using the agricultural waste


i.e. coconut shell, rice straw and rice husk as fuel.
2) To investigate the physico-chemical characteristics of synthesized
hydrotalcite.

7
3) To perform the catalytic performance of the synthesized hydrotalcite in
biodiesel production.
4) To perform the kinetics of the synthesized hydrotalcite in catalyzing the
transesterification reaction of biodiesel production.

1.6 Scope of Study

In this study, three waste candidates are chosen to be the substitute for
saccharose as fuel during the synthesis of hydrotalcite via combustion method. The
metal nitrates were mixed followed by sodium carbonate and the addition of fuel,
which are the candidates. These candidates are coconut shell, rice straw and rice
husk. Then, the solid mixtures were calcined at desired temperature. The calcination
temperature for the synthesis of hydrotalcite also was also varied i.e. 550 °C, 650 °C
and 750 °C.

Calorific value of waste as fuel is one of essential property that allows a better
layered structure of catalyst. The structure of the catalyst should be layered double
hydroxide and contains large surface area. Therefore, to ensure the success of the
study, it is important to determine the calorific values of the agricultural waste
candidates used to as an alternative in the synthesis of hydrotalcite. Bomb calorimeter
was used to determine the calorific value of mentioned fuels. The performance of
these calorific values must match or exceed that of saccharose which was 16.49
MJ/kg (Anuar & Abdullah, 2016). The characteristics or structure of synthesized
hydrotalcites were analysed via XRD, FTIR, and TGA to understand the effect of the
fuel and calcination temperature.

After analyzing the characteristics, the hydrotalcites were subjected into the
performance of catalyzing transesterification reaction of waste cooking oil into
biodiesel with 15:1 methanol to oil molar ratio and 8 wt% catalyst loading. The
parameter for the transesterification reaction is 65 °C and the reaction time is 5 hours.
The performance of hydrotalcite in transesterification reaction is observed via the
successful amount of the FAME yield produced via GC analysis. Synthesized
biodiesel was further analyzed via FTIR to ensure FAME is produced. Then, from the
GC analysis, the best hydrotalcite was undergone three respective cycles to
investigate the robustness of the catalysts.

8
The analysis is followed by the kinetic study of hydrotalcite catalyzed
transesterification reaction. The best catalyst (waste fuelled) and reference
hydrotalcite (sugar fuelled) were subjected into catalytic transesterification reaction at
various reaction temperatures i.e. 55 °C, 60 °C, 65 °C and 70 °C. The aim of the study
is to determine the effect of reaction temperature towards transesterification reaction.
The FAME yield was recorded at each hour for the whole 5-hour reaction of
transesterification reaction. The reaction rate used for the kinetic study is pseudo first
order. Arrhenius equation was used to obtain the activation energy for the best
catalyst (waste fuelled) and reference hydrotalcite (saccharose fuelled) to verify the
efficiency of the catalyst. As a conclusion, the novelty in this study is the substitution
to agricultural waste from saccharose in the preparation of hydrotalcite via combustion
method.

9
CHAPTER 2

LITERATURE REVIEW

2.1 Biodiesel Feedstock

Feedstock in biodiesel plays very important aspects in the production of


biodiesel. It is known that biodiesel can be produced from wide variety of vegetable
oil. Currently, there are more than 95% of edible neat and pure vegetable oil are used
to produce biodiesel i.e. soybean (Aguieiras et al, 2020; Udayakumar et al, 2020),
rapeseed (Udayakumar et al, 2020), sunflower (Dahdah et al, 2020), palm (De & Boxi,
2020) and others. However, to produce good quality of biodiesel, a good quality of
feedstock is needed such as low Free fatty acid (FFA) content (< 3 wt%), low acid
value (< 5 mg KOH/g), low water content (< 0.06 wt%) and others (Atadashi et al,
2012; Leung et al, 2010). Dahdah et al (2020) successfully converted sunflower oil
into biodiesel with yield of 95% using Ca-Mg-Al catalyst. Salimi & Hosseini (2019) also
successfully converting canola oil into biodiesel with 95.43% FAME yield using
ZnO/BiFeO3 catalyst. Coral et al (2019) successfully produce 95.6% FAME yield from
soybean oil using hydrotalcite. Coelho et al (2017) also successfully synthesize
biodiesel with 85% yield using hydrotalcite. It shows that neat or pure vegetable oil
has proven to be successful in the production of biodiesel.

However, the use of edible vegetable oil can cause a negative impact to the
human such as the depletion of food supply, creating high demand in food industry
and others. If the vegetable oil is used on the production of biodiesel and it has high
demand, farmers will shift its plantation to the vegetable oil for biodiesel production
with high demand as it can give a large reward, hence reducing the current food crops
for food as it is to fulfill the market demands. As a result, the crops for food supply will
decrease and the price of food is increasing as the food becomes high demand.
Alexander & Hurt (2007) concluded that a tremendous increase of crops usage for
the fuel which would result in the increase of food cost for consumers. Therefore,

10
researchers provide another alternative with the initiative to prevent the use of food
crop. Several types of non-edible oil that used as a source of biodiesel are castor
(Varma et al, 2007), lesquerella (Kumar & Gupta, 2020), Jatropha curcas and others
(Sundus et al, 2017). Helwani et al (2016) successfully transesterified jatropha curcas
oil into 91.2% biodiesel using hydrotalcite. Chowdhury et al (2019) successfully
converted madhuca indica oil into 81.56% biodiesel yield using CaO from waste
eggshell. It can be summed up that non-edible vegetable oil also has proven can be
successful as feedstock in the production of biodiesel.

However, the use of non-edible vegetable oil for production of biodiesel could
also possesses the same problems as that of an edible vegetable oil. It is due to the
land area requirement to plant the non-edible oil crop in order to be of a land to be
generated and harvested, in which the land is could be used to generate food crops.
It will not be a problem unless the non-edible oil plant is able to grow at an area where
food crops are unable to grow such as degraded land (Goswami & Choudhury, 2019).
As the price of petroleum crude oil has been hitting records height every other day,
vegetable oil has become one of potential sources to produce biodiesel. However, the
production of biodiesel from vegetable oil has been criticized from various non-
governmental organizations with the concern of extinction of animal, deforestation
and food versus fuel agendas (Lam et al 2009). Consequently, it has affected the food
production by land conversion which has resulted in food inflation which occurred in
U.S.A since 1990 (Alexander & Hurt, 2007; Goswami & Choudhury, 2019). With that
reason, an alternative feedstock is required to encounter the problems for assisting in
the growing of the biodiesel production field. Therefore, one of the possible solutions
to overcome the problems of biodiesel feedstock is to use the unwanted used oil which
was discarded by human such as waste cooking oil. As it will not be competing with
the food crops nor depleting the any food supply, it eventually will give benefit to the
community by reducing the amount of discarded waste oil generated by the human
population (Ramli & Farooq, 2015).

11
2.2 Biodiesel from Waste Cooking Oil

Waste cooking oil (WCO) is one of the best sustainable, renewable and
suitable to be used as a feedstock for the production of biodiesel (Ajala et al, 2015;
Allah & Alexandru, 2016). The use of WCO for the production of biodiesel could
reduce the impact in food supply in terms of the competition for the use of land crops
for either the vegetable oil is edible or non-edible (Alexander & Hurt, 2007). WCO is
abundant, low cost and readily available in almost all restaurants which provide fried
food. A survey of the usage and management of WCO has been conducted by
Hanisah et al (2013) and Kabir et al (2014) in Malaysian community where the
researchers found out 50% - 60% of the respondents took a step of disposing the
used cooking oil into the river, bins and drainage system. As a result, it is estimated
that Malaysia has generated a preponderance of 0.5 million ton of waste cooking oil
per year (Ramli & Farooq, 2015). It would be beneficial if the wasted oil is to be
collected and converted into biodiesel. With the use of WCO as feedstock, the cost
production of biodiesel can be effectively reduced up to 70% (Ramli & Farooq, 2015;
Raqeeb & Bhargavi, 2015). However, the WCO requires a proper management where
it requires be collecting and storing properly.

There are various studies were conducted by researchers regarding the use
of WCO as a feedstock to produce biodiesel. Amani et al (2016) successfully
converted waste cooking palm oil into 90% within 3-hour reaction using CsM-SiO2.
Anuar & Abdullah (2016) successfully converted WCO into 70.67% biodiesel in 5-hour
reaction using hydrotalcite as catalyst via stirring method. Zango et al (2019) also
conducted an optimization studies for catalytic conversion of waste vegetable oil into
biodiesel. The researcher found out by utilizing CaO could obtain 70% biodiesel yield
and 90% biodiesel yield by utilizing KOH as catalyst under same reaction
temperature, 65 °C for 4 hours reaction time. The studies show that waste cooking oil
has great potential as feedstock to be convert into biodiesel with an advantage of no
food interference as mentioned and compared with the use of edible and non-edible
oil.

12
2.3 Biodiesel Production

Biodiesel can be produced from various processes and methods. The process
of producing biodiesel has been increasingly positive with latest developed
technology such as micro-emulsion (Antunnes et al, 2017), pyrolysis (Gao et al,
2019), and transesterification and esterification (Krishnamurthy et al, 2020). In
pyrolysis method, biomass is mostly used for conversion towards biodiesel such as
algae biomass, woody biomass (Demiral & Şensöz, 2008), woody biomass (Aho et
al, 2008), millcakes (Pütün et al, 2006), municipal solid waste (Gao et al, 2019) and
others. However, the use of thermal decomposition towards the biomass involves
slow production, low energy yield and excessive air pollution which resulted in less
favorable method (Jahirul et al, 2012). The most adopted and widely used method for
producing biodiesel is homogeneous alkali-catalyzed transesterification of triglyceride
(Atadashi et al, 2012; Rahul et al, 2011; Reyero et al, 2013; Liu et al, 2017). In
microemulsion method, the feedstock oil will be emulsified with alcohol with presence
of surfactant. Another process for production of biodiesel via transesterification
process. In a typical transesterification reaction, triglyceride or triacyglycerol of
vegetable oil or animal fat will react with alcohol, to produce fatty acid methyl ester
(FAME) and glycerol as by-product (Birla et al, 2012; Leung et al, 2010; Mansir et al,
2017). Figure 2.3 shows the transesterification of triglyceride with methanol to
produce FAME or also known as biodiesel. Several journals have reported that
transesterification reaction is the most favorable method due to mild temperature
reaction and fast reaction time. Therefore, with these reasons, transesterification
method is chosen for this study.

Catalyst

Triglyceride Methanol FAME Glycerol

Figure 2.1: General chemical equation for transesterification reaction of triglyceride

13
2.4 Catalyst for Transesterification Reaction
2.4.1 Homogeneous Catalyst

Catalyst is one of the most essential items required in the production of


biodiesel via transesterification reaction. The most common homogeneous base
catalyst used for a commercialize process for producing biodiesel which has been
most favorably performed is potassium hydroxide or sodium hydroxide (Mardhiah et
al, 2017). The use of homogeneous base catalyst in production of biodiesel considers
to be advantageous due to three reasons which are low reaction temperature, high
FAME yield in minimal reaction time and widely available (Lam et al, 2010). Colluci et
al (2005) successfully converted soybean oil to biodiesel with a yield as high as 99.4%
by using molar ratio of alcohol to oil of 6:1 at 40 °C for 15 minutes. Similar findings
from Thanh et al (2010) and Stavarache et al (2003) where they successfully
converted canola oil and soybean into biodiesel with high yield which are 99% and
98% respectively. Table 2.1 shows reviewed base catalysts used in production of
biodiesel via transesterification process. From Table 2.1, the use of homogeneous
catalyst in transesterification requires low reaction time ranging 15 – 20 minutes with
high yield of biodiesel as high as 99.4%. Although the process is fast and simple, the
process is only suitable for neat or pure vegetable oil.

In a typical fresh vegetable oil from farm the free fatty acid (FFA) content is
typically low which makes homogeneous base catalyst is very effective to be used to
convert oil into biodiesel which is >0.5 wt% (Leung & Guo, 2006). However, in WCO,
low cost of feedstock consists of unpredictable and inconsistent value of FFA and
water content. With the usage of sodium hydroxide or potassium hydroxide as
catalyst, the catalyst reacts as side reaction with another component to form
undesirable side product such as formation of soap.

14
Table 2.1: List of homogeneous acid and base catalyst used in transesterification reaction
Amount
Ratio Temperature Time Yield
Feedstock Catalyst Catalyst Source
(wt%) (Alcohol:Oil) (°C) (minutes) (%)
Base Catalyst
Neat vegetable oil NaOH 0.5 6 to 1 25 60 80 Stavarache et al, 2003
Soybean KOH 1.5 6 to 1 40 15 99.4 Colucci et al, 2005
Waste frying oil NaOH 1.1 7 to 1 60 20 88.8 Leung & Guo, 2006
Canola oil KOH 0.7 5 to 1 25 50 99 Thanh et al, 2010
Dairy waste scum KOH 1.2 6 to 1 75 35 96.7 Sivakumar et al, 2011
Madhuca indica KOH 0.75 6 to 1 55 60 95 Singh et al, 2011
WCO KOH 1 6 to 1 70 60 90 Agarwal et al, 2012

Acid Catalyst
Waste frying oil H2SO4 3.8 245 to 1 80 240 99 Zheng et al, 2006
WCO H2SO4 1 90 to 1 95 600 90 Wang et al, 2006
Canola oil AlCl3 5 24 to 1 100 1080 98 Soriano Jr et al, 2009
Canola oil ZnCl2 5 60 to 1 100 1440 48 Soriano Jr et al, 2009

15
However, the use of homogeneous catalyst poses great disadvantages for
biodiesel production. As the homogeneous basic catalyst reacts with FFA in the
feedstock, it produces undesirable side products which are soap and water, known as
saponification, as shown in Figure 2.4 (Endalew et al, 2011). This side reaction could
lead to poor yield of biodiesel. Anya et al (2013) reported that as the value of fatty acid
increases, the yield of biodiesel decreases significantly as low as 20% with the
presence of 1.53% FFA. Therefore, by knowing the parameters which could lead to the
mitigation yield of biodiesel, it is important to ensure the WCO undergo a pretreatment
process to reduce the cause leads to poor yield of biodiesel. Leung et al (2010)
suggested that if the acid value of feedstock is above 5 mg KOH/g, the pretreatment of
feedstock is necessary before the main transesterification process takes place.

FFA Homogeneous Soap Water/


Catalyst Methoxide

Figure 2.2: Undesirable side reaction in transesterification reaction; saponification


(Endalew et al, 2011)

Another catalyst that is worth to be mentioned is homogeneous acid catalyst,


which is another catalyst that used for transesterification reaction. The most common
homogeneous acid catalyst used for transesterification reaction is sulphuric acid or
Lewis acid. However, the utilization of homogeneous acid catalyst in transesterification
reaction requires an excess amount of methanol to oil molar ratio and long reaction
time to produce high FAME yield. Wang et al (2006) reported that 90% of WCO is
successfully converted into biodiesel with mole ratio of alcohol to oil of 20:1 at 95 °C
for 10 hours. Similar findings to Zheng et al (2006), where they used H2SO4 as a
catalyst to convert waste frying oil (WFO) into biodiesel at 80 °C for 4 hours. From
reviewed acid catalysts used in production of biodiesel via transesterification reaction
in Table 2.1, homogeneous acid catalyst used high molar methanol to oil ratio typically
ranging from 24 – 245. The reaction time and reaction temperature for homogeneous
acid catalyst in transesterification reaction is ranging from 4 – 24 hours at 80 – 100 °C.
It can be concluded that the method of transesterification of vegetable oil using acid
catalyst does not gained much attention and unfavorable due to the requirement a large

16
amount of reaction time, high reaction temperature and an excess of alcohol for high
yield of biodiesel (Leung et al, 2010).

In conclusion, homogeneous base catalyst such as KOH or NaOH has been


favorably performed in production of biodiesel (Huaping et al, 2006; Rahul et al, 2011;
Reyero et al, 2013). The use of homogeneous catalyst in transesterification can cause
major drawback. The use of homogeneous acid catalyst such as H2SO4 in esterification
reaction (pretreatment before transesterification) could leads to high sulfur content in
biodiesel product (Mardhiah et al, 2017). The use homogeneous catalyst (NaOH or
KOH) in transesterification reaction could cause in in catalyst recovery difficulties and
reusability, corrosive wastewater generation from purification process, addition of
neutralization steps and difficulty in biodiesel purification process (Abou-Elyazed et al,
2019; Reyero et al, 2013). All the drawbacks and the disadvantages of homogeneous
catalyst could result in additional biodiesel production cost (Ali et al, 2019; Mardhiah et
al, 2017). As a result, it has led researchers anticipate in the application of
heterogeneous catalysts such as CaO, SnCl2, Cs0.6Zr0.4/Al2O3, hydrotalcite, or others
as a reason of heterogeneous catalyst can reduce the major drawback of utilization of
homogeneous catalyst (Borges & Díaz, 2012; Gao et al, 2019; Rahul et al, 2011).

2.4.2 Heterogeneous Catalyst

With the drawback of homogeneous catalyst in transesterification reaction, the


application of heterogeneous catalyst superseded homogeneous catalyst in production
of biodiesel. Heterogeneous catalyst acts in a different phase with reactants in
transesterification reaction which is contrasting to that of homogeneous catalyst.
Heterogeneous catalyst has been receiving positive attention to the researchers due
to the prospect of increased yield at reduced operating costs (Helwani et al, 2013). In
addition, researcher also reported the cost for biodiesel production by heterogeneous
is 4 – 20% less than that of homogeneous catalyst (Ramli et al, 2017). The most
significant reason for the approach to heterogeneous catalyst from homogeneous
catalyst is due to the catalyst could offer an easy catalyst separation via centrifugation
(Borges & Díaz, 2012; Gao et al, 2019; Rahul et al, 2011). With that fact, it has led the
advantage of heterogeneous catalyst to be able to be reuse for the next
transesterification reaction. Some researchers able re-calcine the catalyst to restore
the active site of the used catalyst from clogs of vegetable oil (Amani et al, 2016; Anuar

17
& Abdullah, 2016; Benedictto et al 2019; Onoji et al, 2017). Moreover, the use of
heterogeneous catalysis in transesterification reaction could also prevents the
undesirable reaction such as saponification, which adding more simplification to the
production process and less product loss (Antunes et al, 2008). Regarding
environmental aspects, the use of the heterogeneous catalyst in production of biodiesel
would result in reduced risk that associated with spillage or leakage of hazardous and
corrosive chemicals (Abbaszaadeh et al, 2012). Another additional environmental
benefits can be expected from the use of heterogeneous catalyst in production of
biodiesel is the absence of energy intensive and waste generated from glycerine
purification step (Abbaszaadeh et al, 2012).

The heterogeneous has become a better alternate and more environmentally


friendly catalyst as comparable with homogeneous catalyst which has attracted more
researcher to shift view towards heterogeneous catalyst (Nasreen et al, 2018). Table
2.2 shows reviewed heterogeneous catalyst used for transesterification of vegetable
oil into biodiesel.

18
Table 2.2: List of heterogeneous catalyst used in transesterification reaction

Amount Reaction
Ratio Time Yield Reusability
Feedstock Catalyst Catalyst Temperature Source
(wt%) (Alcohol:Oil) (°C) (hr) (%) (cycle)
WCO CsM-SiO2 3 20 to 1 65 3 90 4 Amani et al, 2016
WCO SiO2 3 20 to 1 65 3 4 - Amani et al, 2016
Waste oil SnCl2.2H2O 5 3.5 to 1 100 3 63 Yes Pereira 2014
Jatropha oil CaO (eggshell) 2.5 8 to 1 65 2.5 95 10 Singh et al 2011
Sunflower oil MgO 3 10 to 3 65 15 45 2 Benedictto et al 2019
Sunflower oil K+/MgO 3 10 to 3 65 0.5 98 2 Benedictto et al 2019
Oleic acid Ui-66(Zr)-NH2 6 39 to 1 60 4 97 3 Abou-Elyazed et al 2019
PURAL
Sunflower oil 2 12 to 1 60 24 50 1 Navajas et al 2010
Hydrotalcite
WCO Hydrotalcite 8 15 to 1 65 5 70.67 3 Anuar & Abdullah 2016
Jatropha oil Hydrotalcite 4 12 to 1 65 6 75.2 1 Helwani et al 2013
Soybean oil Hydrotalcite 20 20 to 1 120 12 43 3 Coelho et al 2017
Hydrotalcite (corn
Soybean oil 20 20 to 1 120 12 73 3 Coelho et al 2017
starch)
Soybean oil Li-Al Hydtotalcite 1 15 to 1 65 1 71.9 - Shumaker et al 2008

19
From Table 2.2, it is observed that heterogeneous catalyst used moderate
amount of molar methanol to oil ratio typically ranging from 8 to 20. The reaction time for
heterogeneous catalyst in transesterification reaction is ranging from 4 – 24 hours. In
addition, the reaction temperature used for transesterification reaction is 60 – 65 °C.
Amani et al (2016) investigated the effect of cesium metal in silica catalyst. The
researchers found out the addition of 30% cesium metal on silica catalyst increases the
catalytic performance of biodiesel synthesis from 4% to 90% in 3 hours under 65 °C
(Amani et al, 2016). Benedictto et al (2019) also examined the effect of K+ in MgO
catalyst. It was found that with addition of 20 wt% of KOH during the synthesis of MgO
catalyst has resulted the increase of catalytic performance of transesterification from
45% to 98% FAME yield under 2 hours. In addition, Xie & Huang (2006) also found out
that the FAME yield increases from 35% up to 87% with the increase of methanol to oil
molar ratio from 4:1 to 25:1. Hence, the use of exact stoichiometric value could reduce
the yield of biodiesel. Xie et al (2006) conducted the effect of methanol to oil molar ratio
in the production of biodiesel using hydrotalcite ranging from 2:1 till 20:1. From the study,
it was found that 15:1 is the best molar ratio of methanol to oil for transesterification
reaction. Therefore, to use the heterogeneous catalyst in transesterification reaction, it
is recommended to use excess of methanol to achieve higher FAME yield. In addition,
with reviewed literature review, it is decided that 15:1 methanol to oil molar ratio for
transesterification reaction is used in this study.

Although heterogeneous catalyst has been gained positive attention from


researchers, there are several drawbacks on use of heterogeneous catalyst. The
drawback of using heterogeneous catalyst in the production of biodiesel are catalyst
robustness, resistivity towards FFA in oil feedstock and water content in oil feedstock.
Robustness is essential property of heterogeneous catalyst. Poor robust of catalyst
could leads to poor FAME yield. Calcium oxide (CaO) is one of favorable catalyst used
in production of biodiesel due to the able to be synthesized via calcium-rich waste
materials are available in nature, such as eggshells (Yusuff et al, 2018), mollusk shells
(Boro et al, 2011) and bone (Obadiah et al, 2012), which can be used as raw materials.
However, CaO derived from waste material shows poor robustness. Boro et al (2011)
synthesized CaO from Turbonilla striatula snail shell. The researcher found out on the
fourth cycle of transesterification; the FAME yield dropped to 20% from 93.3%. CaO also
were reported to be leached during transesterification which could lead to contamination
to the products (Marinković et al, 2016). The poor robustness and leachability of CaO is

20
due to failure to maintain the structure of CaO and transformed to other form such as
Ca(OH)2 (Boro et al, 2011).

Typical FFA content of neat or pure vegetable oil is less than 0.5 wt% (Leung &
Guo, 2006) However, WCO consist of variation of FFA content. Water content in WCO
also can be varied depending on the contamination of the feedstock. Therefore, a high
resistivity towards FFA content is the most desirable heterogeneous catalyst. With these
factors, hydrotalcite was chosen as a catalyst for production of biodiesel due to high FFA
resistance in WCO.

2.5 Hydrotalcite

Hydrotalcite (HT) is a Layered Double Hydroxide (LDH) anionic build up from a


brucite like layers (Dávila et al, 2008). The general chemical formula for layered double
2+ 𝑧−
hydroxide is [𝑀1−𝑥 𝑀𝑥3+ [𝑂𝐻]2 ]𝑥+ [𝐴𝑥/𝑧 ∙ 𝑛𝐻2 𝑂]𝑥− where M2+ is representing a divalent
metal such as Mg2+, Ca2+, Zn2+, Co2+ etc., M3+ is a trivalent metal such as Al3+, Fe3+, Cr3+,
Mn3+ and A with charged z- representing anion such as 𝐶𝑂32− , 𝑆𝑂42− and others (Dávila
et al, 2008; Gomes et al, 2011). Figure 2.3 shows a typical HT structure with Mg as
divalent metal and Al as trivalent metal.

Figure 2.3: Typical structure of Hydrotalcite

HT is also well known as an applicable catalyst for various applications such as


cannizaro reaction, fructose isomerization, photocatalytic degradation of chlorophenol,
transesterification and esterification (Anuar & Abdullah, 2016; Bhojaraj et al, 2019;
Ramos-Ramírez et al, 2018; Yabushita et al, 2019). HT has shown to great potential
heterogeneous basic catalyst due to the catalyst exhibits a considerable large basic

21
surface after calcination (Gomes et al, 2011; Martins et al, 2013). HT has drawn
significant attention in research community due to be able to synthesize from Earth-
abundant and non-toxic elements which are magnesium and aluminium (Coelho et al,
2017). The performance of HT in the transesterification reaction of palm oil to FAME
shows relatively high ranging from 51 – 75% (Navajas et al, 2018) under stirring method
conditions. As mentioned earlier that most heterogeneous catalyst possesses a
drawbacks such as leachability and sensitive towards FFA and water content, HT has
proved to be very robust as it shows no leaching problems in transesterification reaction
(Helwani et al, 2013; Anuar & Abdullah, 2016). HT can also resist the presence of water
content in oil feedstock up to 1 wt% as proven by Zeng et al (2009) and Di Serio et al
(2007) without affecting the FAME yield as the recommended water content in oil for
transesterification reaction is less than 0.5 wt% (Atadashi et al, 2012). In addition, HT
can also be reused for next transesterification process up to 3 consecutive cycles with
minimal reduction of FAME yield (Anuar & Abdullah, 2016; Coelho et al, 2017).
Therefore, HT has been chosen as heterogeneous catalyst in the study for the
production of biodiesel.

According to nomenclature and IUPAC, the ratio of Mg-Al should be 3:1. The
ratio of Mg-Al should be considered as it can affect the yield of biodiesel production. A
comparison between MgO, Al2O3 and Mg-Al ratio of HT varied from 1:1 to 4:1 towards
production of biodiesel conducted by Cantrell et al (2005). It was found that Mg-Al with
3:1 has the most active catalyst in comparison with MgO and Al2O3 for the production of
biodiesel. Xie et al (2006) conducted a research on the effect of Mg and Al ratio in the
synthesis of HT for production of biodiesel. The ratio used in their research was from 2:1
to 4:1. It was found that the best Mg and Al ratio for production of biodiesel in their study
is 3:1 with FAME yield of 68% (Xie et al, 2006). With the mentioned reasons, it is decided
to use 3:1 of the Mg and Al ratio in the synthesis of HT for production of biodiesel.

2.5.1 Hydrotalcite Synthesis Route

There are several methods on the synthesis or preparation of HT such as co-


precipitation, sol-gel, combustion (Dávila et al, 2008). Sol gel method is a common
method for synthesizing TH catalyst. However, this method involves the application of
polymer matrix. The concept of Sol gel method is to homogenize the compound in liquid
and convert it back into solid state in a controlled manner. The advantage of sol gel is it

22
allows the mixing at atomic level and promotes homogeneous compound. If it is calcined
at high temperature, it will form ceramic fibers (Brinker & Scherer, 2013). The advantage
of sol gel methods is that it allows low temperature processing and sintering (Paredes et
al, 2006; Helwani et al, 2016). It also allows the geometry of catalyst to be flexible
(Brinker & Scherer, 2013). However, there are certain drawbacks on using the method.
The method involves several steps which is a complex process. It also requires a very
close monitoring. Co-precipitation method is another common method of preparing HT.
It involves the formation of HT via precipitation under neutral to alkaline condition ranging
from pH 7 to 13. (Paikaray & Hendry, 2014). Similar to sol gel method, the process is
complex, time consuming and requires a very close monitoring as it requires up to 3
days to allow crystallization (Coelho et al, 2017).

Combustion method is another alternative method to synthesize HT. It involves


three steps which are mixing, calcination and rehydration process steps. Combustion
method also offers several advantages on the formation of HT such as high purity,
homogeneity of catalyst and shorter preparation time (Anuar & Abdullah, 2016; Lazarova
et al, 2019). Combustion method also Combustion method has also become an
important technique for the synthesis and processing of advanced ceramics (structural
and functional), catalysts, composites, alloys, and intermetallic and nanomaterial
(Ahmadipour et al, 2012). In the preparation of HT via combustion method, metal
precursor, typically magnesium and aluminium nitrate salt in a desired ratio were mixed
in a solution followed by addition of sodium carbonate and fuel. Then the mixture was
calcined in furnace. In the furnace where the combustion of fuel occurred, the fuel
combusts and forms CO2 and H2O which acts as a source of C and H to construct
complexes with the metal ions (Martunus et al, 2011). The energy of the fuel combustion
initiated the reaction of constructing complexes where CO2 and H2O is absorbed by the
metal until it is promoted to high surface area by contribution of continuous heat energy
supplied to the precursor metal. (Othman et al, 2009; Dávila et al, 2008). Then finally the
metal oxide will be rehydrated with Na2CO3 to reconstruct the layered double hydroxide
structure.

In this study, combustion method was chosen for the synthesis of HT due to easy
steps and requires less time to prepare. The combustion method also offers tunable
properties by the intensity of the combustion reaction through varying the type of fuels
(Lazarova et al, 2019).

23
2.5.2 Combustion method

One of latest development to increase the catalytic activity of LDH catalyst is by


introducing and regulating the fuel used during the synthesis of catalyst via combustion
method. Previous studies shown that with the introduction of organic fuel in the synthesis
of HT via combustion method has proven to be having a positive result on the pore size
and the surface area of LDH structure. The types of organic fuel used in previous studies
are saccharose (Anuar & Abdullah, 2016), corn starch (Coelho et al, 2017), cellulose
(Sobhana et al, 2016), rice starch (Ramimoghadam et al, 2013), pollen grain (Hall et al,
2003), cyclodextrin (Ciobanu et al, 2013) and others. Besides, the use of organic fuel in
the synthesis of HT via combustion method is beneficially cheap, economical,
environmentally and renewable (Coelho et al, 2017; Ramimoghadam et al, 2013).

Sobhana et al (2016) discovered that with the use of cellulose as fuel in the
synthesis of HT via combustion method, the surface area of the synthesized HT
increases significantly as high as 152 m2/g from reference HT (no fuel) which is 36.9
m2/g. Coelho et al (2017) synthesized HT with the use of corn starch as fuel in the
synthesis of HT and found out that the yield of biodiesel with the use of HT synthesized
using corn starch was increased as high as 73% with reference HT (no fuel) of 43%.
Anuar & Abdullah (2016) conducted a study on the effect of fuel used in the synthesis of
HT towards biodiesel production. The fuels used in the study were glucose fructose and
saccharose. It was found that the calorific value of fuel used in the synthesis is directly
proportional to the surface area of the HT hence boost the performance of HT towards
biodiesel production (Anuar & Abdullah, 2016). This occurrence shows that the there is
an interaction between the organic fuel and HT which leads to a significant change to
the structure and the performance of catalyst towards the catalytic activity properties. As
mentioned earlier, in the calcination step in the synthesis of catalyst, the fuel liberate
heat and source of C and H in the form of CO2 and H2O which then these compounds
will be absorbed by the metal oxide periclase like structure and form the construction of
carbonate surface which known as layered structure (Dávila et al, 2008; Patil et al, 2002).

24
2.6 Agricultural Wastes as Fuel for Hydrotalcite Synthesis

In previous studies, saccharose or table sugar is one of most preferable fuel used
in the synthesis of HT via combustion method (Anuar et al, 2013; Anuar & Abdullah,
2016; Coelho et al, 2017; Dávila et al, 2008; Han et al, 2015; Haocui et al, 2019; Helwani
et al, 2013; Martunus et al, 2011; Zhang et al, 2020) where the performance of HT in the
transesterification reaction of oil to FAME shows relatively high ranging from 50%
(Coelho et al, 2017) to 75.2% (Helwani et al, 2013). Saccharose or known as sugar is a
carbohydrate which has molecular formula of C12H22O11 has a calorific value of 16.49
MJ/kg (Anuar et al, 2013; Anuar & Abdullah, 2016; Dávila et al, 2008). The advantage
of using saccharose is inexpensive, commercially available and has high calorific value
(Coelho et al, 2017; Dávila et al, 2008). However, saccharose is a carbohydrate which
often viewed as a neutral health filler which is the principal energy source in most
community throughout the world (Chambers et al, 2019). It is also one of essential food
that is used in daily basis. With the application of food source towards the synthesis of
catalyst to produce energy, which is biodiesel, it might initiate the food vs fuel debate to
a certain society such as a viewpoint paper written by Tomei & Helliwell (2016) on “Food
versus fuel? Going beyond biofuels”, an article by Jean Ziegler (2013), United Nation
Special Rapporteur on “Burning food crops to produce biofuels is a crime against
humanity” and many more. Therefore, agricultural waste is another alternative material
which can be used as fuel substitute in the synthesis of HT via combustion method. As
most agricultural is abundance, waste product and renewable, it will also not interfere
with food products. Besides, the use of agricultural waste has a potential to reduce the
amount of waste. In addition, it can potentially give a new added value to the application
of the waste instead most agricultural waste was not fully utilized and mostly burned in
an open burning (Rosmiza et al, 2014; Shafie, 2015).

Coconut is one of most major crops to be grown in Malaysia. It is the fourth


important crop to be planted in Malaysia after oil palm, rubber and paddy (DOSM, 2018).
According to Food and Agriculture Organization of United Nation (2019), Malaysia has
produced 517,589 tonnes in 2017. According to Banzon (1980) each 1.2 kg of coconut
could produce 0.18 kg coconut shell. Therefore, with 517,589 tonnes of coconut
generated in 2017, it eventually produces a preponderance amount of agricultural waste
approximate of 77 thousand tonnes of coconut shell were generated in that year. For a
long time in Malaysia, the coconut shell had a little to no economic value where the
disposal of the waste was not only costly but has caused an environmental issue (Li et

25
al, 2008). Coconut shell contains several types of carbohydrates. The coconut shell
contains 19.8% cellulose, 68.7% holocellulose and 30.1% lignin (Daud & Ali, 2004) in
which resulted of having a high calorific value of 5500 kcal/kg (Banzon, 1980).

According to Food and Agriculture Organization of United Nation (2019),


Malaysia has a production of 2.901 million tons of paddy in 2017, in which alternatively
produced very high amount of biomass product. Each ton of rice paddy harvested will
generate 0.22 ton of rice husk and 0.29 ton of rice straw (Pode et al, 2015). Therefore,
it can be summed up that Malaysia has produced 638.42 thousand tons of rice husk and
841.55 thousand tons of waste rice straw in that year. It shows that the rice waste is
abundance and a good raw material to be used as a substitute in producing low cost
catalyst. Since the source of fuel used in the synthesis of catalyst can affect the yield in
the production of biodiesel shown from previous work, it is recommended to use the
waste as source of fuel in the synthesis of HT.

As mentioned earlier, several researchers have reported that the agricultural


waste was left after harvest and disintegrated in open burning due to having little to no
economic value (Li et al, 2008; Praveena et al, 2019; Rosmiza et al, 2014). Therefore,
researchers began to utilize the waste to add value to the waste instead was left unused.
Various studies have been conducted to add value to the waste. Sangarunlert et al
(2007) utilized rice husk to produce furfural for the application in petroleum refineries.
Rafiee et al (2012) synthesized nanosilica from rice husk which can be utilized as
catalyst. Babel & Kurniawan (2004) converted coconut shell into charcoal for the removal
of Cr(IV) in wastewater. With the mentioned reason, agricultural waste has the potential
to be used to add more value to the waste to convert it into useful material.

However, abundancy of waste is not the only factor to be taken when choosing
a substitute. The waste should have plenty of similar matching properties to qualify it as
a substitute. One of the most important factors is the calorific value of the waste. In the
synthesis of HT via combustion method (Anuar & Abdullah, 2016; Coelho et al, 2017;
Martunus et al, 2011; Othman et al, 2009), saccharose is introduced as a fuel in aiding
the combustion method to enhance the characteristics and structure of HT. As it may
affect the food vs fuel issue, Table 2.3 shows comparison and possible alternatives for
the replacement of sugar in the synthesis of HT via combustion method.

26
Table 2.3: List of potential waste material used as fuel for synthesization of hydrotalcite
in comparison with current fuel.
Material Calorific value Composition Reference
(MJ/kg) (wt %)
Table 16.49 100% Sucrose Anuar & Abdullah (2016)
sugar
Rice husk 17.50 28.6% Hemicellulose Di Blasi et al (1999)
28.6% Cellulose
24.4% Lignin
18.4% Extractive Matter
Rice straw 15.30 35.7% Hemicellulose Xiao et al (2001)
32.0% Cellulose
22.3% Lignin
10.0% Extractive Matter
Coconut 20.08 48.9% Hemicellulose Daud & Ali (2004)
shell 19.8% Cellulose
30.1% Lignin

From Table 2.3, the parameters in which agricultural wastes matches the
property to the current fuel (saccharose) for synthesis of HT using combustion method
are cellulose and hemicellulose and holocellulose. Cellulose, hemicellulose and
holocellulose are under carbohydrate or saccharide group. Saccharide group is a
biomolecule consist of carbon, hydrogen and oxygen atoms. As mentioned earlier, the
fuel can be served as a fuel to produce energy in combustion (Othman et al, 2009). It
can also provide C and H atoms to construct complexes with metal ions (Martunus et al,
2011). However, the result is strongly dependent on the calorific values of the fuel as
demonstrated in Anuar & Abdullah (2016) work.

27
2.7 Reaction Kinetics

Kinetic study is one of important study to understand the behavior of chemical


reaction of methanolysis or known as transesterification reaction with respect to
triglyceride and methanol with presence of catalyst (Krishnamurthy et al, 2019).
Therefore, the study can be further be used in designing a reactor or pilot plant for cost
estimation (West et al, 2008). Therefore, various kinetic studies were performed by
researchers. Table 2.4 shows list of kinetic studies conducted from previous
researchers. There are many kinetic modellings that have been used by previous
researchers. The use of reaction order is depending on the concentration reactants in
the reaction from zeroth, first and second order reaction. In theory, transesterification
reaction is a reversible reaction of vegetable oil or animal fat known as triglyceride with
an alcohol to form fatty acid methyl ester (FAME) and glycerol as byproduct (Birla et al,
2012, Kaur & Ali; 2014; West et al, 2008). There are few parameters which could affect
the transesterification reaction kinetics such as reaction temperature, type of catalyst,
catalyst loading and alcohol to oil molar ratio.

28
Table 2.4: Activation energy values for various catalysts

Reaction Activation Energy,


Catalyst Oil Order of Reaction Reference
Temperature (°C) Ea (kJ/mol)
CaO WFO 50 – 65 79.00 Pseudo First - Order Birla et al (2012)
AENiCo WFO 60 – 80 23.99 Pseudo First - Order Yusuff et al (2018)
H3PW12O40 WCO 60 – 75 53.99 First - Order Talebian-Kiakalaieh et al (2013)
MCT WCO 60 – 100 21.25 Pseudo First - Order Yahya et al (2018)
BaO/CaO/ZnO Microalgal oil 45 – 65 48.03 Pseudo First - Order Singh et al (2019)
FeCl3 WCO 60 – 90 35.51 Pseudo First - Order Ma et al (2017)
CaO-CaCO3 Soybean oil 40 – 60 42.10 Second Order Hsieh et al (2010)
HT (Pural©) Soybean oil 60 – 120 31.00 Pseudo First - Order Di Serio et al (2012)
HT Fat poultry 60 – 120 56.8 Pseudo First - Order Liu et al (2007)

29
From the Table 2.4, it can be concluded that the range of activation energy for
transesterification reaction is varies from 21.25 KJ/mol – 79.00 KJ/mol and the most
common reaction order used is pseudo first – order. In reaction kinetics, reaction
temperature is crucial in the production of biodiesel. It is due to the reaction temperature
affects FAME yield. Numerous researches were conducted to find the optimum reaction
temperature for the reaction kinetics. From previous research, the optimum reaction
temperature for transesterification via stirring method is ranging from 45 °C – 80 °C (Birla
et al, 2012; Hsieh et al, 2010; Ma et al, 2017; Yusuff et al, 2018). In determination of
activation energy and frequency factor of reaction, only the reaction temperature and the
conversion are involved in the calculation.

Therefore, the use of excess of alcohol in transesterification is considered to be


advantageous and simplifies the calculation steps for determination of activation energy.
Due to the use of excess of reactant (alcohol) the 2nd order rate law of transesterification
is now obeying the pseudo 1st order rate law. It is due to the excess concentration of
alcohol which can be assumed to be constant during the whole reaction reaction takes
place (Birla et al, 2012; Kaur & Ali 2014; Singh et al, 2007). In addition, several studies
show that the use of excess of alcohol in transesterification or methanolysis could
increase the catalytic conversion. Xie & Huang (2006) conducted the effect of molar to
oil ratio in transesterification of soybean oil using KF/ZnO as catalyst. It was found that
with the increase of methanol to oil molar ratio from 4:1 to 10:1, the FAME yield increases
from 35% up to 85% (Xie & Huang, 2006). The studies were extended the effect of molar
to oil ratio up to 25:1 and the FAME yield was recorded at 87%. Therefore, it is
recommended to excess the use of methanol or reactant to achieve higher FAME yield.

For determination the kinetics of the reaction, several measures are required for
the calculation of the kinetics. The first measure was the catalyst used for the overall
reaction was in a sufficient amount with respect to oil (Birla et al, 2012). Therefore, the
reverse reaction and the change of concentration of catalyst during the reaction is
negligible (Birla et al, 2012; Zhang et al, 2010). The equation of transesterification
reaction is as follows:

𝑇𝐺 + 3𝑅𝑂𝐻 → 𝐹𝐴𝑀𝐸 + 𝐺 (Equation 2.1)

30
From equation 2.1, the reaction is assumed to be single step (Birla et al, 2012;
Huaping et al, 2006). Therefore, the rate law of transesterification is expressed as
Equation 2.2 where the overall reaction follows second order reaction law.

−𝑑[𝑇𝐺]
−𝑟𝑎 = = 𝑘′ ∙ [𝑇𝐺] ∙ [𝑅𝑂𝐻]3 (Equation 2.2)
𝑑𝑡

Where [TG] = concentration of triglyceride, [ROH] is concentration of methanol


and k’ is equilibrium rate constant. However, due to excess of methanol or very high
molar ratio methanol to oil, the change of methanol concentration can be considered as
constant throughout the reaction and it behaves as first order reaction. Thus, the reaction
obeys pseudo first order kinetics (Birla et al, 2012; Kaur & Ali, 2014; Singh et al, 2007;
Zhang et al, 2010). The rate can be express as follows:

−𝑑[𝑇𝐺]
−𝑟𝑎 = = 𝑘 ∙ [𝑇𝐺] (Equation 2.3)
𝑑𝑡

Where k is modified rate constant, 𝑘 = 𝑘′ ∙ [𝑅𝑂𝐻]3 . Then, it is assumed that the


initial triglyceride concentration was [TG0] at time t = 0 and at time t, it falls to [TGt]. The
integration of Equation 2.3 where [TG] = [TG0] and at t = t, [TG] = [TGt]. which will be
expressed as Equation 2.4 (Ali et al, 2019; Birla et al, 2012; Kaur & Ali, 2014):

𝑙𝑛[𝑇𝐺0 ] − 𝑙𝑛[𝑇𝐺𝑡 ] = 𝑘 ∙ 𝑡 (Equation 2.4)

The conversion of fatty acid methyl ester or biodiesel consists of methyl


palmitate, methyl stearate, methyl oleate and methyl linoleate. The conversion value of
biodiesel is depending on the yield of biodiesel (Ali et al, 2019). Therefore, from mass
balance,

[𝑇𝐺]
𝐹𝐴𝑀𝐸 = 1 − (Equation 2.5)
[𝑇𝐺0 ]

Or

[𝑇𝐺] = [𝑇𝐺0 ](1 − 𝐹𝐴𝑀𝐸) (Equation 2.6)

31
Where FAME is FAME yield. For in terms of conversion, the rate expression is
given as

𝑑𝐹𝐴𝑀𝐸
= 𝑘 ∙ [1 − 𝐹𝐴𝑀𝐸] (Equation 2.7)
𝑑𝑡
In which, upon integration and arrangement, the final equation is as follows.

𝑙𝑛(1 − 𝐹𝐴𝑀𝐸) = 𝑘 ∙ 𝑡 (Equation 2.8)

The final equation, Equation 2.8 was used to plot ln(1-FAME) versus reaction
time, t. The graph was plotted to determine the reaction rate constant value (k) where
the values can be obtained through the slope of the equation.

Arrhenius model equation was used to determine the activation energy and
frequency factor value of the catalytic transesterification reaction. The equation is as
follows:

−𝐸𝑎
𝑘 = 𝐴𝑒 𝑅𝑇 (Equation 2.9)

Where k is reaction rate constant, A is Frequency Factor, Ea is activation energy,


R is gas constant and T is temperature. By taking natural logarithm into the Arrhenius
equation and rearranged, the new equation is as follows:

−𝐸𝑎 1
𝑙𝑛 𝑘 = ( ) − 𝑙𝑛 𝐴 (Equation 2.10)
𝑅 𝑇

After plotting ln k versus 1/T and value of universal gas constant, R is 8.314
J/K·mol, the activation energy, Ea and frequency factor, A value can be determined.

32
2.8 Concluding Remarks

From the literature, it can be concluded that for this research, HT will be
synthesized via combustion method as the advantages are high purity, homogeneity of
catalyst and shorter preparation time (Anuar & Abdullah, 2016; Lazarova et al, 2019).
As most of drawbacks of heterogeneous catalyst was leachability/robustness and highly
sensitive towards FFA and water content in cooking oil feedstock, HT is very robust and
has low sensitivity towards FFA and water content in cooking oil feedstock. The novelty
of this study is the HT will be synthesized via combustion method and the fuel used in
assisting the combustion method is agricultural waste. Unlike, the previous study, where
HT was synthesized via combustion method where saccharose were used as fuel. Three
types of fuels were selected for use in the preparation of HT i.e. coconut shell, rice straw
and rice husk due to their abundancy and readily available. Then, the HT will be
subjected under transesterification reaction with WCO. The Mg-Al ratio will be used in
this research will be 3:1 as it was highlighted by Cantrell et al (2005) to be the most
active catalyst for production of biodiesel. Based on study conducted by Xie et al (2006),
the molar ratio of methanol to oil will be used for transesterification reaction is 15:1. All
results of HTs synthesized with agricultural waste as fuel will be compared with HT
synthesized with saccharose to determine the effectiveness of using agricultural waste
as fuel in the synthesis of HT.

33
CHAPTER 3

METHODOLOGY

3.1 Materials and Equipment

Coconut shell was collected on local grocery store as a waste from extraction of
grated coconut in Seremban, Negeri Sembilan. Then the coconut shell was grounded to
25 mesh by using pestle and mortar. Rice husk was collected at FAMA Building,
Seremban, Negeri Sembilan. Rice straw was collected from local farmer at Batu Kikir,
Negeri Sembilan, Malaysia. Collected rice straw was cut into small pieces ranging 2 cm
to 4 cm. Then, small pieces of rice straw and rice husk were powdered using a chopper
separately. Collected waste cooking oil from local household was filtered using cloth filter
for the removal of particulate matter and impurities.

For synthesis of HT, magnesium nitrate hexahydrate (HmBG Chemicals),


aluminium nitrate nonahydrate (HmBG Chemicals), sodium carbonate (Merck),
saccharose (Sigma Aldrich) and grounded coconut shell powder were used. Waste
cooking oil, methanol (R&M Chemicals) and synthesized HT were used to produce
biodiesel. n-hexane (Merck), methyl heptadecanoate (Sigma Aldrich), fatty acid methyl
esters (FAME) standards (Sigma Aldrich) were used for fatty acid methyl ester (FAME)
analysis.

For pre-experimental analysis on WCO for determination the average molecular


weight of WCO, potassium hydroxide pellet (Bendosen), phenolphthalein (HmBG
Chemicals) and hydrochloric acid (HmBG Chemicals) were used for determination of
saponification value. Whereas isopropyl alcohol (R&M Chemicals), phenolphthalein
(HmBG Chemicals) and potassium hydroxide pellet (Bendosen) were used for
determination of acid value.

34
Table 3.1: List of all equipment used in this research.
Equipment Brands Purpose of Use

Bomb Calorimeter Cussons Technology Calorific value determination of


agricultural waste fuel.

Drying Oven Memmert For drying HT after recrystallization step.

FTIR Perkin Elmer For FTIR spectra analysis of FAME,


WCO and HT.

Furnace Carbolite For calcination of as-synthesized HT.

Gas Chromatography Perkin Elmer Determining amount of FAME yield in


percentage.

Hotplate Stuart For providing heat in transesterification


reaction.

Hydraulic Press International Crystal Briquetting samples for bomb


Laboratories calorimeter test.

Surface area Analyzer Micromeritics For surface area analysis (BET) of HT.

TGA/DSC Mettler Toledo Thermogravimetric analysis of as-


synthesized HT.

Weighing Balance Mettler Toledo For weighing and pre-calculation of


samples and raw materials.
X-ray Diffractometer PANalytical To verify the structure of HT.

35
3.2 Overall Experimental Flowchart

Preparation of HT via
combustion method

Fuel Type
Analysis
- Saccharose - Rice husk
- Calorific Test
- Rice straw - Coconut shell

Calcination temperature
- 550 °C - 650 °C
- 750 °C

Analysis
Prepared HT - XRD - BET
- TGA - FTIR
Transesterification reaction @ 65°C

Analysis
FAME
- GC - FTIR

Best HT

HT reused up to Reaction temperature


3 cycles in Reusability Reaction
of HT kinetics - 55 °C - 60 °C
transesterification
- 65 °C - 70 °C
reaction

Analysis Activation Analysis


FAME energy
- GC - GC

Figure 3.1: Overall experimental flowchart

The study was initiated with the preparation of HT via combustion method. The
fuel used for the synthesis of HT was saccharose, rice straw, rice husk, and coconut
shell. Then the HT was calcined to various calcination temperature i.e. 550 °C, 650 °C
and 750 °C followed by recrystallization step. The air flowrate used during calcination
step was 0 m/s and the duration was of calcination was 5 minutes. The catalyst was
ready to be used for transesterification reaction and characterization such as XRD, BET,
TGA and FTIR analysis. The fuels used in the preparation of HT such as saccharose,
rice straw, rice husk and coconut shell were put into a bomb calorimetric test for
determination of calorific value.

36
Average molecular weight of WCO was determined prior for pre-calculation of
molar ratio methanol to oil. After obtaining the value, all HTs were undergoing
transesterification with reaction temperature of 65 °C for 5 hours. After all reactions took
place, GC analysis will be carried out to determine the best catalysts. The best catalyst
will be used for the performance in the reusability of catalyst and reaction kinetics. For
reusability of catalyst, the best HT was undergone 3 consecutive transesterifications.
The HTs will be recovered after of each consecutive reaction, recalcined and
recrystallized. For kinetic studies, the best HT was undergone transesterification where
the reaction temperature was varied to 55 °C, 60 °C, 65 °C and 70 °C. Then, the FAME
yield of each hour of transesterification reaction conducted towards the best HT will be
determined prior to determine the activation energy of the catalyst towards
transesterification reaction. The overall experimental flowchart as described in Figure
3.1.

3.3 Preparation of Hydrotalcite

HT was synthesized by using combustion method which was adapted from


Dávila et al (2008) and Anuar & Abdullah (2016). The preparation of HT also was
conducted in the facility of UniKL MICET located in Simpang Empat, Melaka, Malaysia.
The collected coconut shell from grocery store was grounded to 25 mesh by using pestle
and mortar. Whereas rice straw and rice husk were powdered using a chopper
separately. The ratio of Mg:Al was set to 3. For the synthesis of 100 g HT, 67.2 g of
magnesium nitrate hexahydrate and 32.8 g of aluminium nitrate nonahydrate were
placed separately in 2 different beakers containing 80 ml of deionized water. Both
solutions were heated to 80 °C under stirring condition until all solids dissolved. After 5
minutes, both solutions were mixed slowly. 20 wt% of Na2CO3 (20 g) and 10 wt% of fuel
(10 g) from the total weight of nitrate metals were then added into the mixture. Then, the
mixture undergone a vigorous stirring effect and maintained at 80 °C until water had fully
reduced. The resulted paste was calcined in a furnace with calcination temperature of
650 °C for 5 minutes dwelling time with air flowrate of 0 m/s. The fuels used are
saccharose, grounded coconut shell, rice straw and rice husk. Then, the resulting MgAlO
catalysts were ground into powder form. The powdered catalysts were recrystallized by
an immersion in a 0.05 M Na2CO3 solution for 5 minutes. Then the catalysts were filtered
and washed with deionized water 2 times followed by drying in an oven at 40°C for 5
hours. The catalysts were kept in airtight container. The samples were labelled according

37
to the fuel used and calcination temperature tabulated in Table 3.2. HT-SS represents
HT synthesized using saccharose as fuel, HT-CS represents HT synthesized using
grounded coconut shell as fuel, HT-RH and HT-RS represents HT synthesized using
rice husk and rice straw as fuel. The function of magnesium nitrate hexahydrate
aluminium nitrate nonahydrate were to synthesis hydrotalcite. Sodium carbonate were
used to precipitate the metals (magnesium and aluminium). Saccharose, grounded
coconut shell powder, powdered rice straw and powdered rice husk were used as a fuel
for combustion during calcination for formation of layered structure hydrotalcite.

Table 3.2: List of all HT samples in this research


Samples Fuel Calcination Temperature
HT- SS Saccharose 550 °C, 650 °C and 750 °C
HT- RH Rice husk 550 °C, 650 °C and 750 °C
HT- RS Rice straw 550 °C, 650 °C and 750 °C
HT- CS Coconut shell 550 °C, 650 °C and 750 °C

3.4 Fuel Calorific Value Determination

Method of briquetting and determination of calorific value was adapted from


Mu’az et al (2016). Bomb calorimetric test was conducted to determine the calorific value
of fuel (coconut shell, rice husk, rice straw and saccharose) used by using Calorimeter
model P6310. The equipment also equipped with Beckmann thermometer. An
approximate of 1 gram of fuel was weighted and briquetted by using hydraulic presses
up to 28 MPa. Then the briquetted fuel was placed in a crucible where 0.1 mm diameter
of 100 mm length stainless-steel wire was used as ignition wire. As the crucible was
placed in a bomb calorimeter, in contact with the ignition wire. As the bomb calorimeter
was closed tightly, the bomb was compressed with 450 lb/in2 of oxygen gas. Then, the
bomb calorimeter was immersed in a 2000 g of water. Necessary electrical connections
were made, and stirrer was activated to ensure initial temperature was steady. When the
initial temperature was constant, the fuse wire was ignited, and the temperature was
recorded each minute throughout 20 minutes of experiment. The test was conducted in
the facility of UniKL MICET located in Simpang Empat, Melaka, Malaysia.

38
3.5 Average Molecular Weight Determination

In typical transesterification reaction, the molar ratio of methanol to oil used for
calculating the exact molar amount of methanol and oil needs to be placed in a reactor.
Therefore, it is essential to determine the value of average molecular weight of WCO.
For determination of average molecular weight of WCO, acid value and saponification
value were required. The collected waste cooking oil was filtered with cloth filter to
remove solid particulate matter. Then, the WCO was poured in a beaker covered on top
to prevent from any particulate contamination. The determination of average molecular
weight of WCO was adapted Huaping et al (2006) with calculation formula as follows:

𝑀 = 56.1 × 1000 × 3/(𝑆𝑉 − 𝐴𝑉) (Equation 3.1)

M represents average molecular weight of oil where SV is saponification value


(mKOH/moil, mg/g) and AV is acid value (mKOH/moil, mg/g).

3.5.1 Saponification Value Determination

Saponification value is a value that is required to determine the value of average


molecular weight of WCO. Saponification value was determined through a method
adapted from ASTM D5558. At first, an approximate of 1 gram of oil was suspended in
50 ml of 0.1 M KOH and drops of phenolphthalein. The pink solution was heated under
reflux condition for 3 hours. Then the solution was titrated with 0.1 M of HCl till the
solution become colorless. The amount of titrant was recorded. All the value was placed
in the equation below. Based on the equation, SV, V, M and W represents saponification
value, volume of HCl used in titration in milliliter, moles of HCl and mass of oil (sample)
in gram.

𝑉 × 𝑀 × 56.1
𝑆𝑉 = (Equation 3.2)
𝑊

39
3.5.2 Acid Value Determination

Acid value is also another part of value that is required to determine the value of
average molecular weight of WCO. The method of determination of Acid value of WCO
was adapted from EN 14104 by titration. For a typical determination of acid value, a
gram of oil was measured and suspended in a 125 ml of isopropyl alcohol in a conical
flask. 3 drops of phenolphthalein were added into the solution. The mixture was then
titrated using 0.05N KOH until the solution turned into light pink. The volume of titrated
KOH solution was recorded. Blank value was determined by repeating the steps but
without the presence of weight of oil. Below is the equation used to determine the acid
value.

(𝑉 − 𝐵) × 𝑁 × 56.1
𝐴𝑉 = (Equation 3.3)
𝑀

AV represent acid value whereas V, B, N, and M represents volume of titrant in


milliliter, volume titrated KOH solution titrated to blank in milliliter, normality of KOH
solution and mass of oil (sample) weighted in gram respectively. After obtaining SV and
AV value, those values were used for the calculation of WCO average molecular weight
as shown in equation 3.1 in section 3.6. The value of WCO average molecular weight
will be further used in pre-calculation for methanol to oil molar ratio for transesterification
reaction.

3.6 Transesterification Reaction of Biodiesel

The experimental method of transesterification reaction of WCO with methanol


was also adapted from Anuar & Abdullah (2016) where the reaction was conducted on
a double necked flask glass reactor in a water bath equipped with condenser and a
thermometer. The reaction temperature was set at 65 °C in stirring condition.

In a typical transesterification reaction, the weighted oil was initially preheated


for 5 minutes in the reactor for 5 minutes. Then, 8 wt% of catalyst was used from total
weight oil was placed in the reactor followed by 15:1 methanol to oil molar ratio. The
reactor subsequently was submerged into the 65 °C water bath up till the neck of reactor
to ensure the reaction temperature was maintained for the whole transesterification

40
reaction for 5 hours. All HTs were subjected to the transesterification reaction under 65
°C for 5 hours. Figure 3.2 shows the flowchart for biodiesel transesterification reaction
and Figure 3.3 shows the experimental setup for transesterification reaction of biodiesel.

WCO

Filtered via cloth

Prepared HT
Transesterification reaction@ 65 °C - HT-SS - HT-RS
- HT-RH - HT-CS

Analysis
FAME
- GC - FTIR

2 Best HT

Reusability of HT

Up to 3 cycles Analysis
FAME
- GC

Figure 3.2: Transesterification reaction of biodiesel flowchart

41
Cold water out

Cold water in

HT
WCO

Hot water bath

Magnetic
stirrer

Figure 3.3: Experimental setup for transesterification reaction

After the completion of 5 hours reaction, excess of methanol in the products was
removed via rotary evaporator at 80 °C. Then, the products were centrifuged at 3500rpm
for 30 minutes to separate upper layer and mid layer and lower layer which are biodiesel
(FAME) and glycerol (by-product) and the catalyst. The biodiesel layer will be focused in
throughout the research study.

For the reusability of HT, the experimental run only focuses on the most active
HT. After the centrifugation process took place, the catalyst was recovered and washed
with n-hexane. Then the HT was recalcined at 300 °C for 3 hours dwelling time. The HT
was washed with deionized water 2 times followed by drying in an oven at 40°C for 5
hours. Then a new cycle of transesterification reaction was started with new reactants
under same conditions. The experimental run was repeated up to 3 cycles.

42
3.7 Kinetic Studies of Transesterification Reaction

WCO

Filtered via cloth

Reaction Temperature
- 55 °C - 60 °C Best HT
- 65 °C - 70 °C

Transesterification reaction

Analysis Activation
FAME
- GC energy

Figure 3.4: Transesterification reaction of biodiesel process steps for kinetic studies

For kinetic study, the best HTs were subjected to the study. The experimental
method of kinetic study was adapted from Ali et al (2019). The weighted oil was initially
preheated for 5 minutes in the reactor for 5 minutes. Then, 8 wt% of catalyst was used
from total weight oil was placed in the reactor followed by 15:1 methanol to oil molar
ratio. The reactor subsequently was submerged into the reaction temperature water bath
up till the neck of reactor to ensure the reaction temperature was maintained for the
whole transesterification reaction for 5 hours. The reaction temperature carried for the
kinetic study of catalytic transesterification reaction was 55 °C, 60 °C, 65 °C and 70 °C.
The FAME was extracted each hour of the transesterification to obtain the FAME yield.
After obtaining the FAME yield, FAME yield vs time graph, ln (1-FAME) vs time graph
and Arrhenius equation were used to obtain activation energy of HT for
transesterification reaction as shown below:

−𝐸𝑎
𝑘 = 𝐴𝑒 𝑅𝑇 (Equation 3.4)

43
3.8 Catalyst Analysis
3.8.1 XRD Analysis of Hydrotalcite

To determine and verified the formation of layered structure of synthesized


catalyst HTs, the catalysts were subjected into XRD analysis. The method of XRD
analysis on HT was adapted from Liu et al (2016). The XRD analysis was carried out
using X’Pert Pro x-ray diffractometer manufactured by PANalytical. The crystallinity of
the synthesized catalysts was analyzed under 40 kV, 30 mA monochromatic CuKα (λ =
0.15406 nm). The 2θ measurements were recorded over a range from 10° to 80°. The
test was conducted in the facility of Quantum Skynet Solution Sdn. Bhd located in Nilai,
Negeri Sembilan, Malaysia.

3.8.2 BET Analysis of Hydrotalcite

The BET analysis of synthesized of HTs were analyzed to determine the size of
surface area of the HTs. The method of BET analysis on HT was adapted from
Thouchprastichai et al (2018). The analysis was measured from N2 adsorption and
desorption isotherm determined at 77K by using ASAP 2020 Surface Area Analyzer
manufactured by Micromeritics. The test was conducted in the facility of Quantum Skynet
Solution Sdn. Bhd located in Nilai, Negeri Sembilan, Malaysia.

3.8.3 FTIR Analysis of Hydrotalcite

The FTIR spectra of synthesized HTs were analyzed to determine the presence
or absence of specific chemical groups or chemical bands. The method of FTIR analysis
on HT was adapted from Yang et al (2020). The FTIR spectra measurements were
carried out by using Nicolet iS10 FTIR Spectrometer, manufactured by Thermo Fischer
Scientific. The samples were analyzed using ATR method. The ATR crystal used in this
analysis was monolithic diamond. In typical run, the sample was placed on the ATR
crystal and analysis was proceeded. The spectral resolution used in this analysis ranging
from 4000 – 600 cm-1. The test was conducted in the facility of UniKL MICET located in
Simpang Empat, Melaka, Malaysia.

44
3.8.4 TGA Analysis of Hydrotalcite

The characterization analysis of the catalyst was followed by thermogravimetric


decomposition to determine the organic decomposition of synthesized HT. The method
of TGA analysis was adapted from Zhang et al (2020). The analysis was carried out by
using Mettler Toledo TGA/DSC 1 STARe System. In a typical run, an approximate 15 mg
of catalyst was placed on a crucible. Then, it was analyzed under N2 atmosphere (20
cm3/min). The operating temperature of the analysis was 30 °C to 800 °C with ramp
temperature of 10 °C/min. The test was conducted in the facility of UniKL MICET located
in Simpang Empat, Melaka, Malaysia.

3.9 FAME Analysis


3.9.1 Gas Chromatography Analysis of FAME

The FAME yield was quantified by using gas chromatographic analysis. The
method of preparation was adapted from EN-14103. An approximate of 250 mg of
sample was weighted in 10 ml vial. Then, 5 ml of 10 mg/ml methyl heptadecanoate
solution was added in the vial. The mixture was shaken vigorously. 1 μl of diluted sample
was injected into gas chromatography Perkin Elmer AutoSystem XL equipped with FID.
The instrument also equipped with silica capillary column with dimension of 30 m × 0.25
mm × 2.5 µm (Varian Chrompack CP-Sil 8CB) manufactured by Spectralab Scientific.
Carrier gas used for the analysis was hydrogen gas. The injector temperature and
detector temperature were fixed at 250 °C. For column temperature, the initial
temperature was maintained at 100 °C. The ramp rate is 5 °C/min. After reaching 240°C,
the temperature was hold for 5 minutes. The test was conducted in the facility of UniKL
MICET located in Simpang Empat, Melaka, Malaysia.

3.9.2 FTIR Analysis of FAME

The FAME was analyze using FTIR analysis to determine the presence or
absence of specific chemical groups for verifying the existence of FAME in synthesized
products. The method of FTIR analysis on FAME was adapted from Mueller et al (2013).
The FTIR spectra measurements were carried out by using FTIR Spectrometer Perkin

45
Elmer Spectrum RX I. The samples were analyzed using ATR method. The ATR crystal
used in this analysis was monolithic diamond. In typical run, the sample was placed on
the ATR crystal and analysis was proceeded. The FTIR spectra of samples were
recorded in a range of 4000 – 600 cm-1 spectral resolution. The test was conducted in
the facility of UniKL MICET located in Simpang Empat, Melaka, Malaysia.

46
CHAPTER 4

RESULT AND DISCUSSION

4.1 Catalyst Characterization


4.1.1 XRD Analysis of HT

The first stage of characterization of HT is to identify the synthesized


catalyst is HT, a layered double hydroxide crystallite. Therefore, XRD analysis is
significant to verify the successful synthesis of HT. The structure and the
crystallinity of HT was confirmed via X-ray diffraction analysis. Figure 4.1 shows
the XRD pattern of recrystallized HT using saccharose, rice straw, rice husk and
coconut shell as fuel calcined at 550 °C – 850 °C. Based on Figure 4.1, HT was
successfully synthesized as indicated at diffraction peaks of 11.7°, 23.5°, 35.5°,
39.2°, 48.0°, 61.5° and 63.3° as per according to Navajas et al (2018), Shekoohi
et al (2017) and Yang et al (2020) marked as (*).

*
* * * * ** (a)

(b)
Intensity (a.u.)

(c)

(d)

(e)
(f)

(g)

10 20 30 40 50 60 70 80
2θ (degrees)

Figure 4.1: XRD patterns of (a) HT-SS 650, (b) HT-RS 650, (c) HT-RH 650, (d) HT-
CS 550, (e) HT-CS 650, (f) HT-CS 750, (g) HT-CS 850 and (*) HT peaks

47
As per according to previous researchers, the diffraction of peak on 11° and
23° indicated as a layered structure (Yang et al, 2020; Navajas et al, 2018). The
mentioned diffraction peaks exist on HT-SS 650, HT-RS 650, HT-RH 650, HT-CS
550 and HT-CS 650 which confirms the successful synthesis of HT. From the
observation of HT-CS calcined at 550 – 850 °C with 100 °C interval, it can be
concluded that to maintain the structure of HT, the calcination temperature should
not be above than 650 °C. Therefore, as per according to Navajas et al (2018),
Shekoohi et al (2017) and Yang et al (2020), it is confirmed that HT- HT-SS 650,
HT-RS 650, HT-RH 650, HT-CS 550, and HT-CS 650 has the chemical structure
of Mg6Al2CO3(OH)16·4H2O. However, the sample failed to maintain the layered
double hydroxide structure on above 650 °C is due to the heat provided during the
calcination step in synthesis of HT had caused the layered structure on HT to be
collapse by vibrational effect. As high temperature was exposed to the HT, the
molecules or atoms in HT began to vibrate where the layered structure suffers
dehydroxylate (Gao et al, 2018). As a result, the layered structure broken, or the
structure collapse as depicted on peaks of 11° and 23° on HT-CS 750 and HT-CS
850 in Figure 4.1 did not exist. In addition, from the observation on x-ray
diffractograms of HT-CS calcined at 550 °C, 650 °C and 750 °C, the intensity of
layered structure decreases as the calcination temperature increases. During the
calcination process, the It is confirmed due to excessive high temperature
treatment towards the HTs during the synthesis had cause the layered structure to
be collapse (Anuar et al, 2013; Anuar & Abdullah, 2016).

The high vibration effect to the HT structure due to excess exposure to high
temperature also has resulted in the formation of additional phases. In previous
studies, diffraction peaks 18.5° and 20.5° were reported to be bayerite, which is
Al(OH)3 (Fraile et al, 2009). Bayerite compound is presence in HT-CS 750 where
the diffraction peaks of 18.91° and 20.43°. Low intensity 32.9°, 50.1°, 57.5°, 60.5°
diffraction peaks and high 34.5° diffraction peak which found in HT-CS 750 was
assigned to Mg(OH)2 compounds (Selvam et al, 2011). From HT-CS 750 peaks, it
can be concluded that, at calcination temperature 750 °C, HT has been thermally
decomposed into individual compound which are MgO and Al2O3 due to the
vibrational effect to the HT structure from high temperature exposure (Fraile et al,
2009; Selvam et al, 2011). Therefore, the structure of HT-CS 750 became a single
periclase like structure, known as MgO where the peaks were presented at 36.9°,
43.2° and 62.5° (Anuar & Abdullah, 2016; Rahul et al, 2011; Fraile et al, 2009;

48
Selvam et al, 2011). The most significant periclase like structure is formed on all
HT presented except for HT-CS 550 where a sharp diffraction 43.2° peak is
presence. It did not appear on HT-CS 550 is due to the calcination heat is not as
intense as at 650 °C. The peaks 38.3° and 74.5° were contributed to Al2O3
(Suriyanarayanan et al, 2009). It can be expressed that the formation of additional
phase, known as MgO was due to high calcination temperature. Therefore, another
conclusion can be made is that excessive high temperature calcination also
promotes the decomposition to an individual compound, the deconstruction of
layered structure and the disintegration of HT’s memory effect. It can also be
concluded that sample HT-CS750 and HT-CS850 were no longer a layered double
hydroxide, hydrotalcite.

From Figure 4.1, the use of different agricultural waste as fuel in the
synthesis of HT do affect the structure and crystallinity of HT. Table 4.1 shows the
calorific values of agricultural waste (coconut shell, rice husk and rice straw) and
saccharose which were used as fuel in the synthesis of HT.

Table 4.1: Calorific value of tested fuel

Fuel Type Calorific value (MJ/kg)


Saccharose 15.84
Rice Straw 15.92
Rice Husk 16.51
Coconut Shell 17.71

A bomb calorimetric test was conducted to determine the calorific value of


substances which are coconut shell, rice straw, rice husk and saccharose. It was
found that coconut shell has the highest calorific value followed by rice husk, rice
straw and saccharose respectively. The trend of the rice waste calorific values is
similar as reported by Moni et al (2014) where the values for rice husk and rice
straw were 14.97 MJ/kg and 13.74 MJ/kg respectively. During the synthesis of
catalyst, the fuel acts as a source of C and H where their function is to intensify the
combustion process during calcination (Dávila et al, 2008; Martunus et al, 2011).
In the calcination process, the fuel combusted and produce CO2 and H2O (Helwani
et al, 2013). Those compounds can be absorbed on the metal oxide periclase like
structure. Therefore, the CO2 will be acted as a probe molecule on the surface of
oxide compounds where various of adsorption such as monodentate, bidentate
and bridged taken place to promote the construction of carbonate on the surface

49
or known as layered structure (Dávila et al, 2008). High calorific value provided by
the agricultural waste in this study has resulted more CO2 and H2O production for
HT to be absorbed to form layered structure compared to that of reference fuel
(saccharose). The occurrence might also lead to form large surface area (Anuar &
Abdullah, 2016). With large surface area as proven from BET analysis results on
HTs calcined at 650 °C, it provides more active site which promotes a better
possibility for the catalytic reaction to occur, hence increase the formation of FAME
yield.

4.1.2 BET Analysis of HT

The characteristics of HT was further analyzed by using BET. In this study,


the most desired surface characteristic of catalyst is high surface area. Previous
studies have showed that high surface area shows direct proportional to the
conversion towards biodiesel in transesterification reaction (Carvalho et al, 2018;
Ma et al, 2016). During the transesterification reaction, the surface area will provide
the active site allowing more reaction to occur (Tang et al, 2020). Table 4.2 shows
the surface characteristics of synthesized HT. Due to same similar pattern in terms
of the values of surface by other HTs (HT-SS, HT-RS and HT-RH) with various
temperature calcination (550 °C, 650 °C and 750 °C), therefore, it is not shown.

Table 4.2: Surface characteristics of HT


Sample Surface Area (m2/g)
HT-CS 550 79.595
HT-CS 650 115.558
HT-CS 750 56.837
HT-CS 850 28.320
HT-SS 650 28.326
HT-RS 650 31.872
HT-RH 650 74.472

In the data observed in Table 4.2, the increase of calcination temperature


does affect surface area of HT. From calcination temperature 550 °C to 650 °C,
the surface area of HT increases from 79.595 m2/g to 115.558 m2/g. This pattern
is relatively similar to Helwani et al (2013) and Pinthong et al (2019) where increase

50
of calcination up to 650 °C has increase the surface area of catalyst. It is due to
the crystallite structure began to increase as the calcination temperature increase
(Gaber et al, 2013; Kwon et al, 2020). However, from Table 4.2, it is noted that the
further exposure of calcination temperature 750 °C and above in the synthesis HT
has caused different effect. From HT-CS 650 to HT-CS 750, the surface area
reduced about 58.721 m2/g. Whereas from HT-CS 750 to HT-CS 850 shows a
decayed of surface area of 28.517 m2/g. The reduction of surface area on the
synthesized catalyst is due to calcination temperature at 750 °C and above causes
an overheating to HT which leads to a detrimental effect on the structure of HT
such as surface area (Anuar & Abdullah, 2016). The high calcination temperature
also affects the formation of periclase like structure MgO as proved on XRD spectra
in Figure 4.1. From the table, the most desirable synthesized HT would be HT-CS
650 due to high surface area. Periclase MgO is a single layer structure. The single
layer structure will require more time to have more successful transesterification
reaction. Unlike double layered hydroxide structure. If the periclase and double
layered structure were comparing the conversion to FAME, double layered
structure will have a twice efficient as single layer structure as presented by
published study by Cantrell et al (2005). Therefore, double layered hydroxide
structure is very important towards FAME yield. The HT with high surface is
expected to provide more active site for the transesterification reaction to occur to
provide high catalytic performance which resulted in high FAME yield in production
of biodiesel.

From Table 4.2, it is observed that by using agricultural waste as fuel in the
synthesis of HT, the surface area of HT increases. With the use of coconut shell,
surface area (28.326 m2/g) of HT-SS650 was amplified to 115.558 m2/g which
represented on HT-CS 650. The enlargement of surface area trend is similar to
Sobhana et al (2016) where surface area of HT increases from 36.90 m2/g to 152
m2/g with the use of cellulose as fuel for synthesis of HT. In the combustion method,
the fuel added in the preparation of HT would assist the construction of layered
structure by providing enough energy due to the enthalpy release from the fuel
used. A sufficient source of enthalpy energy would result in good development of
layered double hydroxide (Anuar & Abdullah, 2016). Therefore, due to sufficient
energy provided from the fuel used (coconut shell), it has resulted in large and
good development of surface area framework on HT where calorific value of
coconut shell has shown to be highest than any other fuels as shown in Table 4.1.

51
Therefore, it can be concluded that high calorific value has resulted in high
construction of surface area of HT.

4.1.3 FTIR Analysis of HT

The analysis of HT via FTIR was focused on HT-SS 650, HT-RS 650, HT-
RH 650 and HT-CS 650 due to the successful of HT at calcination temperature 650
°C as observed in Figure 4.1. The element constructing HTs was further verified
through FTIR spectra. Figure 4.2 shows the infrared spectra for HT-SS 650, HT-
RS 650, HT-RH 650 and HT-CS 650.

(a) (i) (ii) (iii) (iv) (v) (vi)

(b)
Intensity (a.u.)

(c)

(d)

4000 3500 3000 2500 2000 1500 1000 500


Wavelength (cm-1)

Figure 4.2: FTIR spectra of (a) HT-SS 650, (b) HT-RS 650, (c) HT-RH 650, (d) HT-
CS 650, (i) 3400 – 3500 cm-1, (ii) 1634 – 1643 cm-1, (iii) 1375 – 1400 cm-1, (iv) 1042
– 1092 cm-1, (v) 810 – 825 cm-1 and (vi) 575 – 630 cm-1

Based on Figure 4.2, the constructing element in HT was verified via the
existence peak of 3400 – 3500 cm-1, 1634 – 1643 cm-1, 1375 – 1400 cm-1, 1042 –
1092 cm-1, 810 – 825 cm-1 and 575 – 630 cm-1 as per reported by previous
researchers (Shabanian et al, 2020; Yang et al, 2020).The existence of broad
absorption band on 3400 – 3500 cm-1 indicated the existence of hydroxyl group in
the interlayer of the brucite-like layers (Coelho et al, 2017). The presence of H2O
in the interlayer of the HT structure is confirmed through the existence of 1634 –
1643 cm-1 band (Yang et al, 2020). Strong peak at 1350 – 1400 cm-1 represents

52
carbonate band in where it is located on the interlayer gallery and act as bridge
between cations, hence forming the layered structure (Dixit et al, 2013). The
presence of weak 1042 – 1092 cm-1 band and 800 – 825 cm-1 band conforms the
existence of covalent carbonate bond is due to high calcination temperature (Anuar
& Abdullah, 2016). As the temperature increases, some of LDH was collapsed,
resulting in trapped of CO3- in the brucite-like structure (Anuar & Abdullah, 2016).
On the other hand, the band on 575 – 630 cm-1 responds to the vibration of metal
oxides which are Mg-O and Al-O (Shabanian et al, 2020). Therefore, it can be
concluded that the presence of peak 3400 – 3500 cm-1, 1634 – 1643 cm-1, 1375 –
1400 cm-1, 1042 – 1092 cm-1, 810 – 825 cm-1 and 575 – 630 cm-1 verified the
successful element constructed of HT.

Above all findings, the fuels used during the synthesis of HT displayed an
insignificant effect towards the formation of peaks in the spectra of FTIR thus
confirming no peak was appeared after the synthesis of LDH as claimed by Coelho
et al (2017). It is due to all fuels were combusted and all were degraded at above
500 °C. On the other hand, the fuel does increase the intensity and bonding
characteristics of LDH. However, the fuel does affect the intensity of the band
peaks. The most noticeable peak which is at 1375 – 1400 cm-1 shows that HT-CS
has better intensity compared to other HTs. The intensity of carbonate band at
peak 1375 – 1400 cm-1 increases from 8.81% to 22.88% as compared to the
reference catalyst HT-SS. Another noticeable increase of peak intensity is hydroxyl
group band (3400 – 3500 cm-1) where the intensity increases from 3.42% to
11.73% in comparison with reference HT-SS. As mentioned earlier, due to high
calorific value of fuel added in the synthesis of HT, the fuel has provided a large
amount of C and H atoms and sufficient energy to the catalyst in which the catalyst
is absorbing more C and H source to develop a good layered structure. In addition,
the fuel also did alter and improve the surface area of HT depending on the calorific
value of the fuel used in the synthesis of HT as mentioned in BET analysis as
presented in Table 4.2.

53
4.1.4 TGA Analysis of HT

Thermogravimetric analysis was conducted to study the thermal


decomposition of freshly prepared dried HT samples ranging from 30 °C – 800 °C.
The intention of TGA analysis is to determine organic compound that has been
formed in the HTs via thermal decomposition of the organic in the sample HT-SS
650, HT-RS 650, HT-RH 650 and HT-CS 650. Figure 4.3 show the TGA profiles of
HT samples.

100 0.1(a)

Weight Derivative
0
80
Weight (%)

-0.1

(1/°C)
-0.2
60
-0.3
40 -0.4
-0.5
20 -0.6
30 130 230 330 430 530 630 730
Temperature (°C)

Figure 4.3 (a): TGA profile of HT-SS 650

100 0.1(b)
90
Derivative Weight

0
80
Weight (%)

-0.1
70
(1/°C)

-0.2
60
-0.3
50
40 -0.4
30 -0.5
20 -0.6
30 130 230 330 430 530 630 730
Temperature (°C)
Figure 4.3 (b): TGA profile of HT-RS 650

54
100 0.1(c)

Derivative Weight
0
80
Weight (%)
-0.1

(1/°C)
-0.2
60
-0.3
40 -0.4
-0.5
20 -0.6
30 130 230 330 430 530 630 730
Temperature (°C)

Figure 4.3 (c): TGA profile of HT-RH 650

100 0.1
(d)

Derivative Weight
0
80
-0.1
Weight (%)

60

(1/°C)
-0.2
40 -0.3
-0.4
20
-0.5
0 -0.6
30 130 230 330 430 530 630 730
Temperature (°C)

Figure 4.3 (d): TGA profile of HT-CS 650

It was found that all samples show a significant weight loss. The HT-SS
650 losses 38.84% of total weight. Followed by HT-RS 650 loses 57.56%, HT-RH
650 loses 55.98% and HT-CS 650 losses 79.67% of total weight. It shows that HT-
CS 650 has the highest organic matter such as OH and CO3 which constructed the
layered structure whereas HT-SS 650 has the least layered structure constructed
which was proved by the large surface area constructed on HT-CS 650 as
presented on Table 4.2 and having highest intensity of OH and CO3 band in Figure
4.2. From the profiles, there are 3 major stages of decomposition can be
characterized. The first stage of decomposition occurred from 37 °C - 250 °C where
physical-absorbed H2O molecules was removed known as dehydration step (Dixit
et al, 2013; Ramos-Ramírez et al, 2018). Saccharose fueled HT showed the least
weight loss of 12.89% on the first stage of thermal decomposition. Whereas
grounded coconut shell fueled HT showed the highest weight loss which is 52.54%.

55
Therefore, the chemical composition for hydrotalcite in the first stage of
decomposition is as follows:

𝑀𝑔6 𝐴𝑙2 (𝑂𝐻)16 𝐶𝑂3 ∙ 4𝐻2 𝑂 → 𝑀𝑔6 𝐴𝑙2 (𝑂𝐻)16 𝐶𝑂3 + 4𝐻2 𝑂 at 37 – 250 °C (eq 4.1)

In second stage of decomposition (251 °C – 500 °C), the hydroxyl groups


and interlayer carbonate ions such as CO3- and OH- ion began decomposed (Dixit
et al, 2013; Reyero et al, 2013). The second stage is also known as
dihydroxylation/decarbonation step. It is observed that HT-CS decayed about
14.57% weight followed by HT-RH, HT-RS and HT-SS which decomposes
13.38%, 12.70% and 11.20% of weight. HT-SS loses the least weight percent
which also proved to be having the least interlayer compound as proved with the
least intensity in FTIR in Figure 4.2. HT-CS loses more weight from other HTs due
to having the highest interlayer compound that constructed the layered structure
as proved to have the highest intensity as displayed in Figure 4.2. This result also
proves that fuel used with increase of calorific value has resulted more C and H
atoms supplied to the HT to construct layered structure. Therefore, the higher the
calorific value of fuel, has resulted in more C and H atoms will be formed as
interlayer in HT structure. Therefore, the chemical composition for HT in the second
stage of decomposition is as follows:

𝑀𝑔6 𝐴𝑙2 (𝑂𝐻)16 𝐶𝑂3 → 𝑀𝑔6 𝐴𝑙2 𝑂8 (𝑂𝐻)2 + 7𝐻2 𝑂 + 𝐶𝑂2 at 251 – 500 °C (eq 4.2)

The final stage of thermal decomposition started at 500 °C where the


structure of layered hydroxide is collapse in LDH framework (Anuar & Abdullah,
2016; Ramos-Ramírez et al, 2013; Reyero et al, 2013). The remaining OH- were
collapse, converted into vapor (Lin et al, 2005). As a result, it leads to the formation
of MgO and Al2O3 (Rezvani et al, 2013). The disappearance in diffraction 11.7°
proved the layered structure has collapse.

𝑀𝑔6 𝐴𝑙2 𝑂8 (𝑂𝐻)2 → 6𝑀𝑔𝑂 + 𝐴𝑙2 𝑂3 + 𝐻2 𝑂 at 500 – 800 °C (eq 4.3)

56
4.2 Catalyst Performance
4.2.1 Effect of Fuel Type

The study was initiated with HT synthesized from 4 different types of fuel
which are coconut shell, rice straw, rice husk and saccharose as reference fuel.
The performance of HT in catalyzing the transesterification reaction process was
first begin with the selection of the best fuel. The same calcination temperature
(650 °C) was used to synthesize the catalyst. At this stage, the effect of fuel in HT
synthesization was investigated towards HT catalytic performance in
transesterification reaction of biodiesel production. Each fuel has different amount
of calorific values which can be referred on Table 4.1. Fuel is crucial to enhance
the development of LDH structure during calcination (Anuar & Abdullah, 2016;
Coelho et al, 2017). In FTIR spectra, the difference in structure of HT synthesized
by those fuels was hardly to be observed due to having similar functional groups.
Only the intensity of OH and CO3 within the HT in FTIR are observable. Similar to
TGA analysis, where the decomposition of organic compound in HT such as OH
and CO3 are observable where HT-CS having the highest weight loss. The intensity
of OH and CO3 could lead the difference in catalytic performance of HT. The
performance of the catalytic activity of those HTs are very clearly significant as well
as the spectra of XRD as shown in Figure 4.1. Figure 4.4 shows FAME yield from
different types of fuel used during calcination of HT. Due to same performance in
terms of yield by other HTs with various temperature calcination (550 °C, 650 °C
and 750 °C), therefore, it is not shown.

100 92.86
90 82.94
80 73.88 76.60
FAME Yield (%)

70
60
50
40
30
20
10
0
HT-SS 650 HT-RS 650 HT-RH 650 HT-CS 650
Types of HT

Figure 4.4: Catalytic activity of HT with different types of fuel

57
From the Figure, it is observed that HT-CS has the highest FAME yield
which is 92.86%. Whereas 82.94%, 76.60% and 73.88% are FAME yield for HT
utilized rice husk, rice straw and saccharose as fuel. The result of HT produced by
saccharose are very close to the findings of Helwani et al (2013), Anuar & Abdullah
(2016) which are 71% and 70.67% under stirring condition for 5 hours. In
comparison with the obtained experimentation values with the values mentioned
from literatures where the values are very close has proved the outcome
consistency of HT towards biodiesel production yield. From Figure 4.4, it is
believed that with the change of fuel during synthesis of catalyst, the FAME yield
could rose depending on types of organic fuel used. From the Figure, it can be
seen that the FAME yield is directly proportional to the calorific value and the
surface area of synthesized HT. HT-CS 650 which utilized coconut shell (17.71
MJ/kg) as fuel during the synthesis is having the highest surface area (115.558
m2/g) which successfully having the highest FAME yield of 92.86%. Whereas the
reference HT (HT-SS 650) utilized saccharose (15.84 MJ/kg) during the synthesis
of HT is having the lowest surface area (28.326 m2/g) the lowest FAME yield of
73.88%. The result is somewhat close to the study conducted by Anuar & Abdullah
(2016) where the researchers utilized glucose, fructose and saccharose as fuel in
the synthesis of HT. The researchers discovered that the fuel with highest calorific
value which is saccharose as compared to fructose and glucose which having the
highest catalytic performance FAME yield of 76.45% (Anuar & Abdullah, 2016). It
can be concluded that the performance of HT is directly proportional to surface
area of HT and calorific value of fuel used during the synthesis of HT as proved in
Table 4.1, Table 4.2 and Figure 4.4. The high percentage of OH and CO3 in the
HT also lead to the better catalytic performance of HT. It is due to the OH and CO3
are basic property, which favors the catalytic performance in transesterification
reaction.

4.2.2 Effect of Calcination Temperature

To determine whether calcination temperature affected the catalytic activity


of the HT, transesterification of WCO using HT calcined at different temperatures
was performed. Previous studies show the optimum performance for calcination
temperature of HT is in range between 500 °C - 800 °C (Coelho et al 2017, Anuar
& Abdullah, 2016; Helwani et al, 2013; Reyero et al, 2013). Therefore, the

58
calcination temperature used in this present study were 550 °C, 650 °C and 750
°C. In the transesterification reaction with methanol, the WCO produces FAME
which composing methyl myristate, methyl palmitate, methyl stearate, methyl
oleate and methyl linoleate. Figure 4.5 presents catalytic activity of HT with
different calcination temperature.

100 92.86
90
80
FAME Yield (%)

70 63.83
60 53.91
50
40
30
20
10
0
HT-CS 550 HT-CS 650 HT-CS 750
Types of HT-CS

Figure 4.5: Catalytic activity of HT- CS with different calcination temperature

In this study, it was found that the HT calcined at 650 °C shows the highest
production of FAME yield in the transesterification reaction which is 92.86%. It can
also be observed that HT-CS calcined at 650 °C showed the highest FAME yield
produced from the catalytic reaction in 5 hours reaction time followed by HT at 550
°C and at 750 °C. The increase of FAME yield with the use of HT-CS 550 and HT-
CS 650 has proven that calcination affects the FAME yield. Referring to Table 4.2
and Figure 4.5, from the calcination temperature of 550 °C to 650 °C, it can be
noticed that the FAME yield increases directly proportional to the increased surface
area of BET. HT-CS 550 provides 79.595 m2/g which leads to 63.83% of FAME
yield. With the increase of calcination temperature to 650 °C, an extra of 35.963
m2/g which resulted in increase of FAME yield. Data from Table 4.2 and Figure 4.5
enhanced that the desired HT would be the highest surface area which is HT-CS
650. As mentioned, it is due to higher surface area provided by the catalyst which
more active site is introduced to the reaction, leading to high FAME yield. The high
yield of FAME produced from the transesterification by HT-CS 650 is due to having
highest surface area displayed in Table BET analysis, highest hydroxyl and

59
carbonate intensity displayed in FTIR result and highest intensity of interlayer
hydroxyl and carbonate displayed in TGA result. The high surface area of HT-CS
650 provided in transesterification reaction has resulted in providing more active
site to allow more reaction to occur. High intensity of hydroxyl and carbonate
displayed in FTIR and TGA result has contributed to the basicity of HT where HT-
CS 650 has successfully created a high basic condition where basic condition is
highly required for transesterification reaction to conduct efficiently (Anuar &
Abdullah, 2016, Trejo-Zárraga et al, 2018). This is somewhat similar with Anuar &
Abdullah (2016) in which the highest calcination temperature led to the highest
surface area of catalyst. However, at 750 °C, the FAME yield drops drastically in
which is 53.91%. It is due to structure collapse which was proved and can be
observed on Figure 4.1 where the layered structure did not appear on HT-CS750.
Consequently, it has resulted to the surface area of the catalyst to be reduced
significantly which has led to reduction of catalytic activity. In combination of
obtained data with Table 4.1, Figure 4.2 and Figure 4.5, it can be summed up that
as the calcination temperature exceed optimum calcination temperature, the
layered structure started to collapse and can no longer be recrystallized. The HT-
CS 750 was no longer a layered structure; hence it consists of MgO and Al2O3 as
depicted on the XRD in Figure 4.1. There are few active sites in HT-750, but it was
not as much as HT-CS 650. The FAME yield obtained from transesterification
follows the trend of BET surface area. From the discussion, it can be concluded
that HT-CS 650 was the best HT synthesized in this study and therefore was
proceeded on the reusability study.

4.2.3 Reusability of HT

The study of catalyst reusability is included in the scope of study due to one
of advantage of using heterogeneous catalyst in transesterification reaction. In this
study, the best catalyst which is HT- CS 650 were used for the reusability test. After
each experiment, the catalyst was centrifuged out on the product mixture. The
catalyst was recollected via filtration and washed with n-hexane to remove the
remaining glycerol. The catalyst was filtered and recalcined at 300 °C for 3 h in
order to remove any remaining glycerol that adhered on the surface of the catalyst.
According to Alsamad et al (2018), biodiesel and glycerol decomposed at
temperature 260 °C and 235 °C which recalcination temperature is sufficient. After

60
that, the calcined catalyst was recrystalized and rehydrated with sodium carbonate
solution and dried before use in the next experimental cycle. The catalyst was
repeated up to 3 cycles after the use of fresh batch. Figure 4.6 shows the
performance of catalytic activity for HT-CS 650.

HT-CS 650
100 92.86 90.54 87.84
90 83.77
80
FAME Yield (%)

70
60
50
40
30
20
10
0
Fresh 1st Cycle 2nd Cycle 3rd Cycle
No. of Cycles

Figure 4.6: Reusability cycle in performance of catalytic activity for HT-CS 650
under 5-hour reaction, 15:1 methanol to oil molar ratio, 8 wt% catalyst loading at
65 °C

From the experimental run, it was found that the highest FAME yield
produced by the best active HT which is HT-CS 650 has indicated a slight reduction
towards every cycle. In the fresh batch of HT-CS 650, the FAME yield was 92.86%
respectively. However, on the third cycle, the catalyst performance of HT-CS 650
dropped of 9.09%. The FAME yield for HT-CS 650 at third cycle is 83.77%
reciprocally. The reduction of FAME yield for the best active HT from every cycle
of transesterification reaction is due to minor structure collapse of catalyst. The
collapse occurred due to repeated calcination steps (Anuar & Abdullah, 2016).
Consequently, the carbonate content in the catalyst diminishes during the
repetitive re-calcination process. The slight reduction in FAME yield in
transesterification reaction for HT-CS 650 shows that the HT-CS 650 catalyst has
a better robustness and potentially could undergo several cycles of
transesterification reaction with a consistent FAME yield. The catalytic activity of
HT-CS 650 for the first cycle and onwards was able to be restored via re-calcination
and re-hydration, then the FAME yield is relatively able to be maintain throughout

61
the transesterification reaction repetitive cycles due to its unique property, known
as memory effect (Anuar & Abdullah, 2016; Helwani et al, 2013). However, the
experiment could not be further down to the 3rd cycle due to the loss of catalyst on
the recovery process. Only 60% of the catalyst was recovered after each
transesterification process. Hence, the amount of catalyst was too small to proceed
for the 4th cycle and the reusability experimental was stopped. Throughout the
experimental run, it is proved that HT has reusable for at least 3 cycles with the
minimum reduction of FAME yield of only 9.09% for HT-CS 650.

4.3 FTIR Studies on Biodiesel

The analysis of FTIR towards biodiesel and WCO is to determine the


changes in FTIR band and to verify that FAME or biodiesel is produced. The WCO
was reacted with methanol in a reactor for 5 hours at 65 °C. The catalytic
transesterification reaction of oil with methanol produced FAME and glycerol as
byproduct. The FAME produced from transesterification reaction can be identified
via FTIR spectra. If the feedstock (WCO) and product (FAME) has the same FTIR
spectra peaks, it can significantly prove that there are no reaction occurs to the
reactants. Figure 4.7 shows FTIR spectra of WCO and biodiesels synthesized
using HT as catalyst. Some spectra peaks of WCO and biodiesel are appearing
at the same region due to similar structure such as fatty acid chain, carbonyl C=O
group, C-O group, CH2 and CH3 component. The appearance of peak 3010 cm-1
represents the existence of (C=C) group which exist in oleic acid as triglyceride
and FAME (Peña et al, 2014). Existence of 1742 cm-1 on all peaks represent
carbonyl C=O groups which exist in triglyceride and FAME (Nisar et al, 2017). 2800
– 2995 cm-1 represents the methylene group in which the group exist in the fatty
acid chain in the chemical structure of FAME and triglyceride (Peña et al, 2014).
The existence of similar functional group shows that the FAME and triglyceride is
almost chemically similar. Only few structural compounds which can differentiate
between WCO and FAME which are triglyceride compound and ester compound.

62
(i) (ii) (iii) (iv)
Intensity (a.u.)

1600 1400 1200 1000 800 600


Transmittance (cm-1)
Figure 4.7: FTIR Spectra of (a) WCO (b) Biodiesel produced using HT-SS 650, (c)
HT-CS 650, (d) HT-RH 650, (e) HT-RS 650, (i) 1412-1474 cm-1, (ii) 1342-1386 cm-
1
, (iii) 1152-1220 cm-1 and (iv) 1002-1038 cm-1.

Several significant peaks were found in which can be used as a verification


to differentiate between WCO and biodiesel. The most noticeable peaks were
spotted on range 1196 cm-1 where the peak represent the formation of component
methyl ester (CO)-C-CH3 which represent the FAME (Huaping et al, 2006).
Formation of ethyl ester on range 1435 cm-1 also considered to be most significant
peak as it is still considered as FAME classification (Marwan et al, 2015). Peak on
1377 cm-1 representing triglyceride compound vibration as its presence in WCO
spectra. However, this peak (1377 cm-1) disappear in both biodiesel spectra. It
shows that the effective conversion of triglyceride was successfully converted into
FAME after transesterification reaction as per similar findings with Rabelo et al
(2015). The existence peak on 1002-1038 cm-1 is attributed to C-O vibration on
biodiesel peaks is as similar finding with Nisar et al (2017) which also contributed
the success conversion of triglyceride to biodiesel. Therefore, it can be concluded
that the absence of peak 1377 cm-1 and the presence of peak 1196 cm-1, 1436 cm-
1
and 1014 cm-1 verified the successful catalytic reaction and the conversion of
WCO to palm oil FAME.

63
4.4 Kinetic Studies of Biodiesel Synthesis
4.4.1 Effect of Reaction Temperature and Time on Transesterification
Reaction

The best catalyst with reference catalyst i.e HT-CS 650 and HT-SS 650
were subjected in the kinetic study of transesterification reaction. In kinetic study,
reaction temperature is one of important parameter in transesterification reaction.
The catalytic transesterification reaction was done in various reaction temperatures
i.e. 55 °C, 60 °C, 65 °C and 70 °C to determine the effect of reaction temperature
towards transesterification reaction. Figure 4.8 shows the catalytic activity of HT-
CS 650 and HT-SS 650 under 5-hour reaction time, 15:1 methanol to oil molar
ratio, 8 wt% catalyst loading under reaction temperature varied from 55 °C – 70
°C. The basis of selection for the reaction conditions were highlighted in Chapter
2, Subchapter 2.8; the concluding remarks where the parameters were based on
literature review.

100 55°C 60°C 65°C 70°C (b)


(a)
FAME Yield (%)

80

60

40

20

0
0 1 2 3 4 5
Time (Hour)

100 55°C 60°C 65°C 70°C (c)


FAME Yield (%)

80

60

40

20

0
0 1 2 3 4 5
Time (Hour)
Figure 4.8: Catalytic activity of (a)HT-SS 650 and (b) HT-CS 650 under 5-hour
reaction time, 15:1 methanol to oil molar ratio, 8 wt% catalyst loading with various
reaction temperature.

64
Based on Figure 4.8, the FAME yield increases along the 5-hour reaction.
Therefore, in comparison with reaction temperature from 55 – 70 °C, it was found
that the ideal temperature for 5-hour transesterification reaction is at 65 °C for all
HTs. HT-CS 650 and HT-SS 650 produce high FAME yield at ideal reaction
temperature at 65 °C which are 93.05% and 73.83% respectively. From
temperature 55 °C to 65 °C, it can be noticed that the FAME yield increases with
the increase in reaction temperature. In parallel to previous research where higher
reaction temperature resulted in higher FAME yield as more of energy is provided
and supplied during the catalytic reaction (Ali et al, 2019; Birla et al, 2012; Kaur &
Ali, 2014). However, from reaction temperature 65 °C to 70 °C, the FAME yield
dropped significantly. It is due to the reaction temperature of 70 °C has exceeded
the boiling temperature of methanol which is at 64.7 °C. With that reason of
exceeding methanol boiling temperature, the methanol vaporizes which reducing
the reactant and favouring the glycerol formation (Tan et al, 2015; Yusuff et al,
2013).

From Figure 4.8, it was found that all catalysts are very active on the first
hour of transesterification reaction. For HT-CS 650, the FAME yield for the first
hour of transesterification reaction at 65 °C is 68.71%. Whereas for HT-SS 650,
the FAME yield for the first hour of transesterification reaction at 65 °C is 38.27%.
The increase FAME yield for all catalysts on the next consecutive hour of
transesterification reaction at 65°C was ranging from 0.88 to 14.71%. However,
the reduction of FAME yield on the second hour one to the next consecutive hours
were due to the catalyst was in a prolong contact with the reactants, in which
triglyceride was clogged onto the surface of HT. With 5-hour reaction time, it
caused a significant adsorption of triglyceride to the pore surface which eventually
blocking the active sites of HT (Atadashi et al, 2012). The blockage was deposited
from WCO on the active site of HT has been suspected to have slowed the reaction
to convert triglyceride to FAME. Therefore, for future or upcoming study, 2-step or
multi-step transesterification reaction is recommended to increase the FAME yield
in short duration of time.

65
4.4.2 Transesterification Reaction Kinetics

To obtain the activation energy for HT-CS 650 and HT-SS 650, the reaction
rate constant, k values were determined from the slope value of the Equation 2.8
in Section 2.5. The FAME values from plotting ln(1-FAME) versus reaction time, t
are FAME yields obtained from Figure 4.8. The graph ln(1-FAME) versus reaction
time, t can be observed in Appendix I. After obtaining ln k value with respected to
the reaction temperature, ln k versus 1/T was plotted to obtain activation energy
values. Figure 4.9 shows ln k versus 1/T plot. The activation energy values for HT-
CS 650 and HT-SS 650 can be obtain from the slope of the straight best line on
Figure 4.9. Table 4.3 shows the activation energy values of HT-CS 650 and HT-
SS 650.

HT-SS 650 HT-CS 650


1/T (1/k)
0.00294 0.00296 0.00298 0.003 0.00302 0.00304 0.00306
0
-0.5
-1 y = -29394x + 86.205
-1.5
-2
ln k

-2.5
-3
-3.5 y = -35559x + 103.83
-4
-4.5
-5
Figure 4.9: Arrhenius plot ln k versus 1/T for transesterification reaction for various
HTs (HT-SS 650, HT-RH 650 and HT-CS 650.

Table 4.3 List of activation energy values for synthesized HTs


Catalyst Activation Energy, Ea (kJ/mol) R2
HT-SS 650 295.64 0.999
HT-CS 650 246.68 0.995

As determined from Figure 4.9, the activation energy for HT-CS 650 and
HT-SS 650 are 246.68 kJ/mol and 295 kJ/mol, respectively. All the activation
values obtained in Table 4.3 fall under the range from 40 – 400kJ/mol as mentioned

66
by Xin et al (2009). Although, the activation energy values obtained in the study is
quite high by referring to Table 2.4 in section 2.5, the activation energy in this study
was much lower than the previous reported by Dhawan et al (2018) where the
activation energy for HT in transesterification reaction was found out to be 380
kJ/mol. In addition, HT-CS 650 has shown better activation energy compared to
the reference HT, HT-SS 650. With direct comparison, the synthesized HT was
found to have lower activation energy, which could lead to better successful
reaction compared to HT synthesized by Dhawin et al (2018).

High activation energy value also indicates the degree of temperature -


sensitiveness. With variation of 5 °C in reaction temperature for transesterification,
the FAME yield deviates. As mentioned, at temperature 55 °C, the performance of
HT-CS 650 and HT-SS 650 were very poor which had FAME yield lower than 15%.
At 65 °C, the synthesized HT has the best performance where highest
FAME yield was obtained. Nevertheless, at 70 °C, the performance of HT drops
due to exceeding the methanol boiling temperature resulting less reaction to occur.
Based on the results obtained, it was found that temperature reaction was proven
to be the positive factor for transesterification reaction.

From Table 4.3, it was observed that transesterification reaction using HT-
CS 650 has the lowest activation energy as compared to that of HT-SS 650. The
low activation of energy for the reaction has resulted in highest FAME yield. As the
reaction the HT-CS 650 has the lowest energy compared to other synthesized HT,
more transesterification reaction occurred with the use of HT-CS 650 as catalyst.
The contribution of high catalytic activity of HT-CS 650 is due to high surface area
constructed by the highest calorific value of fuel, which is coconut shell (17.71
MJ/kg) whereas calorific values for saccharose were 15.84 MJ/kg respectively
which can be referred on Table 4.1. As mentioned earlier, in the calcination
process, the fuel with high calorific value combusted and produce CO2 and H2O
which then had taken place to promote more construction of carbonate on the
surface or layered structure (Dávila et al, 2008). High surface area and good
layered structure development of the HT as proved in Figure 4.1 and Table 4.2 has
led to more active site allowing more reactions to occur. Another contribution to
high catalytic activity of HT-CS is due to having high OH and CO3 content in HT-
CS 650 as displayed in TGA analysis in Figure 4.3 which leads to high basicity that
enhances the catalytic activity of transesterification reaction. As transesterification

67
reaction begins, the large amount of OH ion in the HT-CS 650 assisted the alkoxide
ion from alcohol to attack the carbonyl carbon of triglyceride producing
intermediates which would later form FAME (Andrijanto, 2012). The activation
energy for HT-CS 650 is 246.68 kJ/mol as comparable to reference HT, HT-SS
650 which is 295.64 kJ/mol. The pattern of obtained activation energy follows the
trend of calorific value of fuel used for the synthesis of HT. Therefore, it can also
be added that the fuel does affect the reduction of activation energy for
transesterification reaction. In which, it is simultaneously boost the performance of
HT which utilized the highest calorific value fuel.

68
CHAPTER 5

CONCLUSION

5.1 CONCLUSION

Hydrotalcite was successfully synthesized via combustion method with the use
of proposed fuel (coconut shell, rice straw and rice husk) in comparison with reference
fuel (saccharose) at temperature 550 °C and 650 °C. The HT structure development
synthesis was verified through XRD where diffraction of peak on 11° and 23° exist as
indicated as a layered structure. Exposure to higher temperature (>750 °C) led to
layered structure dehydroxylated and collapsed and consequently resulting the
formation of separated metal oxide such as Al2O3 and MgO. Elements constructed in
HT at 650 °C also verified through FTIR with existence of peaks at 3400 – 3500 cm-1,
1634 – 1643 cm-1, 1375 – 1400 cm-1, 1042 – 1092 cm-1, 810 – 825 cm-1 and 575 – 630
cm-1 verified the successful element constructed of HT.

The desired HT for the study is highest surface area on the HT which could
lead to high FAME yield in transesterification. From various calcination temperature
(550 °C – 750 °C), the calcination temperature at 650 °C shows the highest surface
area construction. The catalyst HT CS 650 was found to be having the highest surface
area followed by HT RH 650, HT RS 650 and HT SS 650 due to having highest calorific
value. The trend of surface area formed follows the calorific value of fuel used during
the synthesis of catalyst. Highest calorific value represent highest source of C and H
provided to the HT during in synthesis of HT to construct layered structure. It can be
concluded that high calorific value of fuel used in the synthesis could leads to high
surface area formed on the catalyst. The increase of temperature from 550 °C to 650
°C could leads to high formation of surface area. However, overexposure to the heat
such as calcination temperature at 750 °C and above could reduce the surface area
of HT.

From the study, the use of high calorific value as fuel in the synthesis of HT
(coconut shell), HT-CS 650 has the highest FAME yield in transesterification reaction
at 65 °C which is 92.86%. The FAME yield for HT-SS 650, HT-RH 650 and HT-RS 650

69
are 82.94%, 76.60% and 73.88% respectively. It was also found out that HT-CS
catalyst was calcined at 650 °C indicated the highest FAME yield compared to that of
550 °C and 750 °C. The FAME yield for HT-CS 550 and HT-CS 750 are 63.83% and
53.91% respectively. In also can be concluded that, the FAME yield obtained from
transesterification reaction in the study is directly proportional to the surface area of
HT and the intensity of OH and CO3 ion in HT. In addition, the robustness of HT-CS
650 was tested where it can be reused up to 3 cycles with minimum reduction 9.09%
FAME yield.

In kinetic study, all HTs synthesized via combustion method has shown
excellent ideal catalytic activity for transesterification of WCO at 65 °C reaction
temperature. Increase of reaction temperature from 55 °C to 65 °C leads to increase
of FAME yield. Increase of reaction temperature to more than 70 °C resulted poor
FAME yield. The FAME yield increases directly proportional to the reaction time in
transesterification reaction. The HT-CS 650 shows the lowest activation energy value
(246.68 kJ/mol) compared to reference HT, HT-SS 650 which has resulted in the
highest FAME yield. The kinetic data can also be used to optimize the production of
biodiesel.

5.2 RECOMMENDATION

There are several recommendations could be suggested. From the study, it


was found that agricultural waste such as coconut shell is compatible as an alternative
to saccharose due to have high performance in transesterification reaction. Therefore,
during the synthesis of catalyst, effect of fuel amount ranging 0 – 20 wt% is
recommended for future study to understand towards the structure and the
performance in production of biodiesel. From previous study, it is proven HT could
resist low quality feedstock containing water up to 1 wt% without affecting the yield of
biodiesel (Atadashi et al, 2012). To ensure the resistibility of HT, effect of water content
in WCO feedstock ranging from 0 – 10 wt% of water content feedstock (WCO) should
be conducted for further study. As mentioned in study where the performance of HTs
reduced in terms of FAME yield from one to the next consecutive hours was due to the
catalyst was in a prolong contact with the reactants where it is suspected the reactant
or triglyceride blocking the active sites of HT. Therefore, for future or upcoming study,
2-step or multi-step transesterification reaction is recommended to increase the FAME
yield in short duration of time.

70
REFERENCE

Abbaszaadeh, A., Ghobadian, B., Omidkhah, M. R., & Najafi, G. (2012). Current
biodiesel production technologies: A comparative review. Energy Conversion
and Management, 63, 138-148.

Abou-Elyazed, A. S., Ye, G., Sun, Y., & El-Nahas, A. M. (2019). A Series of UiO-66
(Zr)-Structured Materials with Defects as Heterogeneous Catalysts for
Biodiesel Production. Industrial & Engineering Chemistry Research, 58(48),
21961-21971.

Agarwal, M., Chauhan, G., Chaurasia, S. P., & Singh, K. (2012). Study of catalytic
behavior of KOH as homogeneous and heterogeneous catalyst for biodiesel
production. Journal of the Taiwan Institute of Chemical Engineers, 43(1), 89-
94.

Aguieiras, É. C., Cavalcanti, E. D., da Silva, P. R., Soares, V. F., Fernandez-Lafuente,


R., Assunção, C. L. B., da Silva, J. A. C. & Freire, D. M. (2020). Enzymatic
synthesis of neopentyl glycol-bases biolubricants using biodiesel from soybean
and castor bean as raw materials. Renewable Energy, 148, 689-696.

Ahmadipour, M., Venkateswara Rao, K., & Rajendar, V. (2012). Formation of


nanoscale structure by solution combustion: effect of fuel to oxidizer
ratio. Journal of Nanomaterials, 2012.

Aho, A., Kumar, N., Eränen, K., Salmi, T., Hupa, M., & Murzin, D. Y. (2008). Catalytic
pyrolysis of woody biomass in a fluidized bed reactor: influence of the zeolite
structure. Fuel, 87(12), 2493-2501.

Ajala, O. E., Aberuagba, F., Odetoye, T. E., & Ajala, A. M. (2015). Biodiesel:
Sustainable Energy Replacement to Petroleum‐Based Diesel Fuel–A
Review. ChemBioEng Reviews, 2(3), 145-156.

Alexander, C., & Hurt, C. (2007). Biofuels and their impact on food prices. Purdue
University Cooperative Extension Service.

71
Allah, F. U. M., & Alexandru, G. (2016, August). Waste cooking oil as source for
renewable fuel in Romania. In IOP Conference Series: Materials Science and
Engineering, 147(1),012133

Ali, B., Yusup, S., Quitain, A. T., Bokhari, A., Kida, T., & Chuah, L. F. (2019).
Heterogeneous Catalytic Conversion of Rapeseed Oil to Methyl Esters:
Optimization and Kinetic Study. In Advances in Feedstock Conversion
Technologies for Alternative Fuels and Bioproducts (pp. 221-238). Woodhead
Publishing.

Alsamad, T., Almazrouei, M., Hussain, M. N., & Janajreh, I. (2018). Modeling of
Thermochemical Conversion of Glycerol: Pyrolysis and H2O and CO2
Gasification. Waste and Biomass Valorization, 9(12), 2361-2371.

Amani, H., Asif, M., & Hameed, B. H. (2016). Transesterification of waste cooking palm
oil and palm oil to fatty acid methyl ester using cesium-modified silica
catalyst. Journal of the Taiwan Institute of Chemical Engineers, 58, 226-234.

Andrijanto, E. (2012). The study of heterogeneous catalysts for biodiesel


synthesis (Doctoral dissertation, University of Huddersfield).

Antunes, W. M., de Oliveira Veloso, C., & Henriques, C. A. (2008). Transesterification


of soybean oil with methanol catalyzed by basic solids. Catalysis Today, 133,
548-554.

Anuar, M. R., & Abdullah, A. Z. (2016). Ultrasound-assisted biodiesel production from


waste cooking oil using hydrotalcite prepared by combustion method as
catalyst. Applied Catalysis A: General, 514, 214-223.

Anuar, M. R., Abdullah, A. Z., & Othman, M. R. (2013). Etherification of glycerol to


polyglycerols over hydrotalcite catalyst prepared using a combustion
method. Catalysis Communications, 32, 67-70.

Anya, A. U., Ugwumma, C., Chioma, A. I. N., & Okoro, P. (2013). Effect of Free Fatty
Acid Content on the Yield of Biodiesel Derived from Neem Oil.

72
Atadashi, I. M., Aroua, M. K., Aziz, A. A., & Sulaiman, N. M. N. (2012). The effects of
water on biodiesel production and refining technologies: A review. Renewable
and sustainable energy reviews, 16(5), 3456-3470.

Babel, S., & Kurniawan, T. A. (2004). Cr (VI) removal from synthetic wastewater using
coconut shell charcoal and commercial activated carbon modified with
oxidizing agents and/or chitosan. Chemosphere, 54(7), 951-967.

Banzon, J. A. (1980). The coconut as a renewable energy source. Philippine Journal


of Coconut Studies, 5(1), 31-36.

Benedictto, G. P., Legnoverde, M. S., Tara, J. C., Sotelo, R. M., & Basaldella, E. I.
(2019). Synthesis of K+/MgO heterogeneous catalysts derived from MgCO3 for
biodiesel production. Materials Letters, 246, 199-202.

Bhojaraj, Harley, P., & Rajamathi, M. (2019). Cannizzaro reactions over calcined
hydrotalcite. Applied Clay Science, 174, 86-89.

Birla, A., Singh, B., Upadhyay, S. N., & Sharma, Y. C. (2012). Kinetics studies of
synthesis of biodiesel from waste frying oil using a heterogeneous catalyst
derived from snail shell. Bioresource Technology, 106, 95-100.

Boro, J., Thakur, A. J., & Deka, D. (2011). Solid oxide derived from waste shells of
Turbonilla striatula as a renewable catalyst for biodiesel production. Fuel
Processing Technology, 92(10), 2061-2067.

Borges, M. E., & Díaz, L. (2012). Recent developments on heterogeneous catalysts


for biodiesel production by oil esterification and transesterification reactions: a
review. Renewable and Sustainable Energy Reviews, 16(5), 2839-2849.

Brinker, C. J., & Scherer, G. W. (2013). Sol-gel science: the physics and chemistry of
sol-gel processing. Academic press.

Carvalho, L. S., Vitor Sobrinho, E., Ruiz, D., & Melo, D. M. D. A. (2018). Effect of Urea
Excess on the Properties of the MgAl2O4 Obtained by Microwave-Assisted
Combustion. Materials Research, 21(1).

73
Chambers, E. S., Byrne, C. S., & Frost, G. (2019). Carbohydrate and human health: is
it all about quality?. The Lancet, 393(10170), 384-386.

Chowdhury, S., Dhawane, S. H., Jha, B., Pal, S., Sagar, R., Hossain, A., & Halder, G.
(2019). Biodiesel synthesis from transesterified Madhuca indica oil by waste
egg shell–derived heterogeneous catalyst: parametric optimization by Taguchi
approach. Biomass Conversion and Biorefinery, 1-11.

Ciobanu, A., Ruellan, S., Mallard, I., Landy, D., Gennequin, C., Siffert, S., &
Fourmentin, S. (2013). Cyclodextrin-intercalated layered double hydroxides for
fragrance release. Journal of Inclusion Phenomena and Macrocyclic
Chemistry, 75(3-4), 333-339.

Coelho, A., Perrone, O. M., Gomes, E., Da-Silva, R., Thoméo, J. C., & Boscolo, M.
(2017). Mixed metal oxides from sucrose and cornstarch templated
hydrotalcite-like LDHs as catalysts for ethyl biodiesel synthesis. Applied
Catalysis A: General, 532, 32-39.

Colucci, J. A., Borrero, E. E., & Alape, F. (2005). Biodiesel from an alkaline
transesterification reaction of soybean oil using ultrasonic mixing. Journal of
the American Oil Chemists' Society, 82(7), 525-530.

Coral, N., Brasil, H., Rodrigues, E., da Costa, C. E., & Rumjanek, V. (2019).
Microwave-modified hydrotalcites for the transesterification of soybean
oil. Sustainable Chemistry and Pharmacy, 11, 49-53.

Da Silva César, A., Conejero, M. A., Ribeiro, E. C. B., & Batalha, M. O. (2019).
Competitiveness analysis of “social soybeans” in biodiesel production in
Brazil. Renewable energy, 133, 1147-1157.

Dabi, M., & Saha, U. K. (2019). Application potential of vegetable oils as alternative to
diesel fuels in compression ignition engines: A review. Journal of the Energy
Institute, 92(6), 1710-1726.

74
Dahdah, E., Estephane, J., Haydar, R., Youssef, Y., El Khoury, B., Gennequin, C., ...
& Aouad, S. (2020). Biodiesel production from refined sunflower oil over Ca–
Mg–Al catalysts: Effect of the composition and the thermal
treatment. Renewable Energy, 146, 1242-1248.

Daud, W. M. A. W., & Ali, W. S. W. (2004). Comparison on pore development of


activated carbon produced from palm shell and coconut shell. Bioresource
technology, 93(1), 63-69.

De, A., & Boxi, S. S. (2020). Application of Cu impregnated TiO2 as a heterogeneous


nanocatalyst for the production of biodiesel from palm oil. Fuel, 265, 117019.

Department of Statistics Malaysia (2018) Press release: Selected agricultural


Indicators, Malaysia, 2018. Retrieved from
https://www.dosm.gov.my/v1/index.php?r=column/cthemeByCat&cat=72&bul
_id=UjYxeDNkZ0xOUjhFeHpna20wUUJOUT09&menu_id=Z0VTZGU1UHBU
T1VJMFlpaXRRR0xpdz09

Demiral, I., & Şensöz, S. (2008). The effects of different catalysts on the pyrolysis of
industrial wastes (olive and hazelnut bagasse). Bioresource
technology, 99(17), 8002-8007.

Dhawan, M. S., Barton, S. C., & Yadav, G. D. (2018). Interesterification of triglycerides


with methyl acetate for the co-production biodiesel and triacetin using
hydrotalcite as a heterogenous base catalyst. Catalysis Today, 309 (1), 161-
171.

Di Blasi, C., Buonanno, F., & Branca, C. (1999). Reactivities of some biomass chars
in air. Carbon, 37(8), 1227-1238.

Di Serio, M., Cozzolino, M., Giordano, M., Tesser, R., Patrono, P., & Santacesaria, E.
(2007). From homogeneous to heterogeneous catalysts in biodiesel
production. Industrial & Engineering Chemistry Research, 46(20), 6379-6384.

Di Serio, M., Mallardo, S., Carotenuto, G., Tesser, R., & Santacesaria, E. (2012). Mg/Al
hydrotalcite catalyst for biodiesel production in continuous packed bed
reactors. Catalysis today, 195(1), 54-58.

75
Dixit, M., Mishra, M., Joshi, P. A., & Shah, D. O. (2013). Physico-chemical and catalytic
properties of Mg–Al hydrotalcite and Mg–Al mixed oxide supported copper
catalysts. Journal of Industrial and Engineering Chemistry, 19(2), 458-468.

Endalew, A. K., Kiros, Y., & Zanzi, R. (2011). Inorganic heterogeneous catalysts for
biodiesel production from vegetable oils. Biomass and bioenergy, 35(9), 3787-
3809.

Fontaras, G., Kousoulidou, M., Karavalakis, G., Bakeas, E., & Samaras, Z. (2011).
Impact of straight vegetable oil–diesel blends application on vehicle regulated
and non-regulated emissions over legislated and real world driving
cycles. Biomass and bioenergy, 35(7), 3188-3198.

Food and Agriculture Organization of United Nation (2019). FAOSTAT Crops Malaysia
coconut production 2017. Retrieved from
http://www.fao.org/faostat/en/#data/QC

Food and Agriculture Organization of United Nation (2019). FAOSTAT Crops Malaysia
paddy production 2017. Retrieved from
http://www.fao.org/faostat/en/#data/QC

Fraile, J. M., García, N., Mayoral, J. A., Pires, E., & Roldán, L. (2009). The influence
of alkaline metals on the strong basicity of Mg–Al mixed oxides: the case of
transesterification reactions. Applied Catalysis A: General, 364(1-2), 87-94.

Gao, M., Khalkhali, M., Beck, S., Choi, P., & Zhang, H. (2018). Study of Thermal
Stability of Hydrotalcite and Carbon Dioxide Adsorption Behavior on
Hydrotalcite-Derived Mixed Oxides Using Atomistic Simulations. ACS
omega, 3(9), 12041-12051.

Gao, L., & Goldfarb, J. L. (2019). Solid waste to biofuels and heterogeneous sorbents
via pyrolysis of wheat straw in the presence of fly ash as an in situ
catalyst. Journal of analytical and applied pyrolysis, 137, 96-105.

Gao, L., Teng, G., Xiao, G., & Wei, R. (2010). Biodiesel from palm oil via loading
KF/Ca–Al hydrotalcite catalyst. Biomass and bioenergy, 34(9), 1283-1288.

76
Gärtner, S. O., & Reinhardt, G. A. (2005). Final report biodiesel initiatives in Germany.
PREMIA report. Heidelberg, Germany: IFEU Institute for Energy and
Environmental Research.

Geng, N., Zhang, Y., Sun, Y., & Geng, S. (2019). Optimization of Biodiesel Supply
Chain Produced from Waste Cooking Oil: A Case Study in China. In IOP
Conference Series: Earth and Environmental Science 264(1) 012006.

Gomes, J. F., Puna, J. F., Gonçalves, L. M., & Bordado, J. C. (2011). Study on the use
of MgAl hydrotalcites as solid heterogeneous catalysts for biodiesel
production. Energy, 36(12), 6770-6778.

Goswami, K., & Choudhury, H. K. (2019). Biofuels versus food: Understanding the
trade-offs between climate friendly crop and food security. World Development
Perspectives, 13, 10-17.

Hall, S. R., Bolger, H., & Mann, S. (2003). Morphosynthesis of complex inorganic forms
using pollen grain templates. Chemical Communications, (22), 2784-2785.

Han, C. G., Zhu, C., Saito, G., & Akiyama, T. (2015). Glycine/sucrose-based solution
combustion synthesis of high-purity LiMn2O4 with improved yield as cathode
materials for lithium-ion batteries. Advanced Powder Technology, 26(2), 665-
671.

Hanisah, K., Kumar, S., and Tajul, AY. (2013). The management of waste cooking oil:
a preliminary survey. Health and Environ J, 4, 76-81.

Haocui, Z., Zhourong, X., Mei, Y., Yajie, T., Guozhu, L., Xiangwen, Z., & Guozhu, L.
(2020). Catalytic steam reforming of JP-10 over Ni/SBA-15. International
Journal of Hydrogen Energy.

Helwani, Z., Aziz, N., Bakar, M. Z. A., Mukhtar, H., Kim, J., & Othman, M. R. (2013).
Conversion of Jatropha curcas oil into biodiesel using re-crystallized
hydrotalcite. Energy Conversion and Management, 73, 128-134.

77
Helwani, Z., Aziz, N., Kim, J., & Othman, M. R. (2016). Improving the yield of Jatropha
curcas's FAME through sol–gel derived meso-porous
hydrotalcites. Renewable energy, 86, 68-74.

Hsieh, L. S., Kumar, U., & Wu, J. C. (2010). Continuous production of biodiesel in a
packed-bed reactor using shell–core structural Ca (C3H7O3)2/CaCO3
catalyst. Chemical Engineering Journal, 158(2), 250-256.

Huaping, Z., Zongbin, W. U., Yuanxiong, C., Zhang, P., Shijie, D., Xiaohua, L. I. U., &
Zongqiang, M. A. O. (2006). Preparation of biodiesel catalyzed by solid super
base of calcium oxide and its refining process. Chinese Journal of
Catalysis, 27(5), 391-396.

Jaaffar, A. R., Baharom, N., Jaguli, A. R., Shamim, A., & Shamsuddin, J. (2019). Bio-
Coconut Charcoal and Social Enterprising. In International Conference on
Innovative Applied Energy (IAPE 2019). Oxford University, UK.

Jahirul, M. I., Rasul, M. G., Chowdhury, A. A., & Ashwath, N. (2012). Biofuels
production through biomass pyrolysis – a technological
review. Energies, 5(12), 4952-5001.

Jha, A., & Das, S. (2017). Waste cooking oil–revolution in biodiesel


production. Fermentation Technology, 6(2).

Kabir, I., Yacob, M., and Radam, A. (2014). Households’ awareness, attitudes and
practices regarding waste cooking oil recycling in Petaling, Malaysia. IOSR-
JESTFT, 8(10), 45-51.

Kataria, J., Mohapatra, S. K., Kundu, K., & Singh, R. (2019). Performance and
Emission Studies of a Variable Compression Ratio CI Engine using Bio Diesel
Made from Waste Cooking Oil. International Journal, 7(3), 265-268.

Kaur, N., & Ali, A. (2014). Kinetics and reusability of Zr/CaO as heterogeneous catalyst
for the ethanolysis and methanolysis of Jatropha crucas oil. Fuel processing
technology, 119, 173-184.

78
Khalaf, A. A., Desa, S., & Baharum, S. N. B. (2019). Overview of Selected Native
Seeds in Agricultural Wastes and its Properties. Medico-Legal Update, 19(2),
324-330.

Krishnamurthy, K. N., Sridhara, S. N., & Kumar, C. A. (2020). Optimization and kinetic
study of biodiesel production from Hydnocarpus wightiana oil and dairy waste
scum using snail shell CaO nano catalyst. Renewable Energy, 146, 280-296.

Kumar, U., & Gupta, P. (2020). Modeling and optimization of novel biodiesel production
from non-edible oil with musa balbisiana root using hybrid response surface
methodology along with african buffalo optimization. Reaction Kinetics,
Mechanisms and Catalysis, 1-27.

Kwon, D., Kang, J. Y., An, S., Yang, I., & Jung, J. C. (2020). Tuning the base properties
of Mg–Al hydrotalcite catalysts using their memory effect. Journal of Energy
Chemistry, 46, 229-236.

Lam, M. K., Lee, K. T., & Mohamed, A. R. (2009). Life cycle assessment for the
production of biodiesel: a case study in Malaysia for palm oil versus jatropha
oil. Biofuels, Bioproducts and Biorefining: Innovation for a sustainable
economy, 3(6), 601-612.

Lam, M. K., Tan, K. T., Lee, K. T., & Mohamed, A. R. (2009). Malaysian palm oil:
Surviving the food versus fuel dispute for a sustainable future. Renewable and
Sustainable Energy Reviews, 13(6), 1456-1464.

Lazarova, T., Kovacheva, D., Georgieva, M., Tzankov, D., Tyuliev, G., Spassova, I., &
Naydenov, A. (2019). Tunable nanosized spinel manganese ferrites
synthesized by solution combustion method. Applied Surface Science, 496,
143571.

Leung, D. Y. C., & Guo, Y. (2006). Transesterification of neat and used frying oil:
optimization for biodiesel production. Fuel processing technology, 87(10), 883-
890.

Leung, D. Y., Wu, X., & Leung, M. K. H. (2010). A review on biodiesel production using
catalyzed transesterification. Applied energy, 87(4), 1083-1095.

79
Li, W., Yang, K., Peng, J., Zhang, L., Guo, S., & Xia, H. (2008). Effects of carbonization
temperatures on characteristics of porosity in coconut shell chars and activated
carbons derived from carbonized coconut shell chars. Industrial crops and
products, 28(2), 190-198.

Lin, Y., Adebajo, M., Frost, R., & Kloprogge, J. (2005). Thermogravimetric analysis of
hydrotalcites based on the takovite formula NixZn6-xAl2 (OH) 16 (CO3).
4H2O. Journal of thermal analysis and calorimetry, 81(1), 83-89.

Liu, H., Wierzbicki, D., Debek, R., Motak, M., Grzybek, T., Da Costa, P., & Gálvez, M.
E. (2016). La-promoted Ni-hydrotalcite-derived catalysts for dry reforming of
methane at low temperatures. Fuel, 182, 8-16.

Liu, Y., Lotero, E., Goodwin Jr, J. G., & Mo, X. (2007). Transesterification of poultry fat
with methanol using Mg–Al hydrotalcite derived catalysts. Applied Catalysis A:
General, 331, 138-148.

Liu, Y., Tu, Q., Knothe, G., & Lu, M. (2017). Direct transesterification of spent coffee
grounds for biodiesel production. Fuel, 199, 157-161.

Ma, Y., Wang, Q., Sun, X., Wu, C., & Gao, Z. (2017). Kinetics studies of biodiesel
production from waste cooking oil using FeCl3-modified resin as heterogeneous
catalyst. Renewable Energy, 107, 522-530.

Ma, Y., Wang, Q., Zheng, L., Gao, Z., Wang, Q., & Ma, Y. (2016). Mixed
methanol/ethanol on transesterification of waste cooking oil using Mg/Al
hydrotalcite catalyst. Energy, 107, 523-531.

Mansir, N., Teo, S. H., Ibrahim, M. L., & Hin, T. Y. Y. (2017). Synthesis and application
of waste egg shell derived CaO supported W-Mo mixed oxide catalysts for
FAME production from waste cooking oil: Effect of stoichiometry. Energy
Conversion and Management, 151, 216-226.

Marinković, D. M., Stanković, M. V., Veličković, A. V., Avramović, J. M., Miladinović,


M. R., Stamenković, O. O., ... & Jovanović, D. M. (2016). Calcium oxide as a
promising heterogeneous catalyst for biodiesel production: Current state and
perspectives. Renewable and Sustainable Energy Reviews, 56, 1387-1408.

80
Martins, M. I., Piresb, R. F., Alvesb, M. J., Horib, C. E., Reisb, M. H., & Cardosob, V.
L. (2013). Transesterification of soybean oil for biodiesel production using
hydrotalcite as basic catalyst. Chemical Engineering Transactions, 32,
817, 822.

Martunus, Othman, M. R., & Fernando, W. J. N. (2011). Elevated temperature carbon


dioxide capture via reinforced metal hydrotalcite. Microporous and mesoporous
materials, 138(1-3), 110-117.

Marwan, Suhendrayatna & Indarti, E. (2015). Preparation of biodiesel from microalgae


and palm oil by direct transesterification in a batch microwave reactor. In
Journal of Physics: Conference Series (Vol. 622, No. 1, p. 012040). IOP
Publishing.

Mueller, D., Ferrão, M. F., Marder, L., Da Costa, A. B., & De Cássia de Souza
Schneider, R. (2013). Fourier transform infrared spectroscopy (FTIR) and
multivariate analysis for identification of different vegetable oils used in
biodiesel production. Sensors, 13(4), 4258-4271.

Mu’az, N. M., Dambatta, M. S., Tukur, S. A., Abdullahi, B., & Abdullahi, U. (2016)
Effects of Using Rice Husk and Paper Pulp as Organic Binding Agents on
Calorific Value of Biomass (Sawdust) Briquettes. 2nd National Engineering
Conference, ACICON 20ssss16, 232.

Navajas, A., Campo, I., Arzamendi, G., Hernández, W. Y., Bobadilla, L. F., Centeno,
M. A., ... & Gandía, L. M. (2010). Synthesis of biodiesel from the methanolysis
of sunflower oil using PURAL® Mg–Al hydrotalcites as catalyst
precursors. Applied Catalysis B: Environmental, 100(1-2), 299-309.

Nisar, J., Razaq, R., Farooq, M., Iqbal, M., Khan, R. A., Sayed, M., Shah, A. & ur
Rahman, I. (2017). Enhanced biodiesel production from Jatropha oil using
calcined waste animal bones as catalyst. Renewable Energy, 101, 111-119.

81
Norhafizah, M., Ramadhansyah, P., Siti Nur Amiera, J., Nurfatin Aqeela, M.,
Norhidayah, A. H., Hainin, M. R., & Che Norazman, C. W. (2016). The effect of
coconut shell on engineering properties of porous asphalt mixture. Jurnal
Teknologi, 78(7-2).

Obadiah, A., Swaroopa, G. A., Kumar, S. V., Jeganathan, K. R., & Ramasubbu, A.
(2012). Biodiesel production from palm oil using calcined waste animal bone
as catalyst. Bioresource technology, 116, 512-516.

Onoji, S. E., Iyuke, S. E., Igbafe, A. I., & Daramola, M. O. (2017). Transesterification
of rubber seed oil to biodiesel over a calcined waste rubber seed shell catalyst:
Modeling and optimization of process variables. Energy & Fuels, 31(6), 6109-
6119.

Othman, M. R., Helwani, Z., Martunus & Fernando, W. J. N. (2009). Synthetic


hydrotalcites from different routes and their application as catalysts and gas
adsorbents: a review. Applied Organometallic Chemistry, 23(9), 335-346.

Paikaray, S., & Hendry, M. J. (2014). Formation and crystallization of Mg2+–Fe3+–


SO42−–CO32−-type anionic clays. Applied clay science, 88, 111-122.

Paredes, S. P., Fetter, G., Bosch, P., & Bulbulian, S. (2006). Sol-gel synthesis of
hydrotalcite—like compounds. Journal of materials science, 41(11), 3377-
3382.

Patil, K. C., Aruna, S. T., & Mimani, T. (2002). Combustion synthesis: an


update. Current opinion in solid state and materials science, 6(6), 507-512.

Peña, A. G., Franseschi, F. A., Estrada, M. C., Ramos, V. M., Zarracino, R. G., Loría,
J. C. Z., & Quiroz, A. V. C. (2014). Fourier transform infrared-attenuated total
reflectance (FTIR-ATR) spectroscopy and chemometric techniques for the
determination of adulteration in petrodiesel/biodiesel blends. Química
Nova, 37(3), 392-397.

Pereira, C. O., Portilho, M. F., Henriques, C. A., & Zotin, F. M. (2014). SnSO4 as
catalyst for simultaneous transesterification and esterification of acid soybean
oil. Journal of the Brazilian Chemical Society, 25(12), 2409-2416.

82
Pinthong, P., Praserthdam, P., & Jongsomjit, B. (2019). Effect of calcination
temperature on Mg-Al layered double hydroxides (LDH) as promising catalysts
in oxidative dehydrogenation of ethanol to acetaldehyde. Journal of oleo
science, 68(1), 95-102.

Poudel, J., Karki, S., Sanjel, N., Shah, M., & Oh, S. C. (2017). Comparison of biodiesel
obtained from virgin cooking oil and waste cooking oil using supercritical and
catalytic transesterification. Energies, 10(4), 546.

Praveena, S. M., Rashid, U., & Rashid, S. A. (2019). Application of activated carbon
from banana stem waste for removal of heavy metal ions in greywater using a
Box–Behnken design approach. Environmental technology, 1-12.

Pütün, E., Uzun, B. B., & Pütün, A. E. (2006). Fixed-bed catalytic pyrolysis of cotton-
seed cake: effects of pyrolysis temperature, natural zeolite content and
sweeping gas flow rate. Bioresource Technology, 97(5), 701-710.

Rabelo, S. N., Ferraz, V. P., Oliveira, L. S., & Franca, A. S. (2015). FTIR analysis for
quantification of fatty acid methyl esters in biodiesel produced by microwave-
assisted transesterification. International Journal of Environmental Science and
Development, 6(12), 964.

Rafiee, E., Shahebrahimi, S., Feyzi, M., & Shaterzadeh, M. (2012). Optimization of
synthesis and characterization of nanosilica produced from rice husk (a
common waste material). International nano letters, 2(1), 29.

Ramimoghadam, D., Hussein, M. Z. B., & Taufiq-Yap, Y. H. (2013). Hydrothermal


synthesis of zinc oxide nanoparticles using rice as soft biotemplate. Chemistry
Central Journal, 7(1), 136.

Ramli, A., & Farooq, M. (2015). Optimization of process parameters for the production
of biodiesel from waste cooking oil in the presence of bifunctional γ-Al2O3-CeO2
supported catalysts. Malaysian Journal of Analytical Sciences, 19(1), 8-19.

83
Ramli, A., Farooq, M., Naeem, A., Khan, S., Hummayun, M., Iqbal, A., Sohail, A. &
Shah, L. A. (2017). Bifunctional heterogeneous catalysts for biodiesel
production using low cost feedstocks: a future perspective. Frontiers in
Bioenergy and Biofuels, 285.

Ramos-Ramírez, E., Gutiérrez-Ortega, N., Tzompantzi, F., Gómez, C. M., Del Angel,
G., Herrera-Pérez, G., Seraf´ın-Muñoz, A. H. & Tzompantzi-Flores, C. (2018).
Activated Hydrotalcites Obtained by Coprecipitation as Photocatalysts for the
Degradation of 2, 4, 6-Trichlorophenol. Advances in Materials Science and
Engineering, 2018.

Raqeeb, M. A., & Bhargavi, R. (2015). Biodiesel production from waste cooking
oil. Journal of Chemical and Pharmaceutical Research, 7(12), 670-681.

Rauch, A., & Thöne, M. (2012). Biofuels–at what cost? Mandating ethanol and
biodiesel consumption in Germany. FiFo Institute for the Global Subsidies
Initiative (GSI) of the International Institute for Sustainable Development (IISD),
Geneva, Switzerland.

Rahul, R., Satyarthi, J. K., & Srinivas, D. (2011). Lanthanum and zinc incorporated
hydrotalcites as solid base catalysts for biodiesel and biolubricants production.

REN21, R. (2018). Global Status Report, REN21 Secretariat, Paris. France. Tech.
Rep.

Reyero, I., Velasco, I., Sanz, O., Montes, M., Arzamendi, G., & Gandía, L. M. (2013).
Structured catalysts based on Mg–Al hydrotalcite for the synthesis of
biodiesel. Catalysis today, 216, 211-219.

Rezvani, Z., Rad, F. A., & Khodam, F. (2015). Synthesis and characterization of Mg–
Al-layered double hydroxides intercalated with cubane-1,4-dicarboxylate
anions. Dalton Transactions, 44(3), 988-996.

Rosmiza, M. Z., Davies, W. P., Rosniza Aznie, CR, Mazdi, M., & Jabil, M. J. (2017).
Farmers’ knowledge on potential uses of rice straw: an assessment in MADA
and Sekinchan, Malaysia. Geografia-Malaysian Journal of Society and
Space, 10(5).

84
Sahoo, S. K., Khandelwal, V., & Manik, G. (2019). Synthesis and characterization of
low viscous and highly acrylated epoxidized methyl ester based green
adhesives derived from linseed oil. International Journal of Adhesion and
Adhesives, 89, 174-177.

Said, N. H., Ani, F. N., & Said, M. F. M. (2015). Review of the production of biodiesel
from waste cooking oil using solid catalysts. Journal of Mechanical Engineering
and Sciences, 8, 1302-1311.

Salimi, Z., & Hosseini, S. A. (2019). Study and optimization of conditions of biodiesel
production from edible oils using ZnO/BiFeO3 nano magnetic
catalyst. Fuel, 239, 1204-1212.

Sangarunlert, W., Piumsomboon, P., & Ngamprasertsith, S. (2007). Furfural


production by acid hydrolysis and supercritical carbon dioxide extraction from
rice husk. Korean Journal of Chemical Engineering, 24(6), 936-941.

Saxena, R. (2019). An Overview of Kusum Based Biodiesel: An Optimum Fuel.

Selvam, N. C. S., Kumar, R. T., Kennedy, L. J., & Vijaya, J. J. (2011). Comparative
study of microwave and conventional methods for the preparation and optical
properties of novel MgO-micro and nano-structures. Journal of Alloys and
Compounds, 509(41), 9809-9815.

Shabanian, M., Hajibeygi, M., & Raeisi, A. (2020). FTIR characterization of layered
double hydroxides and modified layered double hydroxides. In Layered Double
Hydroxide Polymer Nanocomposites (pp. 77-101). Woodhead Publishing.

Shafie, S. M. (2015). Paddy residue based power generation in Malaysia:


Environmental assessment using LCA approach. ARPN Journal of Engineering
and Applied Sciences, 10, 6643-6648.

Sharma, Y. C., Singh, B., & Korstad, J. (2011). Latest developments on application of
heterogenous basic catalysts for an efficient and eco friendly synthesis of
biodiesel: A review. Fuel, 90(4), 1309-1324.

85
Shekoohi, K., Hosseini, F. S., Haghighi, A. H., & Sahrayian, A. (2017). Synthesis of
some Mg/Co-Al type nano hydrotalcites and characterization. MethodsX, 4, 86-
94.

Silveira, E. G., Barcelos, L. F. T., Perez, V. H., Justo, O. R., Ramirez, L. C., de Moraes
Rêgo Filho, L., & de Castro, M. P. P. (2019). Biodiesel production from non-
edible forage turnip oil by extruded catalyst. Industrial Crops and
Products, 139, 111503.

Singh, A. K., & Fernando, S. D. (2007). Reaction kinetics of soybean oil


transesterification using heterogeneous metal oxide catalysts. Chemical
Engineering & Technology: Industrial Chemistry‐Plant Equipment‐Process
Engineering‐Biotechnology, 30(12), 1716-1720.

Singh, B., Bux, F., & Sharma, Y. C. (2011). Comparison of homogeneous and
heterogeneous catalysis for synthesis of biodiesel from M. indica oil. Chemical
Industry and Chemical Engineering Quarterly, 17(2), 117-124.

Singh, R., Kumar, A., & Chandra Sharma, Y. (2019). Biodiesel production from
microalgal oil using Barium–Calcium–Zinc mixed oxide base catalyst:
Optimization and kinetic studies. Energy & Fuels, 33(2), 1175-1184.

Sinico, S. (2006, August). Germany Phasing Out Price Protection for Biodiesel. DW.

Sivakumar, P., Anbarasu, K., & Renganathan, S. (2011). Bio-diesel production by


alkali catalyzed transesterification of dairy waste scum. Fuel, 90(1), 147-151.

Sobhana, S. L., Bogati, D. R., Reza, M., Gustafsson, J., & Fardim, P. (2016). Cellulose
biotemplates for layered double hydroxides networks. Microporous and
Mesoporous Materials, 225, 66-73.

Soriano Jr, N. U., Venditti, R., & Argyropoulos, D. S. (2009). Biodiesel synthesis via
homogeneous Lewis acid-catalyzed transesterification. Fuel, 88(3), 560-565.

Stavarache, C., Vinatoru, M., Nishimura, R., & Maeda, Y. (2003). Conversion of
vegetable oil to biodiesel using ultrasonic irradiation. Chemistry Letters, 32(8),
716-717.

86
Sundus, F., Fazal, M. A., and Masjuki, H. H. (2017). Tribology with biodiesel: A study
on enhancing biodiesel stability and its fuel properties. Renewable and
Sustainable Energy Reviews, 70, 399-412.

Suriyanarayanan, N., Nithin, K. K., & Bernardo, E. (2009). Mullite glass ceramics
production from coal ash and alumina by high temperature plasma. Journal of
Non-Oxide Glasses, 1(4), 247-260.

Talebian-Kiakalaieh, A., Amin, N. A. S., Zarei, A., & Noshadi, I. (2013).


Transesterification of waste cooking oil by heteropoly acid (HPA) catalyst:
optimization and kinetic model. Applied energy, 102, 283-292.

Tan, Y. H., Abdullah, M. O., Nolasco-Hipolito, C., & Taufiq-Yap, Y. H. (2015). Waste
ostrich-and chicken-eggshells as heterogeneous base catalyst for biodiesel
production from used cooking oil: Catalyst characterization and biodiesel yield
performance. Applied Energy, 160, 58-70.

Taufiq-Yap, Y. H., Lee, H. V., Hussein, M. Z., & Yunus, R. (2011). Calcium-based
mixed oxide catalysts for methanolysis of Jatropha curcas oil to
biodiesel. biomass and bioenergy, 35(2), 827-834.

Thangaraj, B., Solomon, P. R., Muniyandi, B., Ranganathan, S., & Lin, L. (2019).
Catalysis in biodiesel production—a review. Clean Energy, 3(1), 2-23.

Thanh, L. T., Okitsu, K., Sadanaga, Y., Takenaka, N., Maeda, Y., & Bandow, H. (2010).
Ultrasound-assisted production of biodiesel fuel from vegetable oils in a small
scale circulation process. Bioresource Technology, 101(2), 639-645.

Thouchprasitchai, N., Pintuyothin, N., & Pongstabodee, S. (2018). Optimization of CO2


adsorption capacity and cyclical adsorption/desorption on
tetraethylenepentamine-supported surface-modified hydrotalcite. Journal of
Environmental Sciences, 65, 293-305.

Tomei, J., & Helliwell, R. (2016). Food versus fuel? Going beyond biofuels. Land use
policy, 56, 320-326.

87
Trejo-Zárraga, F., de Jesús Hernández-Loyo, F., Chavarría-Hernández, J. C., &
Sotelo-Boyás, R. (2018). Kinetics of transesterification processes for biodiesel
production. In Biofuels: state of development (pp. 149-179). IntechOpen.

Tyson, K. S. (2001). Biodiesel--Clean, Green Diesel Fuel: Great Fleet Fuel Gaining
Popularity Rapidly.

Udayakumar, R., Shah, V. J., & Venkatesh, S. V. (2020). Performance and Emission
Characteristics of Biodiesel from Rapeseed and Soybean in CI Engine.
In Intelligent Manufacturing and Energy Sustainability (pp. 209-219). Springer,
Singapore.

U.S. Energy Information Administration (2017) U.S. biodiesel production still


increasing despite expiration of tax credit. Retrieved from
https://www.eia.gov/todayinenergy/detail.php?id=34152

Wahab, A. G. (2019) Malaysia: Biofuels Annual. Retrieved from


https://www.fas.usda.gov/data/malaysia-biofuels-annual-2

Wang, Y., Ou, S., Liu, P., Xue, F., and Tang, S. (2006). Comparison of two different
processes to synthesize biodiesel by waste cooking oil. Journal of Molecular
Catalysis A: Chemical, 252(1), 107-112.

Wang, W., Gowdagiri, S., & Oehlschlaeger, M. A. (2014). The high-temperature


autoignition of biodiesels and biodiesel components. Combustion and
flame, 161(12), 3014-3021.

West, A. H., Posarac, D., & Ellis, N. (2008). Assessment of four biodiesel production
processes using HYSYS. Plant. Bioresource technology, 99(14), 6587-6601.

Xie, W., Peng, H., & Chen, L. (2006). Calcined Mg–Al hydrotalcites as solid base
catalysts for methanolysis of soybean oil. Journal of Molecular Catalysis A:
Chemical, 246(1-2), 24-32.

Xie, W., & Huang, X. (2006). Synthesis of biodiesel from soybean oil using
heterogeneous KF/ZnO catalyst. Catalysis Letters, 107(1-2), 53-59.

88
Xin, J., Imahara, H., & Saka, S. (2009). Kinetics on the oxidation of biodiesel stabilized
with antioxidant. Fuel, 88(2), 282-286.

Varma, M. N., & Madras, G. (2007). Synthesis of biodiesel from castor oil and linseed
oil in supercritical fluids. Industrial & engineering chemistry research, 46(1), 1-
6.

Yabushita, M., Shibayama, N., Nakajima, K., & Fukuoka, A. (2019). Selective glucose-
to-fructose isomerization in ethanol catalyzed by hydrotalcites. ACS
Catalysis, 9(3), 2101-2109.

Yahya, N. Y., Ngadi, N., Wong, S., & Hassan, O. (2018). Transesterification of used
cooking oil (UCO) catalyzed by mesoporous calcium titanate: kinetic and
thermodynamic studies. Energy conversion and management, 164, 210-218.

Yang, Y., Yang, M., Zheng, Z., & Zhang, X. (2020). Highly effective adsorption removal
of perfluorooctanoic acid (PFOA) from aqueous solution using calcined layer-
like Mg-Al hydrotalcites nanosheets. Environmental Science and Pollution
Research, 1-13.

Yusuff, A. S., Adeniyi, O. D., Olutoye, M. A., & Akpan, U. G. (2018). Kinetic study of
transesterification of waste frying oil to biodiesel using anthill eggshell-Ni-Co
mixed oxide composite catalyst. Petroleum & Coal, 60(1).

Zailani, S., Iranmanesh, M., Sean Hyun, S., & Ali, M. H. (2019). Barriers of biodiesel
adoption by transportation companies: A case of Malaysian transportation
industry. Sustainability, 11(3), 931.

Zango, Z. U., Kadir, H. A., Imam, S. S., Muhammad, A. I., & Abu, I. G. (2019).
Optimization studies for catalytic conversion of waste vegetable oil to
biodiesel. American Journal of Chemistry, 9(2), 27-32.

Zeng, H. Y., Liao, K. B., Deng, X., Jiang, H., & Zhang, F. (2009). Characterization of
the lipase immobilized on Mg–Al hydrotalcite for biodiesel. Process
biochemistry, 44(8), 791-798.

89
Zhang, J., Ren, M., Li, X., Hao, Q., Chen, H., & Ma, X. (2020). Ni-based catalysts
prepared for CO2 reforming and decomposition of methane. Energy
Conversion and Management, 205, 112419.

Zhang, L., Sheng, B., Xin, Z., Liu, Q., & Sun, S. (2010). Kinetics of transesterification
of palm oil and dimethyl carbonate for biodiesel production at the catalysis of
heterogeneous base catalyst. Bioresource Technology, 101(21), 8144-8150.

Zhang, N., Santos, R. M., & Šiller, L. (2021). CO2 mineralisation of brines with
regenerative hydrotalcites in a cyclical process. Chemical Engineering
Journal, 404, 126450.

Zheng, S., Kates, M., Dubé, M. A., & McLean, D. D. (2006). Acid-catalyzed production
of biodiesel from waste frying oil. Biomass and bioenergy, 30(3), 267-272.

Ziegler, J. (2013). Burning food crops to produce biofuels is a crime against


humanity. The Guardian, 26.

90
APPENDICES

91
Appendix A: Example of Bomb Calorimeter Data
Sample Name: SS
Weight of Fuel
Crucible (g) Crucible + Fuel (g) Fuel (g), F
6.82 7.9043 1.0843

Weight of Water
Beaker + Water (g) Beaker (g) Water (g), W
2486.32 515.22 1971.1

Time and Temperature Reading


Time (minutes) Temperature (°C)
0 1.740
1 1.740
2 1.740
3 1.740 (sample ignited)
4 1.740
5 1.990
6 2.990
7 3.220
8 3.250
9 3.300
10 3.240
11 3.390
12 3.440
13 3.435 (max temperature)
14 3.430
15 3.420
16 3.405
17 3.395
18 3.385
19 3.375
20 3.365

92
Appendix B: Example of Bomb Calorimeter Calculation
Radiation Correction
n = number of minutes between the ignition and the attainment at end of test
v1 = rate of temperature fall in degree per minute at end of test
v = rate of temperature rise in degree per minute at beginning of test

For v1, taken from 14th to the 18th minute


(3.430 − 3.420) + (3.420 − 3.405) + (3.405 − 3.395) + (3.395 − 3.385)
v1 =
4
𝑣 1 = 0.01125
Hence,

1
−v + v1
Radiation Correction = n v + ( ), n = 5, v=0
2
−0 + 0.01125
Radiation Correction = 5 ∙ 0.01125 + ( )
2
Radiation Correction = 0.108 °C

Acutal Temperature = Max Temperature − Temperature at Ignition

Acutal Temperature = 3.435 − 1.740

Acutal Temperature = 1.695 °C

Corrected Temperature = Actual Temperature + Radiation Correction

Corrected Temperature = 1.695 + 0.108

Corrected Temperature = 1.802 °C

𝑊𝑎𝑡𝑒𝑟 𝑉𝑎𝑙𝑢𝑒 = 259.753


𝐻𝑒𝑎𝑡 𝐴𝑏𝑠𝑜𝑟𝑏𝑒𝑑 𝑏𝑦 𝑊𝑎𝑡𝑒𝑟 = 𝐶𝑜𝑟𝑟𝑒𝑐𝑡𝑒𝑑 𝑇𝑒𝑚𝑝𝑒𝑟𝑎𝑡𝑢𝑟𝑒 × (𝑊 + 𝑊𝑎𝑡𝑒𝑟 𝑉𝑎𝑙𝑢𝑒)
𝐻𝑒𝑎𝑡 𝐴𝑏𝑠𝑜𝑟𝑏𝑒𝑑 𝑏𝑦 𝑊𝑎𝑡𝑒𝑟 = 1.802 × (1971.1 + 259.753)
𝐻𝑒𝑎𝑡 𝐴𝑏𝑠𝑜𝑟𝑏𝑒𝑑 𝑏𝑦 𝑊𝑎𝑡𝑒𝑟 = 4019.718 𝐶𝑎𝑙𝑜𝑟𝑖𝑒𝑠

𝐻𝑒𝑎𝑡 𝐴𝑏𝑠𝑜𝑟𝑏𝑒𝑑 𝑏𝑦 𝑊𝑎𝑡𝑒𝑟


𝐶𝑎𝑙𝑜𝑟𝑖𝑓𝑖𝑐 𝑉𝑎𝑙𝑢𝑒 =
𝐹𝑢𝑒𝑙
4019.718
𝐶𝑎𝑙𝑜𝑟𝑖𝑓𝑖𝑐 𝑉𝑎𝑙𝑢𝑒 =
1.0843
𝐶𝑎𝑙𝑜𝑟𝑖𝑓𝑖𝑐 𝑉𝑎𝑙𝑢𝑒 = 3707.201 𝐶𝑎𝑙𝑜𝑟𝑖𝑒/𝑔

93
Appendix C: Example of Acid Value and Saponification Value
Determination of WCO For Average Molecular Weight Oil
Acid Value Determination
Molality of KOH (N), N 0.05
Blank value (ml), B 0.3

Weight Conical flask (g) 110.53


Weight Conical flask + Oil (g) 111.54
Weight Oil (g), m 1.01

Initial Volume Titre (ml) 30


End Volume Titre (ml) 30.6
Total Volume Titre (ml), T 0.6

(𝑇 − 𝐵) × 𝑁 × 56.1
𝐴𝑐𝑖𝑑 𝑉𝑎𝑙𝑢𝑒 =
𝑚
(0.6 − 0.3) × 0.05 × 56.1
𝐴𝑐𝑖𝑑 𝑉𝑎𝑙𝑢𝑒 =
1.01
𝐴𝑐𝑖𝑑 𝑉𝑎𝑙𝑢𝑒 = 0.833 𝑚𝑔 𝐾𝑂𝐻/𝑔 𝑓𝑎𝑡

Saponification Value Determination


Concentration of HCl (M), C 0.1
Initial Volume Titre (ml) 0
End Volume Titre(ml) 40
Total Volume Titre (ml), T 40

Weight Conical Flask (g) 114.91


Weight Conical Flask + Oil (g) 115.93
Weight Oil (g), m 1.02

𝑇 × 𝐶 × 56
𝑆𝑎𝑝𝑜𝑛𝑖𝑓𝑖𝑐𝑎𝑡𝑖𝑜𝑛 𝑉𝑎𝑙𝑢𝑒 =
𝑚
40 × 0.1 × 56.11 1𝑙 1000𝑚𝑔
𝑆𝑎𝑝𝑜𝑛𝑖𝑓𝑖𝑐𝑎𝑡𝑖𝑜𝑛 𝑉𝑎𝑙𝑢𝑒 = × ×
1.02 1000𝑚𝑙 1𝑔
𝑆𝑎𝑝𝑜𝑛𝑖𝑓𝑖𝑐𝑎𝑡𝑖𝑜𝑛 𝑉𝑎𝑙𝑢𝑒 = 220.039 𝑚𝑔 𝐾𝑂𝐻/𝑔 𝑓𝑎𝑡

94
Appendix D: Example of Average Molecular Weight Oil
Determination of WCO
Average Molecular Weight Oil Determination
Saponification Value (mgKOH/g), SV 211.877
Acid Value (mgKOH/g), AV 0.792

56.1 × 1000 × 3
𝐴𝑣𝑒𝑟𝑎𝑔𝑒 𝑀𝑊 𝑂𝑖𝑙 =
(𝑆𝑉 − 𝐴𝑉)
56.1 × 1000 × 3
𝐴𝑣𝑒𝑟𝑎𝑔𝑒 𝑀𝑊 𝑂𝑖𝑙 =
(211.877 − 0.792)
𝐴𝑣𝑒𝑟𝑎𝑔𝑒 𝑀𝑊 𝑂𝑖𝑙 = 797.31 𝑔/𝑚𝑜𝑙

95
Appendix E: Example of HT Preparation
Calculation of Mg-Al based HT composition
Molecular weight of Mg = 256.41 𝑔/𝑚𝑜𝑙
Molecular weight of Al = 375.13 g/mol

𝑅𝑎𝑡𝑖𝑜 𝑜𝑓 𝑀𝑔
𝑀𝑊 𝑜𝑓 𝑀𝑔 × 3 𝑚𝑜𝑙
=
(𝑀𝑊 𝑜𝑓 𝑀𝑔 × 3 𝑚𝑜𝑙) + (𝑀𝑊 𝑜𝑓 𝐴𝑙 × 1 𝑚𝑜𝑙)
256.41 𝑔/𝑚𝑜𝑙 × 3 𝑚𝑜𝑙
𝑅𝑎𝑡𝑖𝑜 𝑜𝑓 𝑀𝑔 =
(256.41 𝑔/𝑚𝑜𝑙 × 3 𝑚𝑜𝑙) + (375.13 𝑔/𝑚𝑜𝑙 × 1 𝑚𝑜𝑙)
𝑅𝑎𝑡𝑖𝑜 𝑜𝑓 𝑀𝑔 = 0.672

𝑅𝑎𝑡𝑖𝑜 𝑜𝑓 𝐴𝑙
𝑀𝑊 𝑜𝑓 𝐴𝑙 × 1 𝑚𝑜𝑙
=
(𝑀𝑊 𝑜𝑓 𝑀𝑔 × 3 𝑚𝑜𝑙) + (𝑀𝑊 𝑜𝑓 𝐴𝑙 × 1 𝑚𝑜𝑙)
375.13 𝑔/𝑚𝑜𝑙 × 1 𝑚𝑜𝑙
𝑅𝑎𝑡𝑖𝑜 𝑜𝑓 𝐴𝑙 =
(256.41 𝑔/𝑚𝑜𝑙 × 3 𝑚𝑜𝑙) + (375.13 𝑔/𝑚𝑜𝑙 × 1 𝑚𝑜𝑙)
𝑅𝑎𝑡𝑖𝑜 𝑜𝑓 𝐴𝑙 = 0.328

𝑊𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑀𝑔 = 0.672 × 𝑇𝑜𝑡𝑎𝑙 𝑤𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑠𝑎𝑚𝑝𝑙𝑒


𝑊𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝐴𝑙 = 0.328 × 𝑇𝑜𝑡𝑎𝑙 𝑤𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑠𝑎𝑚𝑝𝑙𝑒
𝑊𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑆𝑜𝑑𝑖𝑢𝑚 𝐶𝑎𝑟𝑏𝑜𝑛𝑎𝑡𝑒 = 0.2 × 𝑇𝑜𝑡𝑎𝑙 𝑤𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑠𝑎𝑚𝑝𝑙𝑒
𝑊𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝐹𝑢𝑒𝑙 = 0.1 × 𝑇𝑜𝑡𝑎𝑙 𝑤𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑠𝑎𝑚𝑝𝑙𝑒

96
Appendix F: Example of GC Data

Response (mV)

Time [min]

Name Time (minute)


N Hexane 1.35
Methyl Palmitate 7.84
Internal Standard (Methyl Heptadecanoate) 8.72
Methyl Stearate 9.64
Methyl Oleate 9.85
Methyl Linoleate 10.19

97
Appendix G: Example of FAME Yield Calculation

Sample Name HT SS 650


Concentration Internal standard (g/l), CIS 10
Volume Standard (ml), VIS 0.15
Weight Sample (mg), m 7.5

∑ 𝐴 − 𝐴𝐼𝑆 𝐶𝐼𝑆 × 𝑉𝐼𝑆


𝐹𝐴𝑀𝐸 𝑌𝑖𝑒𝑙𝑑 = ×
𝐴𝐼𝑆 𝑚
Whereas
∑A = Total peak area C14:0 – C24:1
AIS = Internal standard (methyl heptadecanoate) peak area
CIS = Concentration of internal standard solution in g/l
VIS = volume of internal standard used in ml
m = mass of sample in mg

Name Time (minute) Area (μVs)


Methyl Palmitate 1.35 653572.27
Internal Standard (Methyl Heptadecanoate) 7.84 439005.36
Methyl Stearate 8.72 51193.98
Methyl Oleate 9.64 748337.65
Methyl Linoleate 9.85 174341.24

∑ 𝐴 = 653572.27 + 439005.36 + 51193.98 + 748337.65 + 174341.24

∑ 𝐴 = 2066450.50

∑ 𝐴 − 𝐴𝐼𝑆 𝐶𝐼𝑆 × 𝑉𝐼𝑆


𝐹𝐴𝑀𝐸 𝑌𝑖𝑒𝑙𝑑 = × × 100%
𝐴𝐼𝑆 𝑚
2066450.50 − 439005.36 10 × 0.15
𝐹𝐴𝑀𝐸 𝑌𝑖𝑒𝑙𝑑 = × × 100%
439005.36 7.5
𝐹𝐴𝑀𝐸 𝑌𝑖𝑒𝑙𝑑 = 74.14%

98
Appendix H: Example of Kinetic Study Calculation
Convert Yield FAME into ln(1-FAME)
Yield FAME HT SS 650 at 5hr (reaction temperature = 65 °C) = 73.88%

𝑙𝑛(1 − 𝐹𝐴𝑀𝐸) = 𝑙𝑛(1 − 73.88%)


𝑙𝑛(1 − 𝐹𝐴𝑀𝐸) = 1.342

99
Appendix I: ln(1-FAME) versus t graph
2 55 ℃ 60 ℃ 65 ℃ (a)

1.5
ln(1-FAME)

y = 0.2687x + 0.1268
1

y = 0.054x + 0.007
0.5
y = 0.0109x - 0.0083

0
0 1 2 3 4 5
Time (hours)

2.5 55 ℃ 60 ℃ 65 ℃ (b)

2 y = 0.3344x + 0.2675
ln(1-FAME)

1.5

1
y = 0.127x + 0.1111

0.5
y = 0.0172x - 0.0123
0
0 1 2 3 4 5
Time (Hours)

3 55 ℃ 60 ℃ 65 ℃ (c)

2.5
y = 0.4602x + 0.3542
2
ln(1-FAME)

1.5
y = 0.147x + 0.1765
1

0.5 y = 0.0381x - 0.0077

0
0 1 2 3 4 5
Time (Hours)
ln(1-FAME) versus t graph for (a) HT-SS 650, (b) HT-RH 650 and (c) HT-CS 650

100
Appendix J: Activation Energy Calculation

For HT-SS 650 from Figure 4.9,


Slope, m = -35559
ln A = 103.83
R = 8.314 J K/mol

−𝐸𝑎
∴ = −35559
𝑅
−35559 (8.314)
𝐸𝑎 =
−1
𝐸𝑎 = 295637 𝐽 𝐾/𝑚𝑜𝑙
𝐸𝑎 = 295.64 𝑘𝐽 𝐾/𝑚𝑜𝑙

𝑙𝑛 𝐴 = 103.83
𝐴 = 𝑒 103.83
𝐴 = 1.24 × 1045

101

You might also like