+ All Categories
Home > Documents > LATEEF, FATAI ABIOLA

LATEEF, FATAI ABIOLA

Date post: 22-Dec-2021
Category:
Upload: others
View: 5 times
Download: 0 times
Share this document with a friend
126
BATCH PRODUCTION OF FATTY ACID PROPYL ESTERS (BIODIESEL) FROM OILSEED CROPS BY LATEEF, FATAI ABIOLA PG/M.Sc./07/43230 A THESIS SUBMITTED TO THE SCHOOL OF POST GRADUATE STUDIES IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE AWARD OF THE DEGREE OF MASTER OF SCIENCE IN THE DEPARTMENT OF PURE AND INDUSTRIAL CHEMISTRY, UNIVERSITY OF NIGERIA, NSUKKA, ENUGU STATE MARCH, 2010
Transcript
Page 1: LATEEF, FATAI ABIOLA

BATCH PRODUCTION OF FATTY ACID PROPYL ESTERS

(BIODIESEL) FROM OILSEED CROPS

BY

LATEEF, FATAI ABIOLA

PG/M.Sc./07/43230

A THESIS

SUBMITTED TO THE SCHOOL OF POST GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

AWARD OF THE DEGREE OF MASTER OF SCIENCE IN THE

DEPARTMENT OF PURE AND INDUSTRIAL CHEMISTRY,

UNIVERSITY OF NIGERIA, NSUKKA,

ENUGU STATE

MARCH, 2010

Page 2: LATEEF, FATAI ABIOLA

ii

APPROVAL PAGE

LATEEF, Fatai Abiola a postgraduate student in the Department of

Pure and Industrial Chemistry with Registration Number PG/M.Sc/07/43230

has satisfactorily completed the requirements for course and research work

for the degree of M.Sc. in Industrial Chemistry.

__________________ ____________________

Prof. O.D. Onukwuli Dr. U.C. Okoro

Supervisor Supervisor

_____________________

Dr. P.O. Ukoha

Head of Department

Page 3: LATEEF, FATAI ABIOLA

iii

DEDICATION

This work is dedicated to God and to Popoola Folarin (SAN),

Opaluwa E.H.O and Abeh T. for their financial and moral support.

Page 4: LATEEF, FATAI ABIOLA

iv

ACKNOWLEDGEMENTS

I would like to thank my supervisors, Prof. O.D. Onukwuli and Dr.

U.C. Okoro for firstly, giving me this opportunity to have a part in this

project and more importantly giving me support throughout both my course

and project work. I would also like to say that their faith in me has given me

confidence in my ability to be an industrial chemist from this work, I have

learnt to appreciate hard work as the basis of success in all human

endeavuors. I appreciate the spirit of hardwork that they have instilled in me.

I would also like to thank Mr. P.M. Ejikeme for his assistance and

input on cloudy issues and the many offer of much needed help in writing

this thesis. Also in this category are all the lecturers (both academic and non-

academics) of the department of Pure and Industrial Chemistry, UNN;

Departments of Science Laboratory Technology, and Food Science and

Technology, Federal Polytechnic, Idah.

I would also like to say thank you to Mr. B.I.D. Obidiegwu of

Analytical laboratory, Soil Science department, UNN for his input in

analytical chemistry work and Mr. Fidelis Eze of Fluid Mechanics

Laboratory, Civil Engineering Department, UNN.

I thank all my well wishers and admirers who stood by me during my

masters programme till the end. Thanks to Mr. M.K. Egbekun, Dr.E.M Odin,

Late I.E.O. Salawo, Mr Okorie, E., Oresanya Adebisi, Daodu Niyi, Afolabi

Safara, Adeyemo Nurudeen, Atiga Simeon, Eguda Felix, Ajayi Ebenezer,

Dosunmu Oladipupo, Tijani Memuna, the entire Muslim Community UNN,

Osayuki Godwin and Joseph Jesuorobo of ICT unit, FPI Idah for the

provision of materials used in writing this thesis free of charge.

Finally, I would like to thank my parents, my extended family

members for standing by me, without them I would not have been able to

complete much of what I have done and become who I am. They are the

light that shines my way and the drive for my ever persistent determination.

Page 5: LATEEF, FATAI ABIOLA

v

ABSTRACT

Biodiesel fuels were prepared from castor, palm kernel and groundnut oils

through alkali transesterification reaction using sodium hydroxide (NaOH)

as the catalyst at 65oC, residence time of 60 minutes, 0.50g optimal catalyst

weight for castor oil and palm kernel oil and 0.40g for groundnut oil in an

air-tight 250ml batch reactor at 1:6 oil/alcohol molar ratio. The biodiesels

produced were characterized as alternative diesel fuels through ASTM D

6751, EN 14214 specification standards for the fuel properties: specific

gravity, viscosity, calorific (combustion) value, saponification value, iodine

value and acid value. The castor oil (CSO), palm kernel oil (PKO) and

groundnut oil (GNO) biodiesels displayed temperature dependent behaviour

when their respective viscosities were measured at 30, 45, 60 and 75oC. In

the kinetic studies, biodiesels were produced at 5, 10, 20 and 60 minutes, the

volume of biodiesel and their various percentage (%) conversions evaluated

at different temperatures of 30, 45, 60 and 75oC. The CSO, PKO and GNO

biodiesels were also measured for their storage/oxidative stabilities

compared with commercial petrodiesel. The biodiesels produced were of

good fuel properties with respect to ASTM D 6751 and EN 14214

specification standards (except kinematic viscosity of castor oil biodiesel).

The viscosities of castor oil biodiesel at different temperatures were in the

range of 6.89-8.12 mm2/s which were higher than that of European biodiesel

standards (EN 14214 : 3.5–5.0) and American Society of Testing Material

Standards (ASTM D 6751 : 1.9-6.0). However, promising results which

conforms to above specification standards were realized when castor oil

biodiesels were blended with commercial petrodiesel. The specific gravities

recorded for CSO, PKO and GNO biodiesel were higher than the values

obtained for petrodiesel: 1.049 times that of the petrodiesel for CSO, 1.086

times that of petrodiesel for PKO and 1.078 times that of petrodiesel for

GNO at 28oC. In the oxidative storage stability test, commercial petrodiesel

has the highest oxidative stability than biodiesel produced from CSO, PKO

and GNO oils. The infrared spectra result of a B100 biodiesel shows that the

strong ester peaks near 1750 (the C=O vibration) and around 1170-1120cm-1

(C-O vibration).

Page 6: LATEEF, FATAI ABIOLA

vi

TABLE OF CONTENTS

Title Page - - - - - - - - - i

Certification - - - - - - - - ii

Dedication - - - - - - - - - iii

Acknowledgements - - - - - - - iv

Abstract - - - - - - - - - v

Table of Contents - - - - - - - - vi

List of Figures - - - - - - - - x

List of Tables - - - - - - - - xii

CHAPTER ONE

INTRODUCTION

1.1 Preamble - - - - - - - - 1

1.2 Justification of Study - - - - - - 4

1.3 Aim and Objectives of Research - - - - 5

CHAPTER TWO

LITERATURE REVIEW

2.1 History of Biodiesel - - - - - - 7

2.2 Chemical Foundations of Biodiesel-Making - - - 7

2.3 Biodiesel Production: Process Overview - - - 12

2.3.1 Direct Use and Blending - - - - - - 12

2.3.2 Microemulsion - - - - - - - 13

2.3.3 Thermal Cracking (Pyrolysis) - - - - - 15

2.3.4 Transesterification - - - - - - 16

2.3.5 Alcohols - - - - - - - 17

2.3.6 Esterification - - - - - - - 17

2.3.7 Aminolysis - - - - - - - 18

2.3.8 Biocatalyst - - - - - - - 18

Page 7: LATEEF, FATAI ABIOLA

vii

2.4 Oil Extraction and Expression Methods - - - - 19

2.5 Oil Purification Methods - - - - - - 20

2.6 Lipid Chemistry - - - - - - 22

2.6.1 Fats and Oils - - - - - - 23

2.6.2 Transesterification of Glycerides - - - - 24

2.6.3 Fatty Acids - - - - - - - 25

2.7 Physical and Chemical Properties of Lipids - - - 28

2.7.1 Physical Properties - - - - - - 28

2.7.2 Chemical Properties of Fats and Oil - - - - 29

2.8 Rancidity - - - - - - - 31

2.9 Storage and Oxidative Stability of Fatty acid Esters - 32

2.10 Additization of Esters - - - - - - 34

2.11 Rheological Properties of Fluids - - - - - 35

2.11.1 Biodiesel Rheology: Fluid flow phenomena - - 35

2.11.2 Viscosity - - - - - - 36

2.11.3 Kinematic Viscosity - - - - - - 36

2.12 Biodiesel Specifications and Properties - - - - 37

2.13 Biodiesel Production Process Options - - - - 39

2.13.1 Batch Processing - - - - - - - 40

2.13.2 Continuous Process Systems - - - - - 43

2.14 Spectroscopy for Biofuel Analysis - - - - 43

2.14.1 Use of Molecular Spectra as Aids in the Identification of

Organic Structures - - - - - - 43

2.14.2 Infrared Spectra - - - - - - - 45

CHAPTER THREE

RESEARCH METHODOLOGY

3.1 Raw materials - - - - - - - 46

3.2 Apparatus - - - - - - - 46

3.3 Extraction of Oil from Various Seeds and Solvent Recovery 46

Page 8: LATEEF, FATAI ABIOLA

viii

3.3.1 Extraction of Oil from Castor Seed - - - - 46

3.3.2 Extraction of Oils from Palm-Kernel Seed - - - 47

3.3.3 Extraction oil Oils from Groundnut Seed - - - 47

3.4 Pre-treatment of Oils - - - - - - 47

3.5 Determination of the Optimal Catalyst Weight - - 48

3.6 Determination of Kinematic Viscosity of Crude, Kinetics

Biodiesel Samples - - - - - - 49

3.7 Alkali-Catalyzed Batch Production of Biodiesel - - 50

3.8 Characterization of the Crude, Refined Oils, Biodiesel and

Petrodiesel - - - - - - - 51

3.8.1 Saponification Value (SV) Determination - - - 51

3.8.2 Acid Value (AV) Determination - - - - - 51

3.8.3 Iodine Value (IV) Determination - - - - - 51

3.8.4 Specific Gravity Determination - - - - - 52

3.8.5 Peroxide Value Determination - - - - - 52

3.8.6 Calorific (heating/combustion) Value Determination using

Bomb Calorimeter - - - - - - 53

3.9 Investigation of Temperature Dependence of Biodiesel Kinematic

Viscosity and Specific Gravity / Biodiesel-Petrodiesel Blending 54

3.10 Transesterification Kinetics in a Batch Reactor - - 55

3.11 Infrared Spectroscopy-Biodiesel Product Analysis Procedure 56

CHAPTER FOUR

RESULTS AND DISCUSSION

4.1 Extraction of Crude Plant Oils - - - - - 58

4.2 Pre-treatment of Oils - - - - - - 58

4.3 Optimal Catalyst Weight Test - - - - - 59

4.4 Properties of Crude, Refined and Biodiesel Obtained - 61

4.4.1 Saponification Value - - - - - - 61

4.4.2 Iodine Value - - - - - - - 62

4.4.3 Acid Value - - - - - - - 62

Page 9: LATEEF, FATAI ABIOLA

ix

4.5 Physical Characterization of Castor, Palm Kernel and Groundnut

Oils Biodiesel in Comparison with Conventional Diesel

(Petrodiesel) - - - - - - - 62

4.5.1 Specific Gravity - - - - - - - 63

4.5.2 Calorific Value - - - - - - - 63

4.5.3 Kinematic Viscosity - - - - - - 64

4.6 Alkali-Catalyzed Batch Production of Biodiesel at 65oC - 64

4.7 Kinematic Viscosity of Biodiesel at Different Temperatures 66

4.8 Castor Oil Biodiesel/Petrodiesel Blending - - - 72

4.9 Kinetic Studies Result at Various Temperatures - - 73

4.10 Biodiesel Storage/Oxidative Stability Measurement - 76

4.11 Infrared Spectroscopy-Biodiesel Product Analysis Result - 78

CHAPTER FIVE

CONCLUSION AND RECOMMENDATIONS

5.1 Conclusion - - - - - - - - 85

5.2 Recommendation - - - - - - - 86

References - - - - - - - - 88

Appendices

Page 10: LATEEF, FATAI ABIOLA

x

LIST OF FIGURES

Fig. 1.1: Transesterification Reaction of Triglycerides - - 3

Figure 2.1: Molecule Structure of an Idealized Fatty Acid - 8

Figure 2.2: Molecular Structure of Soap - - - - 8

Figure 2.3: Molecular Structure of Glycerol - - - - 8

Figure 2.4: Molecular structure of methanol, ethanol, 1–propanol, and

1–butanol - - - - - - - - 9

Figure 2.5: Biodiesel molecules (a) methyl ester (b) propyl ester - 9

Figure 2.6: Molecular structure of triglycerides - - - - 10

Figure 2.7: Molecular structures of octadecanoic acid (a) cetane

(hexadecane) molecule (b) ethyl ester 10

Fig. 2.8: Transesterification of Triglycerides - - - - 16

Figure 2.9: Esterification - - - - - - - 18

Fig 2.10: Aminolysis of triglycerides - - - - - 18

Figure 2.11: Typical Oil Extraction Process - - - - 19

Fig 2.12: Fats and Oils - - - - - - - 24

Figure 2.13: Three consecutive and reversible reactions - - 24

Figure 2.14: Part of the hydrocarbon chain of a saturated fatty acid 25

Figure 2.15: Part of the hydrocarbon chain of a unsaturated fatty acid 25

Figure 2.16: Oxidation Behaviour of Vegetable Oils and Animal Fats 33

Fig 2.17: Butylated Hydroxyanisole (BHA) - - - - 34

Figure 2.18: Antioxidants for oils and fats - - - - 35

Figure 2.19: Batch Reaction Process - - - - - 42

Figure 2.20: Plug Flow Reaction System - - - - 43

Fig. 4.1: Variation of Percentage (%) Yield of Biodiesel with % (w/v)

of Catalyst (CSO, PKO and GNO) - - - - 60

Fig. 4.2: Graph of lnV CSO against 1/T (K-1

) - - - 67

Fig. 4.3: Graph of lnV PKO against 1/T (K-1

) - - - 68

Fig. 4.4: Graph of lnV GNO against 1/T (K-1

) - - - 69

Page 11: LATEEF, FATAI ABIOLA

xi

Fig. 4.5: Graph of Specific Gravity of CSO, PKO and GNO

at different temperatures - - - - 71

Fig 4.6: Graph of Percentage (%) Conversion of biodiesel against

time (min) Kinetic result at 30oC - - - - 74

Fig 4.7: Graph of % Conversion of biodiesel against time (min)

Kinetic result at 45oC - - - - - - 74

Fig 4.8: Graph of Percentage (%) Conversion of biodiesel against

time (min) Kinetic result at 60oC - - - - 75

Fig 4.9: Graph of Percentage (%) Conversion of biodiesel against

time (min) Kinetic result at 75oC - - - - 75

Page 12: LATEEF, FATAI ABIOLA

xii

LIST OF TABLES

Table 1.1: Nigeria Vegetable Oil Profile - - - - 6

Table 1.2: Nigeria’s Palm Kernel Oil Profile - - - - 6

Table 2.1: Known problems, probable cause and potential solutions for

using straight vegetable oil in diesels - - - 14

Table 2.2 Compositional data of pyrolysis of oils - - - 15

Table 2.3: Typical Assay Showing the Percentage of Constituent Fatty

Acids - - - - - - - - 26

Table 2.4: Castor Oil fatty acids - - - - - 27

Table 2.5: Biodiesel Chain Length - - - - - 27

Table 2.6: Some common fatty acid - - - - - 28

Table 2.7: Iodine values of various groups of oils and fats - 30

Table 2.8: ASTM Biodiesel Standard D 6751a - - - 38

Table 2.9: European Biodiesel Standards EN 14214 for Vehicle Use

and EN 14213 for Heating Oil Use - - - 39

Table 2.10: Kinematic Viscosity of Oils - - - - 40

Table 4.1: Yield of Oils - - - - - - 58

Table 4.2: Percentage loss on Pretreatment - - - - 58

Table 4.3 Result for the Characterization of Crude, Refined and

Biodiesel - - - - - - - 61

Table 4.4 Physical Characterization of Biodiesel and Petrodiesel 63

Table 4.5 Results of Alkali-Catalyzed Batch Production of Biodiesel

at 65oC - - - - - - - 64

Table 4.6: Variation of Kinematic Viscosity of Biodiesel with

Temperature - - - - - - - 66

Table 4.7: Simplification of Andrade Equation of Biodiesel Sample 66

Table 4:8: Variation of Specific Gravity of Biodiesel Samples with

Temperatures - - - - - - 70

Table 4.9: Comparison of Kinematic viscosity and Specific Gravity

of Castor oil Biodiesel and Petrodiesel Blend with

Unblended Biodiesel - - - - - 73

Table 4.10: Peroxide Values of Castor, Palm kernel, Groundnut Oils

and Petrodiesel - - - - - - 76

Page 13: LATEEF, FATAI ABIOLA

1

CHAPTER ONE

INTRODUCTION

1.1 Preamble

The high energy demand, pollution problems, as well as global

consensus that fossil energy sources are finite, make it increasingly

necessary to develop the renewable energy source of limitless duration,

smaller environmental impacts, technically feasible and readily available

(Gupta et al, 2008; Meher et al., 2004; Alamu et al, 2008; Bamgboye and

Hansen, 2007; Fukuda et al, Sharma et al, 2008). Biodiesel have been

reported as a promising long-tern renewable energy source (Tapasvi et al,

2005).

Biodiesel is pursued not only for the consideration of the future

shortage of petroleum supplies but also for the well being of the environment

(Zhao et al, 2007; Ma and Hanna, 1999). Diesel fuels have an essential

function in the industrial economy of a developing country and used for

transport of industrial and agricultural goods and operation of diesel tractor.

Various plant oils have been converted into biodiesel and they work well in

diesel engines (Pimentel and Patrick, 2005). One possible alternative to

fossil fuel is the use of oil of plant origin like vegetable oil (Wenzel et al,

2006; Meher et al, 2004; Gerpen, 2005; Noureddini and Zhu, 1997) waste fat

and oil (Refaat et al, 2008, Gupta et al, 2007; Zhang et al, 2003); although

other crops such as mustard, hemp, jatropha and even algae show great

potential as a source of raw materials for biodiesel production (Baroutian et

al, 2008; Chhetri et al, 2008).

Available statistical data ranked Nigeria as one of the twenty largest

oil producers in the world with a whooping 36.220 million barrels of

proven oil reserves, 184.160 Trillion cubic feet natural gas reserves at the

beginning of year 2009 and end of the year 2007 respectively (EIA, 2009).

Page 14: LATEEF, FATAI ABIOLA

2

Nigeria is the eight largest producer of oil and nineth largest producer of

natural gas. Fossil fuels account for over ninety percent of its revenue

generation (NNPC Newsletter, 2009). This amazing oil wealth

notwithstanding, the Energy Commission of Nigeria (ECN) expressed fears

over future depletion of these fossil fuels and its severe environmental

impacts (Alamu et al, 2007a).

Petroleum-based energy sources pose severe threats to the

environment from hazardous emissions. Huge level of fossil fuel combustion

have resulted in the concentration of carbondioxide which causes dramatic

global climate change, air pollution (Kyu-Wan et al, 2007).

Biodiesel, an environmental friendly fuel, has many merits. It is

derived from renewable, domestic resource (thereby relieving reliance on

petroleum fuel imports), biodegradable and non-toxic (Zhang et al, 2003;

Gerpen, 2004; Canakci and Gerpen, 2001). Compared to petroleum-based

diesel, biodiesel has a more favourable combustion emission profile, such as

low emissions of carbon monoxide, particulate emission and unburned

hydrocarbons. Biodiesel has a relatively high flash point, lubricating

properties that reduce engine wear and extend engine life (Zhang et al, 2003,

Gerpen, 2004; Alamu et al, 2007a).

Plant-oils occupy a prominent position in the development of

alternative fuels although, there have been many problems associated with

using it directly in diesel engine (especially in direct ignition engine). These

include carbon deposits, oil ring stickening, thickening of lubricants and high

viscosity (Knothe et al, 2005; Krisnangkura et al, 2005; Knothe, 2001;

Alamu et al, 2008).

Biodiesel is produced by transesterifying the parent oil or fat with an

alcohol (Knothe, 2006; Darnoko et al, 2000). Alcohols such as methanol,

ethanol, 1-propanol and butanol have been used for biodiesel production.

Using alcohol of higher molecular weight improves the cold flow properties

Page 15: LATEEF, FATAI ABIOLA

3

of the resulting ester, at the cost of a less efficient transesterification reaction

(Wikipedia, the free encyclopedia: Biodiesel). Alkaline catalyst such as

sodium hydroxide (NaOH) and potassium hydroxide (KOH) are the most

commonly used in transesterification, since their reaction is much faster than

an acid-catalyzed reaction (Titipong, 2006).

Two approaches for transesterification of plant-oils for production of

biodiesel are suggested: chemical one using alkali catalyst (NaOH, KOH or

alkoxides) or acid catalyst (strong acids H2SO4, H3PO4) (Jitputti et al, 2004;

Shah et al, 2004; Zhang et al, 2003). The product of the reaction is a mixture

of esters, which is known as biodiesel and glycerol, which is a high value co-

product. The second approach is enzymatic one, in which lipase-catalyzed

transesterification is carried out in non-aqueous environment.

H2C – OCOR1 ROCOR1 H2C – OH

H2C – OCOR2 + ROH ROCOR2 + H C – OH

H2C – OCOR3 ROCOR3 H2C – OH

Triglyceride Alcohol Mixtures of Glycerol

Alkyl ester

Fig. 1.1: Transesterification reaction of triglycerides

R1, R2 and R3 are long hydrocarbon chains, sometimes called fatty

acid chains.

Triglycerides, as the main component of plant oil, consist of three long

chain fatty acids esterified to a glycerol backbone. When triglycerides react

with an alcohol (e.g 1-propanol), the three fatty acid chains are released from

the glycerol skeleton and combine with alcohol to yield fatty acid alkyl

esters. Glycerol is produced as a by-product. The transesterification reaction

is a reversible reaction and the transformation occurs essentially by mixing

the reactants. However, the presence of a catalyst considerably accelerates

Catalyst

Page 16: LATEEF, FATAI ABIOLA

4

the adjustment of the equilibrium (Ma and Hanna, 1999). Stoichiometric

reaction requires 1mol of a triglycerides and 3mol of alcohol. However an

excess of alcohol is used to increase the yields of the alkyl esters and to

allow its phase separation to be formed (Schuchardt et al, 1998).

There are two major methods available for biodiesel production: batch

transesterification processes and continuous process (Leevijit et al, 2004).

The batch reactor has the advantage of high conversions that can be obtained

by leaving the reactant in the reactor for long periods of time (Fogler, 2006).

1.2 JUSTIFICATION OF STUDY

Various plant oils have been converted into biodiesel and they work

well in diesel engines. Several researchers have used biodiesel as alternative

fuel in the existing compression engine (CE) without any modification.

Promising results have been obtained by running CE on plant oil based

biodiesels: soybean (US), rapeseed (Europe), oil palm (South-East Asia) e.t.c.

which are commercialized in these countries (Alamu, 2007b). However,

Nigeria, having the greatest potential in this area, because of the availability

of these raw materials for the production of biorenewable and

environmentally friendly fuel (biodiesel), is yet to make remarkable impact

on its production and usage.

Plant oil is an important renewable feedstock in the long-term (2016–

2025) vision of providing secure, abundant, cost effective and clean source

of energy for Nigeria. Common plant oil in the country includes palm oil,

palm kernel oil, groundnut oil, cottonseed, soybean bean oil etc. Profile for

these oils is presented in Tables 1.1 and 1. 2.

Nigeria is rated to be the second world’s largest producer of palm-

kernel oil and groundnut oil after Malaysia and India respectively (Asiedu,

1989). It was also reported that castor oil plant originates in Africa. However,

industrial use of these plants oil has been limited to soap, detergent,

Page 17: LATEEF, FATAI ABIOLA

5

lubricants, paints e.t.c. This shows that despite the abundance of these plant

oils, it has been underutilized in Nigeria. Successful reports on

transesterification of some Nigerian oils in the preparation of biodiesel is an

indication of better industrial utilization of these plant oils in Nigeria, as

considerable research efforts are now focusing on this alternative diesel fuel

worldwide (Knothe, 1999).

1.3 AIM AND OBJECTIVES OF RESEARCH

This project is aimed at the production of biodiesel using various plant

oils: castor oil, palm kernel and groundnut oils on 6:1 propanol-oil molar

ratio and optimal weight of sodium hydroxide as the catalyst.

• To carryout physical and chemical characterization on the biodiesel

produced.

• To determine the optimal catalyst weight for the transesterification

reaction.

• To evaluate biodiesel kinematic viscosity at various temperatures.

• To investigate how biodiesel production reaction proceeds at different

time (Reaction Kinetics).

• To determine the oxidative stability of the biodiesel produced.

• To analyze the biodiesel sample by Fourier Series Infra-red

Spectroscopy.

• To make recommendations based on the findings from the study.

Page 18: LATEEF, FATAI ABIOLA

6

Table 1.1: Nigeria Vegetable Oil Profile (2006)

Commodity Quantity (tons) Percentage Share

Palm oil 800,000 50

Palm kernel oil 270,000 17

Others: peanuts, cottonseed, soybean 260,000 16

Imports 270,000 17

National requirement 1,600,000 100

Source: Alamu et al, 2007a

Table 1.2: Nigeria’s Palm Kernel Oil Profile

Commodity USDA Revised

Estimate (2004)

USDA Revised

Estimate (2005)

USDA Revised

Estimate (2006)

(1000 tons)

Beginning stock 10 10 10

Production 272 272 275

Imports 1 1 1

Total supply 283 283 286

Exports 1 1 1

Industrial domestic

consumption

86 86 86

Food use domestic

consumption

186 186 189

Total domestic

Consumption

272 272 275

Source: Alamu et al, 2007a

Page 19: LATEEF, FATAI ABIOLA

7

CHAPTER TWO

LITERATURE REVIEW

2.1 HISTORY OF BIODIESEL

The diesel engine was invented by Otto Diesel in 1982. His engine

was designed to run on a wide variety of fuels. Although he demonstrated a

diesel engine running on peanut oil at the Paris exhibition of 1900, the first

commercial diesel engines ran on kerosene. The new engine first appeared in

transportation vehicles in ships in the 1900’s. They showed up in trains in

1914, but “did not seriously displace the steam engine until after World

War”. Diesels were first used in automobile in 1924 (Turner, 2005).

On August 31,1937, G. Chavanne of the University of Brussels

(Belgium) was granted a patent for a procedure for the transformation of

vegetable oils for their uses as fuels (Belgian Patent 422,877). This patent

described the alcoholysis (often referred to as transesterification) of

vegetable oils using ethanol (an alcohol) in order to separate the fatty acids

from the glycerol by replacing the glycerol with short linear alcohols. This

appears to be the first account of the production of what is known as

‘biodiesel’ today (Wikipedia, the free encyclopedia, 2008).

2.2 Chemical Foundations of Biodiesel-making

Biodiesel is a renewable, biodegradable, environmentally benign fuel

use in diesel engine (Titipong, 2006). Either virgin vegetable oil or waste

vegetable oil (wvo) can be used to make quality fuel. Fats are converted to

biodiesel through a chemical reaction involving alcohol and a catalyst.

It is instructive to think of the chemistry of biodiesel in terms of

building blocks that comprise the larger molecules involved in the biodiesel-

making reactions. Fatty acids are a component of both vegetable oil and

biodiesel. In chemical terms, they are carboxylic acids of the form:

Page 20: LATEEF, FATAI ABIOLA

8

H2 H2 H2 H2 H2 H2 H2 H2 O C C C C C C C C H3C C C C C C C C C OH H2 H2 H2 H2 H2 H2 H2 H2

Figure 2.1: Molecule structure of an idealized fatty acid

Fatty acids which are not bound to some other molecule are known as

free fatty acids. When reacted with base, a fatty acid loses a hydrogen atom

to form soap.

H2 H2 H2 H2 H2 H2 H2 H2 O C C C C C C C C H3C C C C C C C C C O H2 H2 H2 H2 H2 H2 H2 H2

Figure 2.2: Molecular structure of soap

Chemically, soap is the salt of a fatty acid. The structures of the fatty

acids shown in this section are highly idealized. Real fatty acids vary in the

number of carbon atoms, and in the number of double bonds. Glycerol, a

component of vegetable oil and a by-product of biodiesel production, has the

following form:

HO CH2

HO CH

CH2 HO

Figure 2.3: Molecular structure of glycerol

Alcohols are organic compound of the form R–OH, where R is a

hydrocarbon. Typical alcohols used in biodiesel-making are methanol,

ethanol, 1-propanol, and 1-butanol:

Page 21: LATEEF, FATAI ABIOLA

9

H2 H2 H2 H2

H3C–OH C C OH C C

H3C OH H3C C H3C C OH H2 H2

Figure 2.4: Molecular structure of methanol, ethanol, 1–propanol, and 1–

butanol

Transesterification is sometimes called alcoholysis, or if by a specific

alcohol, by corresponding names such as methanolysis, or ethanolysis or

propanolysis. Chemically, biodiesel is a fatty acid alkyl ester:

H2 H2 H2 H2 H2 H2 H2 H2 O H3 C C C C C C C C C _ (a) H3C C C C C C C C C O H2 H2 H2 H2 H2 H2 H2 H2

H2 H2 H2 H2 H2 H2 H2 H2 O H2 H3 C C C C C C C C C C (b) H3C C C C C C C C C O C H2 H2 H2 H2 H2 H2 H2 H2 H2

Figure 2.5: Biodiesel molecules (a) methyl ester (b) propyl ester

The biodiesel ester contains a fatty acid chain on one side, and a

hydrocarbon called an alkane on the other. Thus, biodiesel is a fatty acid

alkyl ester. Usually, the form of the alkane is specified as in “methyl ester”,

“ethyl ester” or “propyl ester”.

Vegetable oil is a mixture of many compounds, primarily triglycerides

and free fatty acids. Triglyceride is a tri-ester of glycerol and three fatty

acids.

Page 22: LATEEF, FATAI ABIOLA

10

H2 H2 H2 H2 H2 H2 H2 H2 O C C C C C C C C H3C C C C C C C C C O H2 H2 H2 H2 H2 H2 H2 H2

H2 H2 H2 H2 H2 H2 H2 H2 O C C C C C C C C CH _ H3C C C C C C C C C O H2 H2 H2 H2 H2 H2 H2 H2

H2 H2 H2 H2 H2 H2 H2 H2 O C C C C C C C C CH2 _ H3C C C C C C C C C O H2 H2 H2 H2 H2 H2 H2 H2

Figure 2.6: Molecular structure of triglycerides

Virgin oil contains a low percentage of free fatty acids. Waste

vegetable oil contains a higher amount of FFA’s because the frying process

breaks down triglycerides molecules (Turner, 2005).

Petroleum diesel and biodiesel are both mixtures of organic

compounds. The idealized petroleum molecule is cetane, a pure paraffin.

Compared to cetane, alkyl esters are somewhat longer, and more importantly,

contain two oxygen atoms.

H2 H2 H2 H2 H2 H2 H2 C C C C C C C CH3 (a) C C C C C C C C H3 H2 H2 H2 H2 H2 H2 H2

H2 H2 H2 H2 H2 H2 H2 H2 O H2 C C C C C C C C C (b) H3C C C C C C C C C O CH3 H2 H2 H2 H2 H2 H2 H2 H2

Figure 2.7: (a) cetane (hexadecane) molecule

(b) ethyl ester of octadecanoic acid

Since combustion is an oxidation reaction, the heating value of cetane,

which contains no oxygen atoms, is higher than that of biodiesel. For this

reason, diesel engines running biodiesel, experience a loss of power on the

order of 5% (Turner, 2005).

CH2

Page 23: LATEEF, FATAI ABIOLA

11

The principal ways of making biodiesel are by transesterification of

triglycerides and esterification of free fatty acids.

The esterification reaction of triglycerides is as follows:

Tryglyceride + 3Alcohol 3Biodiesel + Glycerol

The esterification reaction of free fatty acids:

FFA + Alcohol Biodiesel + Water

In the first reaction, a tri-ester is converted to three individual esters,

thus the term transesterification. In the second reaction, a new ester is

created, thus it is called esterification.

Transesterification reaction can be base-catalyzed, acid catalyzed, or

enzymatic. The base-catalyzed reaction takes about one hour or more at

room temperature (Gerpen et al, 2004; Ma & Hanna 1999). It suffers from

competing saponification reactions, which convert the same ingredients as

well as any free fatty acid to soap. Acid-catalyzed and enzymatic

transesterification require three to four days to complete (Turner, 2005). The

acid-catalyzed reaction also requires heat.

There is no competing saponification reaction with acid-catalyzed and

enzymatic reactions. In fact, even free fatty acids are converted to biodiesel

by esterification.

The associated acid-catalyzed esterification reaction requires only

about two hours to completion (Turner, 2005). A combined strategy called

the two-stage process can be used to maximize the amount of biodiesel

produced, while minimizing the amount of soap produced (Turner, 2005).

The first stage is acid-catalyzed esterification of the free fatty acids. This is

followed by base-catalyzed transesterification. This approach is especially

effective for waste vegetable oil and animal fats, which have high free fatty

acid content.

This thesis only deals with base-catalyzed transesterification.

acid catalyst

catalyst

Page 24: LATEEF, FATAI ABIOLA

12

2.3 BIODIESEL PRODUCTION: PROCESS OVERVIEW

There are several generally accepted ways to make biodiesel, some

more common than others e.g. transesterification and blending, and several

others that are more recent developments e.g. reaction with supercritical

methanol. An overview of these processes is as follows:

1. Direct use and Blending, which is the use of pure vegetable oils or the

blending with diesel fuel in various ratios,

2. Micoremulsions with simple alcohols,

3. Thermal cracking (pyrolysis) to alkanes, alkenes e.t.c.,

4. Transesterification (alcoholysis) which consists of several sub

categories;

(i) Esterification

(ii) Aminolysis

5. Other forms of catalysis

(i) Biocatalysis

(ii) Catalyst free

2.3.1 Direct Use and Blending

The direct use of vegetable oil (palm kernel oil, groundnut oil e.t.c) in

diesel engines is problematic and has many inherent failings. It has only

been researched extensively for the past couple of decades, but has been

experimented with for almost a hundred years. Although some diesel engines

can run pure vegetable oils, turbocharged direct injection engines such as

trucks are prone to many problems (Khan, 2002). Direct use has not been

satisfactory because of viscous nature of vegetable oil. However, methyl,

ethyl, propyl and butyl esters of fatty acids present in oils have proved

promising enough to be called biodiesel (Shah et al, 2003).

Page 25: LATEEF, FATAI ABIOLA

13

Neat vegetable oils (such as palm oil, groundnut oil e.t.c) were

primarily considered as alternatives for diesel fuel but their very high

viscosity, at room temperature made them unsuitable in diesel engine

(Krisnangkura et al, 2005). Even after heating to around 80oC, it is still six

times as viscous as diesel. This leads to problems with flow of oils from the

fuel tank to the engine, blockages in filters and subsequent engine power

losses. Even if preheating is used to lower the viscosity, difficulties may still

be encountered with starting due to the temperatures required for oils to give

off ignitable vapour. Further, engines can suffer coking and gumming which

leads to sticking of piston rings due to multibonded compounds undergoing

pyrolysis. Polyunsaturated fatty acids also undergo oxidation in storage

causing gum formation and at high temperatures where complex oxidative

and thermal polymerization can occur (Ma and Hanna, 1999). See table 2.1.

2.3.2 Microemulsion

A microemulsion is designed to tackle the problem of the high

viscosity of pure vegetable oils by reducing the viscosity of oils with

solvents such as simple alcohols. Microemulsions are defined as colloidal

equilibrium dispersions of optically isotropic fluid microstructures, with

dimensions generally in 1–150nm range formed spontaneously from two

normally immiscible liquids and one or more ionic or non-ionic compounds

(Khan, 2002).

Page 26: LATEEF, FATAI ABIOLA

14

Table 2.1

Known problems, probable cause and potential solutions for using straight vegetable oil in diesels (Ma and Hanna,1999).

Problem Probable cause Potential solution

Short-term

1. Cold weather starting

High viscosity, low cetane, and low flash point

of vegetable oils

Preheat fuel prior to injection. Chemically alter fuel

to an ester

2. Plugging and gumming of filters,

lines and injectors

Natural gums (phosphatides) in vegetable oil.

Other ash

Partially refine the oil to remove gums. Filter to

4-microns

3. Engine knocking

Very low cetane of some oils. Improper injection

timing.

Adjust injection timing. Use higher compression

engines. Preheat fuel prior to injection. Chemically

alter fuel to an ester

Long-term

4. Coking of injectors on piston

and head of engine

High viscosity of vegetable oil, incomplete

combustion of fuel. Poor combustion at part

load with vegetable oils

Heat fuel prior to injection. Switch engine to diesel

fuel when operation at part load. Chemically alter

the vegetable oil to an ester

5. Carbon deposits on piston

and head of engine

High viscosity of vegetable oil, incomplete

combustion of fuel. Poor combustion at part

load with vegetable oils

Heat fuel prior to injection. Switch engine to diesel

fuel when operation at part load. Chemically alter

the vegetable oil to an ester

6. Excessive engine wear

High viscosity of vegetable oil, incomplete

combustion of fuel. Poor combustion at part

load with vegetable oils. Possibly free fatty acids

in vegetable oil. Dilution of engine lubricating

oil due to blow-by of vegetable oil

Heat fuel prior to injection. Switch engine to diesel

fuel when operation at part load. Chemically alter

the vegetable oil to an ester. Increase motor oil

changes. Motor oil additives to inhibit oxidation

7. Failure of engine lubricating

oil due to polymerization

Collection of polyunsaturated vegetable oil

blow-by in crankcase to the point where

polymerization occurs

Heat fuel prior to injection. Switch engine to diesel

fuel when operation at part load. Chemically alter

the vegetable oil to an ester. Increase motor oil

changes. Motor oil additives to inhibit oxidation.

14

Page 27: LATEEF, FATAI ABIOLA

15

Table 2.2: Compositional data of pyrolysis of oils (Ma and Hanna, 1999)

Percent by weight

High Oil safflower Soybean

Alkanes 40.9 29.9

Alkenes 22.0 24.9

Alkadienes 13.0 10.9

Aromatics 2.2 1.9

Unresolved unsaturates 10.1 5.1

Carboxylic acids 16.1 9.6

Unidentified 12.7 12.6

The performances of ionic and non-ionic microemulsions were found

to be similar to diesel fuel, over short term testing. They also achieved good

spray characteristics, with explosive vapourization which improved the

combustion characteristics (Ma and Hanna, 1999). In longer-term testing, no

significant deterioration in performance was observed, however significant

injector needle sticking, carbon deposits, incomplete combustion and

increasing viscosity of lubricating oils were reported (Ma and Hanna, 1999).

2.3.3 Thermal Cracking (Pyrolysis)

Pyrolysis is defined as the conversion of one substance into another by

means of heat or by heat with the aid of a catalyst (Ma and Hanna, 1999).

The pyrolysis of fats has been investigated for more than 100 years,

especially in countries where there is shortage of petroleum deposits. Typical

catalysts that can be employed in pyrolysis are SiO2 and Al2O3 (Khan, 2002).

Unlike direct blending, fats can be pyrolysed successfully to produce many

smaller chain compounds. Typical breakdown of compounds found from

pyrolysis of safflower and soybean oil are listed in Table 2.2 above.

Page 28: LATEEF, FATAI ABIOLA

16

2.3.4 Transesterification

Transesterification of triacylglycerols (triglycerides) is a simple

process that converts vegetable oils into fuel for diesel engine (Warzyniak et

al, 2005). The stoichiometry for the reaction is 3:1 alcohol to oils (lipids),

however in practice this is usually

H2C – OCOR1 ROCOR' H2C – OH

H2C – OCOR2 + 3R'OH ROCOR' + H C – OH

H2C – OCOR3 ROCOR' H2C – OH

Triglyceride Alcohol Esters Glycerol

Fig. 2.8: Transesterification of Triglycerides

The raw materials for biodiesel production are: vegetable oils and

animal fat, alcohols and catalysts.

Oils and fats are essentially esters of glycerol and fatty acids, derived

from plant and animal sources. Different oils have different acid

compositions and hence different viscosities. Oils are normally liquid at

ambient temperature, fats are normally solid (Lewis, 1987). Vegetable oil is

an important renewable feedstock in the long term (2016 – 2025) vision of

providing secure, abundant, cost effective and clean source of energy for

Nigeria (Alamu et al, 2007a). Common vegetable oil in the country includes

palm oil, palm kernel oil, groundnut oil, peanuts, cottonseed and soybean.

Animal fat such as tallow, lard, chicken fats are useful in biodiesel

production.

Catalysts are necessary to reduce the time and energy requirement for

the transesterification reaction. The catalyst lowers the activation energy of

the reaction by providing an alternative path that avoids the slow, rate-

determining step of the uncatalyzed reaction (Atkins and Paula, 2002).

Catalysts are classified as alkali, acid or enzyme. Alkali-catalyzed

transesterification is much faster than acid catalyzed. However, if a glyceride

Catalyst

Page 29: LATEEF, FATAI ABIOLA

17

has a higher free fatty acid content and more water, acid-catalyzed

transesterification is suitable. The acids could be sulfuric acid, phosphoric

acid, hydrochloric acid or organic sulfonic acid. Alkali includes sodium

hydroxide, sodium methoxide, potassium hydroxide, potassium methoxide

e.t.c. Sodium hydroxide is cheaper and also chosen because is used widely in

large-scale processing (Ma and Hanna, 1999).

Transesterification will occur without the aid of catalyst, however at

temperatures below 300oC the rate is very slow. It has been said that there

are, from a broad perspective, two methods to producing biodiesel and that is

with and without a catalyst (Khan, 2002).

2.3.5 Alcohols

Aliphatic monohydric alcohols are monohydroxyl derivatives of

alkanes and have a general molecular formula CnH2n+1OH, or simply ROH

(Murray, 1997). Generally, in biodiesel production a large excess of alcohol

is used to shift the equilibrium to the right (Zhang et al, 2003). The most

commonly used primary alcohol in biodiesel production is methanol,

although other alcohols such as ethanol, propanols, butanols and amyl

alcohol can also be used. The major factors that influence the choice of

alcohol for transesterification are: cost of the alcohol, amount of the alcohol

needed for the reaction, ease of recovery, and recycling of the alcohol and

fuel tax incentives.

2.3.6 Esterification

The formation of esters occurs through a condensation reaction known

as esterification. This requires two reactants, carboxylic acids (fatty acid)

and alcohols (Khan, 2002). Esterification reactions are acids catalyzed and

proceed slowly in the absence of strong acids such as sulphuric acid,

Page 30: LATEEF, FATAI ABIOLA

18

phosphoric acid, sulfonic acids and hydrochloric acid. The equation for the

esterification reaction is shown below.

R– C – OH + R'OH R – C – OR' + H2O

Free fatty acid Alcohol Esters Water

Figure 2.9: Esterification

2.3.7 Aminolysis

Esters undergo nucleophilic substitution at their acyl carbon atoms

when they are treated with primary or secondary amines. These reactions are

slow but are synthetically useful (Khan, 2002).

O

R1 – C – N – R

5

H2C – COO – R1 R4 O H2C – OH

H2C – COO – R2 + 3 H – N – R3 R

2 – C – N – R

5 + H2C – OH

H2C – COO – R3 O H2C – OH

R3 – C – N – R

5

Triglycerides Amine Amides Glycerol

Fig 2.10: Aminolysis of triglycerides

2.3.8 Biocatalyst

Biocatalysts are usually lipases; however conditions need to be well

controlled to maintain the activity of the catalyst (Khan, 2002). Hydrolytic

enzymes are generally used as catalysts as they are readily available and are

easily handled. They are stable, do not require co-enzymes and will often

tolerate organic solvent. The second approach in producing biodiesel is

enzymatic one, in which lipase-catalyzed transesterification is carried out in

nonaqeous environments. The main problem of enzyme catalyzed process is

the high cost of the lipases used as catalyst (Royon et al, 2006). Chemical

transformation is efficient in terms of reaction. Though, biocatalysts allow

H+

heat

Page 31: LATEEF, FATAI ABIOLA

19

synthesis of specific alkyl esters, easy recovery of glycerol, and

transesterification of glycerides with high free fatty acid content (Shah et al,

2003) but recent patents and articles have shown that reaction yields and

times are still unfavourable compared to base-catalyzed transesterification

for commercial application.

2.4 OIL EXTRACTION AND EXPRESSION METHODS

Fats and oils are derived from oil seed and animal sources. In order to

get high quality oil, various techniques are used in the processing. The

process of obtaining oil from seeds involves the separation of oil from oil-

bearing material by mechanical means, chemical means e.t.c.

There are generally three broad techniques or methods of oil

extraction – Distillation, Solvent extraction and Expression (Eckhardt and

Kawaguchi, 1997).

Distillation:

Most essential oils are extracted using steam distillation. As the steam

break down the plant, its essential oils are released in a vapourized form.

When these pass through cooling tanks, the volatile essential oil return to

liquid form and are separated and is easily isolated as pure essential plant oil.

Figure 2.11: Typical Oil Extraction Process

Whole seed Cleaning Mechanical press

Purification processing Crude oil Seed meal

Usable oil

Page 32: LATEEF, FATAI ABIOLA

20

Solvent Extraction

It is very common in large-scale operations to remove the oil from

cracked seeds at low temperature with a non-toxic fat solvent such as hexane.

When a plant contains very little oil or when the odorous properties of

flavour and plant materials would be destroyed or altered by steam or water

distillation, solvent extraction is used. The solvent is percolated through the

plant (seed) material in order to produce concrete. The plants (seeds)

constituents including essential oils, fatty acids and waxes are dissolved by a

solvent. After the solvent is distilled off, the remaining constituents make up

the concrete. Alcohol is used to extract the essential oil from the other

constituents. Since the waxes and fatty acids are not alcohol soluble, they are

separated. The alcohol is then released through secondary distillation,

leaving the absolute oil behind.

Expression

Expression, also known as cold pressing, can also be used for the

extraction of oil. Various types of mechanical presses are used to squeeze oil

from oil seeds. The oil-bearing material is squeezed through a tapering outlet

in the mechanical pressing and filtered to get pure essential oil.

2.5 OIL PURIFICATION METHODS

Crude fats and oils are obtained directly from the extraction of the oil

seed. Crude fats and oils contain varying substances that may influence

undesirable flavour, colour, or quality. These substances are removed

through a series of processing steps. The purification processing can be

divided into seven types.

1. Degumming:

Most oils undergo treatment known as degumming. The bulk of

certain phosphatides such as lecithin are separated through this operation.

Page 33: LATEEF, FATAI ABIOLA

21

The processing consists of mixing the oil with water or steam for 30mins.

The gummy residue is dehydrated and the oil is then passed through

centrifugal separators. Larger amounts of water or steam are used to prepare

oil which is more gummed.

2. Refining:

The process of refining reduces the free fatty acid, phospholipids,

carbohydrates, or proteins. The most widely practices form of refining

method is alkali treatment. By treatment of the fats and oils with an alkali

solution, the free fatty acid converts into water soluble soaps. After the alkali

treatment, the fats and oils are washed with water to remove residual water

soluble soaps.

3. Bleaching:

The bleaching process is removing colouring materials such as

chlorophyll and carotene and purifying the fats and oils. The method is by

adsorption of the colour producing substances on an adsorbent material.

Bentonite, silical gel and activated carbon are used as bleaching adsorbent.

4. Deodourization

Deodourization is a vacuum steam distillation process for the purpose

of removing undesirable flavours and odours, mostly arising from oxidation,

in fats and oils. Using steam under reduced pressure, the volatile compounds

are removed from fats and oils. The deodourization utilizes the differences in

volatility between off-flavour and off-odour substances and the triglycerides.

5. Fractionation:

Fractionation is the removal of solids at a given temperature. There are

three kinds of fractionation process such as crystallization, winterization, and

pressing. Crystallization is the widespread technique. A mixture of

triglycerides is separated into different melting points based on solubility at

selected temperature. Next, a small quantity of material is crystallized to

avoid hazes of liquid fractions at refrigeration temperatures, this process is

Page 34: LATEEF, FATAI ABIOLA

22

called winterization. Many oils, including cottonseed and hydrogenated

soyabean oils, are winterized. Pressing process is used to separate liquid oil

from solid fat. The process squeezes or presses the liquid oil. This process is

used to produce hard butter.

6. Hydrogenation

In hydrogenation, hydrogen is added directly to react with unsaturated

(double bond) oil in the presence of nickel catalyst. The need for the

hydrogenation is based on: (1) converting liquid oils to the semi-solid forms

and (2) increasing the oxidation and thermal stability of fats and oils. This

process greatly influences the desired stability and properties of many edible

oil products. The hydrogenation process is easily controlled and can be

stopped at any point. A gradual increase in the melting point of fat and oil is

one of the advantages. If the double bonds are eliminated entirely with

hydrogenation, the product is a hard brittle solid at room temperature.

Margarines are typical examples.

7. Interesterification

Interesterification is a rearrangement or redistribution of the fatty

acids on the glycerol. The fatty acids can be described in random or directed

under some conditions. In addition, this process does not change the degree

of unsaturation or the isomeric state of the fatty acids.

2.6 LIPID CHEMISTRY

LIPIDS

Lipids are group of compounds soluble in organic solvents such as

hexane, benzene, carbontetrachloride, ether e.t.c. but sparingly soluble in

water. They contain carbon, hydrogen and oxygen.

Page 35: LATEEF, FATAI ABIOLA

23

CLASSIFICATION OF LIPIDS

Lipids are classified into:

1. Simple (neutral) lipids

(a) Fats and oils (b) Waxes

2. Compound lipids

(a) Phospholipids (phosphatides)

(i) Lecithins (ii) Cephalins (iii) Sphingomyelins

(b) Glycolipids

(i) Cerebrosides (ii) Sialic acid (iii) Gangliosine

3. Derived lipids

(a) Sterols (b) Bile acids

2.6.1 FATS AND OILS

Fats and oils are esters of fatty acids and glycerol. Glycerol is a

trichydric alcohol, that is, it has 3OH groups. Each OH group of glycerol

reacts with the COOH group of fatty acid to form a molecule of fat and oil.

When one molecule of glycerol reacts with a molecule of fatty acid, a

monoglyceride is formed. When two molecules of fatty acids combine with a

molecule of glycerol, diglyceride is formed. While three molecules of fatty

acids reacting with a glycerol molecule give rise to triglycerides (Onimawo

and Akubor, 2005).

Page 36: LATEEF, FATAI ABIOLA

24

O O

CH2OH HO – C – R1 CH2 – O – C – R1

O O

CH2OH + HO – C – R2 CH2 – O – C – R2

O O

CH2OH HO – C – R3 CH2 – O – C – R3

(1 mole of glycerol) (3 molecules of fatty acids) (1 molecule of

triglycerides)

R1 = R2 = R3 (identical FA)

R1 ≠ R2 ≠ R3 (different FA)

Fig 2.12: Fats and Oils

Fats and oils are principally mixture of triglycerides. It may be

composed of glycerol combined with three different or with three molecules

of the same kind. The simplest type of triglyceride is one in which all three

fatty acids are the same as shown in the scheme above.

2.6.2 TRANSESTERIFICATION OF GLYCERIDES

Transesterification, also called alcoholysis, is the displacement of

alcohol from an ester by another alcohol. Suitable alcohols include methanol,

ethanol, propanol, butanol and amyl alcohol. This process has been widely

used to reduce the viscosity of triglycerides, thereby enhancing the physical

properties of renewable fuels to improve engine performance (Fukuda et al,

2001).

Triglyceride (TG) + R'OH Diglyceride (DG) + R'COOR1

Diglyceride (DG) + R'OH Monoglyceride (MG) + R'COOR2

Monoglyceride (MG) + R'OH Glycerol (GL) + R'COOR3

Figure 2.13: Three consecutive and reversible reactions. R1, R2, R3 and R'

represent alkyl group since naturally occurring fats and oils are mixtures of

different triglycerides, they may contain a number of different fatty acids.

Catalyst

Catalyst

Catalyst

Page 37: LATEEF, FATAI ABIOLA

25

O

– C

OH

2.6.3 FATTY ACIDS

The essential components of lipids are carboxylic acids, known as

fatty acids. The general formula of a fatty acid (alkanoic acid is RCOOH

where R represents a hydrocarbon chain. The carboxyl functional group,

contains both a carbonyl and hydroxyl functional group (Murray, 1985).

They are divided into two groups, that is saturated and unsaturated. In

saturated fatty acids, the hydrocarbon chain is saturated with hydrogen.

O H H H H H H H

HO – C – C – C – C – C – C – C – C – R

H H H H H H H H

Figure 2.14: Part of the hydrocarbon chain of a saturated fatty acid

The end carbon atom to which the oxygen atom and hydroxyl (OH)

radical are attached to form a carboxyl group gives the molecule its acid

name. Saturated fatty acids, are characterized by single bonds. Unsaturated

fatty acid, on the other hand, do not have the hydrogen chain saturated with

hydrogen and has one or more double bond.

O H H H H H H H

HO – C – C – C – C – C – C – C – C – R

H H H H H H H H

Figure 2.15: Part of the hydrocarbon chain of a unsaturated fatty acid

Page 38: LATEEF, FATAI ABIOLA

26

Table 2.3: Typical Assay Showing the Percentage of Constituent Fatty Acids

Page 39: LATEEF, FATAI ABIOLA

27

Table 2.4: Castor Oil Fatty Acids

Average Composition of Castor seed oil/fatty acid chain

Acid name Average percentage Range

Ricinoleic acid 85 – 95%

Oleic acid 6 – 2%

Linoleic acid 5 – 1%

Linoleic acid 1 – 0.5%

Stearic acid 1 – 0.5%

Palmitic acid 1 – 0.5%

Dihydroxystearic acid 0.5 – 0.3%

Other 0.5 – 0.2%

Source: Wikipedia, the free encyclopedia from http://en.wikipedia.org/

wiki/castoroil

Table 2.5: Biodiesel Chain Length

Fatty Acid Composition (%)

Carbon

chain length

Soy oil Castor oil Palm oil Palm

kernel oil

Animal fat

C8 4.4

C10 3.7

C12 48.3

C14 1.1 15.6 3.0

C16 11.0 44.0 7.8 26.0

C18 87.6 97.5 53.8 19.8 70.0

C20

Source: PAC catching the wave – IR Application for Biofuel

Page 40: LATEEF, FATAI ABIOLA

28

Table 2.6: Some common fatty acid

Type Common

name

Systematic

name

Formula No. of

double

bond

Position

of

double

bond

Saturated

Butyric Butanoic C3H7COOH

Caproic Hexanoic C5H11COOH

Caprylic Octanoic C7H15COOH

Capric Decanoic C9H19COOH

Lauric Dodecanoic C11H23COOH

Myristic Tetradecanoic C13H27COOH

Palmitic Hexadecanoic C15H31COOH

Stearic Octadecanoic C17H35COOH

Arachidic Eicosani C19H39COOH

Mono

unsaturated

Oleic 9-

octadecanoic

C17H33COOH 1 9

Poly

unsaturated

Linoleic 9,12-

octadecanoic

C17H31COOH 2 9, 12

Linolenic 9, 12, 15-

octadecanoic

C17H29COOH 3 9, 12, 15

Arachidonic 5,8,11,14-

eicosatetranoic

C19H31COOH 4 5,8,11,14

Source: Onimawo and Akubor (2005)

2.7 PHYSICAL AND CHEMICAL PROPERTIES OF LIPIDS

Some of the important physical and chemical tests that shall be

considered in this work are: refractive index, specific gravity, viscosity,

iodine value, saponification value, peroxide value.

2.7.1 PHYSICAL PROPERTIES

The physical properties of the fats and oils are often used to identify

them. Physical and chemical properties characterization of oils are of

practical importance in understanding of the components of the lipids, and

Page 41: LATEEF, FATAI ABIOLA

29

their determination will usually be sufficient for confirming the identity and

uses of most oils and fat (Meyer, 2004). These properties of lipids depend on

factors such as sources, degree of saturation, length of carbon chains,

molecular structures of the triglycerides and processing method.

Specific Gravity (S.G) of oils

Specific gravity is one of the analytical constant used in identifying

oils and fats. Specific gravity is a measure of the ratio of the weight gravity

or mass gravity of a substance to that of a given standard, usually water.

The specific gravity of oils is related to the degree of unsaturation of

the component fatty acids and their average molecular weight. The higher

their degree of unsaturation and the lower the molecular weight. The specific

gravity of individual oils, however, does not vary widely, it lies between

0.910 – 0.924. The degree of unsaturation of component of fatty acids has

but little effect on specific gravity (Joslyn, 1970).

2.7.2 CHEMICAL PROPERTIES OF FATS AND OIL

IODINE VALUE

The iodine value of oil is defined as the weight of iodine absorbed by

100 parts by weight of the sample. The glycerides of the unsaturated fatty

acids present (particularly of the oleic acid series) unite with a definite

amount of halogen and the iodine value is therefore a measure of the degree

of unsaturation. It is constant for a particular oil or fat, but the exact figure

obtained depends on the particular technique employed. The ranges of

figures for the iodine values of the various groups of oils and fats are as

follows:

Page 42: LATEEF, FATAI ABIOLA

30

Table 2.7: Iodine values of various groups of oils and fats

Groups Examples Range of iodine values

Waxes – Very small

Animal fats Butter, dripping, land 30 – 70

Non-drying oils Olive oil, arachis oil 80 – 110

Semi-drying oil Cottonseed oil, soya oil 80 – 140

Drying oils Linseed oil, sunflower

oil

125 – 200

Source: David (1976)

The iodine value is often the most useful figure for identifying an oil

or at least placing it into particular group. It should be noted that the less

unsaturated fats with low values are solid at room temperature, or conversely,

oils that are more highly unsaturated are liquids. Further point of interest is

that, in general, the greater the degree of unsaturation (i.e. the higher the

iodine value), the greater is the liability of the oil or fat to go rancid by

oxidation.

The iodine value is usually determined by Wij’s method.

SAPONIFICATION VALUE

The saponification value of an oil or fat is defined as the number of

milligram of potassium hydroxide required to saponify one gram of the oil or

fat. Saponification value (S.V) is a good indicator of the average molecular

weight of the constituent fatty acids (Gupta et al, 2007). It gives the number

in milligram potassium hydroxide of the base required to neutralize the free

fatty acids in the lipid, and to saponify the triglycerides in 1gram of the lipid.

As many oils have somewhat similar values (e.g. those in the olive oil

series fall within the range 188 – 196), the saponification value is not, in

general, as useful for identification purposes as the iodine value. The

saponification value is of most use for detecting the presence of palm-kernel

Page 43: LATEEF, FATAI ABIOLA

31

oil (SV 247) and coconut oil (SV 255), which contain a high proportion of

the lower fatty acids (David, 1976).

ACID VALUE OR FREE FATTY ACIDS (FFA)

The acid value of an oil or fat is defined as the number of mg of

potassium hydroxide required to neutralize the free acid in 1g of the sample.

The result is often expressed as the percentage of free acidity.

The acid value is a measure of the extent to which the glycerides in the

oil have been decomposed by lipase action. The decomposition is

accelerated by heat and light. As rancidity is usually accompanied by free

fatty acid formation, the determination is often used as general indication of

the condition and edibility of oils (David, 1976).

2.8 RANCIDITY

Fats and oil undergo changes during storage which result in the

production of an unpleasant taste and odour, which is commonly referred to

as rancidity. Rancidity is brought about by the action of air (oxidative

rancidity) or by micro-organisms (Ketonic rancidity). Oxidative rancidity is

accelerated by exposure to heat and light, by moisture and by the presence of

traces of certain metals (e.g. copper, nickel, iron). It is now generally

accepted that oxygen is taken up by the fat (or oil) with the formation of

compounds which react as peroxides. In general, the greater the degree of

unsaturation (the higher the iodine), the greater is the liability of the fat (or

oil) to oxidative rancidity. When the concentration of ‘peroxides’ reaches a

certain level, complex chemical changes occur and volatile products are

formed which are mainly responsible for the rancid taste and odour

(David,1976).

With most oils and fats, the free fatty acidity increases during storage

but, with refined oils particularly, the free fatty acidity figure is not

Page 44: LATEEF, FATAI ABIOLA

32

necessarily related to the extent to which rancidity has progressed. On the

other hand, although the ‘peroxides’ are possibly not directly responsible for

the taste and odour of rancid fats, the concentration of them as represented

by the peroxide value is often useful for assessing the extent to which

spoilage has advanced (David, 1976). Fortification of fats and oils with

antioxidants extends the storage time and protect essential nutrients (Krause

and Hunscher,1999). The main substrates for oxidation of lipids are the

unsaturated fatty acids which generally oxidize faster in a free state than

when they form part of triglycerides or phospholipids (Alais and Linden,

1999).

2.9 STORAGE AND OXIDATIVE STABILITY OF FATTY ACID

ESTERS

Storage stability refers to the ability of the fuel to resist chemical

changes during long term storage. The changes usually consist of oxidation

due to contact with oxygen from the air (Ferrari et al, 2005). Fatty acid

composition of the biodiesel fuel is an important factor in determining

stability towards air. Generally, the polyunsaturated fatty acids (C18:2,

linoleic acid; C18:3 linolenic acid) are most susceptible to oxidation. The

changes can be catalyzed by the presence of certain metals (including those

making up the storage container) and light. If water is present, hydrolysis can

also occur. The chemical changes associated with oxidation usually produce

hydroperoxides that can, in turn, produce short chain fatty acids, aldehydes

and ketones. Under the right conditions, the hydroperoxides can also

polymerize. Therefore, oxidation is usually denoted by an increase in acid

value and viscosity of the fuel. Often these changes are accompanied by a

darkening of the biodiesel colour from yellow to brown and the development

of a “paint” smell. When water is present, the esters can hydrolyse to long

chain free fatty acids which can also cause the acid value to increase.

Page 45: LATEEF, FATAI ABIOLA

33

There is currently no generally accepted method for measuring the

stability of biodiesel, the techniques generally used for petroleum-based

fuels, such as ASTM D 2274, have shown to be incompatible with biodiesel.

Other procedures, such as the oil stability index or the Rancimat apparatus,

which are widely used in the fats and oils industry, seem to be more

appropriate for use with biodiesel. However, the engine industry has no

experience with these tests and acceptable values are not known. Also, the

validity of accelerated testing methods has not been established or correlated

to actual engine problems. If biodiesel’s acid number, viscosity or sediment

content increase to the point where they exceed biodiesel’s ASTM limits, the

fuel should not be used as transportation fuel.

Figure 2.16: Oxidation Behaviour of Vegetable Oils and Animal Fats

Page 46: LATEEF, FATAI ABIOLA

34

2.10 Additization of Esters

Biodiesel, because it contains large numbers of molecules with double

bonds, is much less oxidatively stable than petroleum-based diesel fuel

(Gerpen et al, 2004).

Additives such as BHT (Butyl hydroquinone) and TBHQ (t-butyl

hydroquinone) are common in the food industry and have been found to

enhance the storage stability of biodiesel. Biodiesel produced from oils

naturally contain some antioxidants (tocopherols, i.e vitamin E), providing

some protection against oxidation (some tocopherol is lost during refining of

the oil prior to biodiesel production). Thermal process during biodiesel

pretreatment can also degrade some naturally occurring antioxidants in oils

and fats (Ferrari et al, 2005). Any fuel that will be store for more than 6

months, whether it is diesel fuel or biodiesel, should be treated with an

antioxidant additive.

Figure 2.17 shows typical antioxidants in oils and fats.

3 BHA (3-tert-butyl-4-hydroxyanisole 2BHA(2-tert-butyl-4-, hydroxyanisole

2 tert-butyl-4-methoxyphenol) 3 tert-butyl-4-methoxyphenol)

Fig 2.17: Butylated Hydroxyanisole (BHA)

OH

OCH3

C(CH3)3

C(CH3)3

OH

OCH3

Page 47: LATEEF, FATAI ABIOLA

35

Butylated Hydroxytoluene (BHT) Propyl Gallate (PG)

(2, 6 – di-tert-butyl-o-cresol; 4- methyl (Propyl gallate (n-propyl-

– 2, 6 – di-tert-butylphenol) 3,4,5 – trihydroxybenzoate)

Mono-tert-butylhydroquinone (TBHQ)

Figure 2.18: Antioxidants for oils and fats

2.11 Rheological Properties of Fluids

2.11.1 Biodiesel Rheology: Fluid flow phenomena

Rheology is defined as the science of the deformation and flow of

matter (Jacobs, 1999). The behaviour of a flowing fluid depends strongly on

whether the fluid is under the influence of solid boundaries. In the region

where the influence of the wall is small, the shear stress may be negligible

and the fluid behaviour may approach that of an ideal fluid, one that is

incompressible and has zero viscosity. The flow of such an ideal fluid is

called potential flow and is completely described by the principles of

Newtonian mechanics and conservation of mass. The relationship between

the shear stress and shear rate in a real fluid are part of science and Rheology.

Gases and most liquid are Newtonian. In a Newtonian fluid, the shear stress

(CH3)3 – C

CH3

OH

C(CH3)3

C – ( CH3)3

CH3

COOC3H7

OH HO

C(CH3)3

OH

OH

Page 48: LATEEF, FATAI ABIOLA

36

is proportional to the shear rate, and the proportionality constant is called the

viscosity.

τv = µ (du/dy)

where τv – shear stress

µ - viscosity

du/dy – shear rate

In SI unit τv is measured in Newton per square meter and µ is

kilograms per meter second or pascal-second. Viscosity data are generally

reported in millipascal-seconds or in centipoises (cP = 0.01P = 1m Pa.s)

(McCabe et al, 2005)

2.11.2 Viscosity

Viscosity can be simply defined as the internal friction acting within a

fluid, that is, its resistance to flow. A fluid in a glass, when inverted, is

subjected to gravitational forces, some fluids flow easily out of the glass,

some with difficulty and some not at all. Viscosity is therefore, the measure

of the rate of flow (Lewis, 1987). The viscosity generally increases with

molecular weight and decreases rapidly with increasing temperature. The

main effect of temperature change comes not from the increase in average

velocity, but from the slight expansion of the liquid, which makes it easier

for the molecules to slide past each another.

The viscosity is a strongly non-linear function of the temperature, but

a good approximation for temperatures below the normal boiling point is

ln µ = A + B

T

2.11.3 Kinematic Viscosity

This is the ratio of absolute viscosity to the density of a fluid µ/ ρ. The

property is called the kinematic viscosity and designated by υ. In SI, the unit

for υ is square meter per second. For liquids, kinematic viscosities vary with

Page 49: LATEEF, FATAI ABIOLA

37

temperature over a somewhat narrower range than absolute viscosities

(McCabe et al, 2005).

BACKGROUND: VISCOSITY OF BIODIESEL

Neat vegetable oils were primarily considered as alternatives for diesel

fuel but their very high viscosity at room temperature made them unsuitable

in diesel engines. Esters of lower alcohol (methanol, ethanol, propanol) of

plant or animal oils are very much lower in viscosities than neat oils.

Reducing viscosity is the major reason why vegetable oils or fats are

transesterified to biodiesel because the high viscosity of neat vegetable oils

or fat ultimately lead to operational problems such as engine deposits

(Knothe and Steidley, 2005; Krisnangkura et al, 2005). The viscosity of

biodiesel is slightly greater than that of petrodiesel but approximately an

order of magnitude less than that of the parent vegetable oil or fat.The best

known equation that correlates liquid viscosity and temperature is the

Andrade equation, given by:

µ = AeB/T

where A and B are constants. T is the

absolute temperature. The equation can be used for predicting viscosity up to

approximately the normal boiling point of the fluid (Krisnangkura et al,

2005).

The viscosity values are recorded for some common vegetable oils in

table 2.10.

2.12 Biodiesel Specifications and Properties

The biodiesel standard is framed as a set of property specifications

measured by specific ASTM test methods. The standard of biodiesel is

ASTM 6751 – 02.

ASTM D 6571 – 02 sets forth the specifications that must be met for a

fatty acid ester product to carry the designation “biodiesel fuel” or “B100” or

Page 50: LATEEF, FATAI ABIOLA

38

for use in blends with any petroleum-derived diesel fuel defined by ASTM D

975, Grades 1-D, 2-D, and low sulfur 1-D and 2-D.

The values of the various biodiesel properties specified by ASTM D

6751 are listed in Table 2.8. Each of these properties and the test method

used to measure it are also in Table 2.8. Also, Table 2.9 is for European

Biodiesel Standards EN 14212 for vehicle use and EN 14213 for heating oil

use.

Table 2.8:

Page 51: LATEEF, FATAI ABIOLA

39

Table 2.9:

39

Page 52: LATEEF, FATAI ABIOLA

40

Table 2.10: Kinematic Viscosity of oils

Oil Temperature (oF) Kinematic Viscosity (cSt)

Almond 100 43.2

Olive 100 46.7

Rapeseed 100 50.6

Cottonseed 100 35.9

Soyabean 100 28.5

Linseed 100 29.6

Sunflower 100 33.3

Castor 100 293.4

Coconut 100 29.8

Palm-kernel 100 30.9

Source: Levis, J. M. (1987).

2.13 Biodiesel Production Process Options

There are two major ways in which biodiesel can be produced. These

are: Batch processing and Continuous processing.

2.13.1 Batch Processing

The simplest method for producing alcohol esters is to use a batch,

stirred tank reactor. Alcohol to triglyceride ratios from 4:1 to 20:1

(mole:mole) have been reported with a 6:1 ratio most common (Sharma et al,

2008). The reactor may be sealed or equipped with a reflux condenser. The

operating temperature is usually about 65oC, although temperatures from

25oC to 85

oC have been reported (Gerpen et al, 2005). The most commonly

used catalyst is sodium hydroxide, with potassium hydroxide also used.

Typical catalyst loading range from 0.3% to about 1.5% (Gerpen et al, 2005).

Thorough mixing is necessary at the beginning of the reaction to bring

the oil, catalyst and alcohol into intimate contact. Towards the end of the

Page 53: LATEEF, FATAI ABIOLA

41

reaction, less mixing can help increase the extent of reaction by allowing the

inhibitory product, glycerol, to phase separate from the ester-oil phase.

Completions of 85% to 94% are reported (Gerpen et al, 2005).

Some groups use a two-step reaction, with glycerol removal between

steps, to increase the final reaction extent to 95+% (Sharma et al, 2008).

Higher temperature and higher alcohol:oil ratios can enhance the percent

completion. Typical reaction times range from 20 minutes to more than one

hour (Sharma et al, 2008).

Figure 2.19 shows a process flow diagram for a typical batch system.

The oil is first charged to the system, followed by the catalyst and alcohol

(methanol, ethanol, 1-propanol or butanol). The system is agitated during the

reaction time. Then agitation is stopped. In some process, the reaction

mixture is allowed to settle in the reactor to give an initial separation of the

esters and glycerol. In order processes, the reaction mixture is pumped into

settling vessel, or is separated using a centrifuge.

The alcohol is removed from both the glycerol and ester stream using

an evaporator or a flash unit. The esters are neutralized, washed gently using

warm, slightly acid water to remove residual methanol and salts, and then

dried. The finished biodiesel is then transferred to storage. The glycerol

stream is neutralized and washed with soft water. The glycerol is then sent to

the glycerol refining section.

Page 54: LATEEF, FATAI ABIOLA

42

Figure 2.19: Batch Reaction Process

For yellow grease and animal fats, the system is slightly modified with

the addition of an acid esterification vessel and storage for the acid catalyst.

The feedstock is sometimes dried (down to 0.4% water) and filtered before

loading the acid esterification tank. The sulphuric acid and alcohol (e.g

methanol, 1-propanol) mixture is added and sometimes the system is

pressurized or a co-solvent is added. Glycerol is not produced. If a two-step

acid treatment is used, the stirring is suspended until the alcohol phase

separates and is removed. Fresh alcohol is added and the stirring resumes.

Once the conversion of the fatty acids to methyl esters has reached

equilibrium, the alcohol/water/acid mixture is removed by settling or with a

centrifuge. The remaining mixture is neutralized or sent straight into

transesterification where it will be neutralized using excess base catalysts.

Any remaining free fatty acids will be converted into soaps in the

transesterification stage.

Page 55: LATEEF, FATAI ABIOLA

43

2.13.2 Continuous Process Systems

Continuous process of biodiesel production uses a Continuous Stirred

Tank Reactors (CSTRs) in series. The CSTRs can be varied in volume to

allow for a longer residence time in CSTR to achieve a greater extent of

reaction. Glycerol is usually decanted from the stream leaving the first CSTR

before charging the stream into the second CSTR. An essential element in

the design of a CSTR is sufficient mixing input to ensure that the

composition throughout the reactor is essentially constant. This has the effect

of increasing the dispersion of the glycerol product in the ester phase.

Continuous stirred tank reactor however, has disadvantage of low conversion

per reactor volume (Gerpen et al, 2004).

Figure 2.20: Plug Flow Reaction System

2.14 SPECTROSCOPY FOR BIOFUEL ANALYSIS

2.14.1 Use of Molecular Spectra as Aids in the Identification of Organic

Structures

The absorption of electromagnetic radiation by some part of the

molecule may be used to help gain precise information about structure. In

Page 56: LATEEF, FATAI ABIOLA

44

the case of infrared (I.R) spectroscopy, the radiation is passed through the

sample under analysis and the spectrum is recorded.

During the early stages of analysis, it if often advantageous to study

first the infrared spectrum, bearing in mind the evidence already obtained

from the infrared spectrum. Consideration of the molecular formula will

often allow the rejection of a number of alternative interpretations consistent

with the sample piece of spectroscopic evidence (Murray, 1985). Molecules

have different bond structures which absorb unique wavelength of light. I.R

measures how light interact with fuel components. The amount of light

absorbed is proportional to that components concentration in the fuel.

Infrared (I.R) region produces primary absorbances that give

fundamental knowledge of the types of chemical groups presents in the fuel.

The benefits include fast analysis, no sample preparation, no waste

chemicals, no consumables, portable/automated instruments. I.R application

for biofuel includes feedstock analysis, determination of product blends,

final product quality and contamination (Ritz and Nash, 2004)

Spectroscopic methods are being increasingly utilized for quality

control purposes (Knothe, 1999). The analytical issues with biodiesels have

two sources. The production facilities and terminal facilities need to ensure

quality (completion of transesterification, glycerol removal etc) while testing

laboratory and regulatory agents must ensure the labeled blend levels are

present. Infrared provides a rapid, precise and accurate tool for this analysis

when these needs are taken into account (Bradley, 2007).

Page 57: LATEEF, FATAI ABIOLA

45

2.14.2 Infrared Spectra

Carbonyl stretching vibrations Wavenumber range/cm-1

1. Aldehyde, aliphatic, saturated 1470 – 1720

aromatic 1715 – 1695

2. Ketone, R – CO – R’ 1725 – 1705

R – CO – Ar 1700 – 1680

Cyclic, saturated 1775 – 1750

3. Carboxylic acid, R – COOH (diver) 1725 – 1700

Ar – COOH 1700 – 1680

(O - C - - - O, intermolecularity 2700 – 2500

hydrogen bonded (several bonds)

4. Ester, R – COOR’ 1750 – 1735

Ar – COOR 1730 – 1715

R – COOAr 1800 – 1730

(Murray, 1985)

Page 58: LATEEF, FATAI ABIOLA

46

CHAPTER THREE

MATERIALS AND METHODS

3.1 Raw Materials

Castor seeds, Palm kernel seeds and groundnut seeds were obtained

from Ogige market in Nsukka, Enugu State.

3.2 Apparatus

The apparatus used in the laboratory are as follows:

Centrifuge (Hettich Universal II, Serial number 27712), Oven (Labor

Muszeripari Muvek, Type: LP-301), Thermostated water-bath (Laboratory

Thermal Equipment, Serial no: 72294150), Electronic weighing balance

(Electronic Scale High Precision Mettler Toledo, Model AD01), Manual

blender (Lander YCIASA 2E), Mechanical grinder, Extraction column,

Viscometer (Ferranti Portable Viscometer, Model: VL,VH), Rotary

evaporator (Rotavapor-R), Electric heater (STC Model no: 0016), and

Nicolet Avartar 310FTIR (Fourier Transform Infrared Spectroscopy, Model

no:310).

3.3 Extraction of Oil from Various Seeds

3.3.1 Extraction of Oils from Castor Seed

The castor seeds were de-shelled by hand picking and the shell were

separated from the seed. The seeds were spread and sundried for three

consecutive days. The seeds of initial weight 5.735kg reduced to 5.435kg

after sun drying. The sun-dried seeds were left in the oven for three hours for

dehydration (removal of moisture present in the seed) and then chopped into

smaller bits with hand and later with mortal pestle. This was done in order to

create large surface area of contact with the solvent for maximum extraction.

Absolute ethanol (2.5 litres) was used as the extracting solvent and

Page 59: LATEEF, FATAI ABIOLA

47

extraction column (cold extraction method) was used for the extraction of oil

from the seeds.

3.3.2 Extraction of Oils from Palm-Kernel Seed

Palm kernel seeds (3.281kg) were ground with mechanical grinder.

The ground palm kernel seeds were soaked with the 2.5 litres of solvent (n-

hexane) in an extraction column and left standing for 72 hours, after which

the tap was opened and the oil collected.

3.3.3 Extraction oil Oils from Groundnut Seed

Groundnut seeds (4.298kg) were sundried, after which it was ground

with the aid of manual blender, and then soaked with the 2.5 litres solvent

(n-hexane) in an extraction column. The extraction column was left standing

for 72 hours after which the tap was opened and the oil collected.

The solvent from the three seeds (castor, palm-kernel and groundnut)

were recovered with rotary evaporator (model ROTAVAPOR–R). The

solvent recovered were stored for further use.

3.4 Pre-treatment of Oils

Measuring cylinder was used to measure 300ml of oil and the oil was

heated to 75oC in a 500ml beaker using electric heater (STC Model no:

0016). The oil was mixed with 0.1 (v/v)% of 85% phosphoric acid, distilled

water to about 0.2 wt % of the oil. Magnetic stirrer was placed at the bottom

of the 500ml beaker and it was used to homogenized the oil for about 30

minutes. Gum which result from the homogenization was allowed to settle

and the oil was decanted into another 500ml beaker for refining.

In the refining step, temperature of the oil was restored to 75oC under

magnetic stirrer and 9.5wt% of NaOH solution was added gradually, as the

mixture was continuously homogenized. This process converts the free fatty

Page 60: LATEEF, FATAI ABIOLA

48

acids into soaps. 15 wt% distilled water of the total mixture was used to

wash the oil free of soap in a 500ml separating funnel. The washed oil was

later dried at 105oC using oven (LP-301) for about 30 minutes.

3.5 Determination of the Optimal Catalyst Weight

The alkali used for transesterification affects the yield of oil samples

into biodeisel. Hence, the optimal catalysts tests were studied for each oil

sample. (See table B2 in Appendix II for the process parameters).

Transesterification

The laboratory scale transesterification reactors (batch reactor) to

produce propyl esters from castor oil, palm kernel oil and groundnut oil were

carried out in a 200ml conical flask (air-tight flask) and mounted on a

magnetic stirrer. The magnetic stirrer was set to a constant speed throughout

the experiment, to ensure uniform agitation and thorough homogenization of

the reaction mixture.

Optimal catalyst tests were determined for each oil sample using 50ml

of the refined oils and the volume of 1-propanol used on the basis of 3:1, 1-

propanol to oil molar ratio. The catalyst used is sodium hydroxide (NaOH)

pellet. The weights of the catalyst were varied from 0.05, 0.10, 0.15, 0.20,

0.25, 0.30, 0.35 and 0.40g.

NaOH pellets was dissolved in 15ml of 1-propanol and the mixture

stirred for 15 minutes to form sodium methoxide (CH3ONa) in an air-tight

conical flask.

This sodium methoxide was introduced gently into the heated oil in

the reactor and the entire content was brought to a temperature of 55oC and

then held at this temperature for 60 minutes. The reaction product mixture of

the transesterification were allowed to separate into two phases by standing

for 6 hours in a separating funnel (100ml) so as to separate glycerol from the

biodiesel. The two layers – superior (biodiesel) and inferior (glycerol) were

separated by washing with warm distilled water to remove impurities. The

Page 61: LATEEF, FATAI ABIOLA

49

denser soapy mixtures was carefully drained from the bottom of the 100ml

separating funnel, leaving behind the superior biodiesel layer. The volumes

of the biodiesel obtained were determined in a measuring cylinder. Graphs of

biodiesel yield fraction against catalyst weight per volume of oil were

estimated and the graphs plotted.

Note: from the volume ratio of sample at 3:1 alcohol/oil molar ratio, it

was observed that 50cm3 of refined oil requires 15cm

3 of alcohol.

3.6 Determination of Kinematic Viscosity of Crude Plant oils and

Biodiesel Samples

Kinematic viscosity values of biodiesel samples was determined with

Ferranti portable viscometers (Model VL for PKO and GNO, Model VH for

CSO) at 30oC following the standard method as outlined in the Ferranti

portable viscometer manual. About 150ml of sample were measured into a

300ml beaker and placed under an outer cylinder. The outer cylinder was

immersed in the sample fluid by allowing the cylinder to rotate until the

reading was stable. The viscometer was raised above the fluid and tilted to

allow the sample to flow from the annulus back into the container and the

readings were taken from the Ferranti Portable Viscometer calibrations by

selecting the appropriate speed.

Dynamic and kinematic viscosity data which are related by density as

a factor were determined for crude plant oils (castor oil, palm kernel oil and

groundnut oil) and biodiesel produced from these oils. The viscosities in

poises, at a given speed and cylinder combination, were obtained by

multiplying the instrument reading by the appropriate Multiplying Factor

given on the calibration chart.

Page 62: LATEEF, FATAI ABIOLA

50

3.7 Alkali-Catalyzed Batch Production of Biodiesel

The transesterification reaction was carried out by reacting oils (castor,

palm kernel and groundnut oils) with 1-propanol in the presence of a basic

catalyst - sodium hydroxide (NaOH) pellet, analytical grade (Joechem UNN)

and the biodiesel fuels were processed in a batch type reactor at 6:1 alcohol

to oil molar ratio. Excess alcohol was used in order to shift the equilibrium

to the right. (See table B3 in Appendix II for the process parameters).

Procedure:

One hundred milli-litre each of the three oil samples (castor, palm

kernel and groundnut oils) was heated to 65oC and placed in a 250ml flat

bottom flask-batch reactor (at 6:1 alcohol to oil molar ratio, 100ml of castor

and palm kernel oil requires 60ml of 1-propanol and 100ml of groundnut oil

requires 50ml of 1-propanol). The optimal catalyst (NaOH) weight earlier

calculated for the three oils (0.50g for castor and palm kernel oils and 0.40g

for groundnut oil) was dissolved into the alcohol by vigorous stirring in a

separate air-tight container of 200ml. The alcohol-optimal catalyst weight

mixtures were poured into the oils and the final mixture stirred vigorously

for 60 minutes in an air-tight container. The reaction product mixture wass

allowed to separate into two phases at the end of the reaction; ester and crude

glycerol, by standing for 15 hours in a separating funnel so as to separate

glycerol from the biodiesel. The tap of the separating funnel was opened to

evacuate the lower layer (glycerol) and the crude biodiesel was left in the

separating funnel.

Fifty milli-litre warm distilled water at 45oC was used to wash the

crude biodiesel thrice and dried in the oven at 105oC for 60 minutes. The

volumes of the biodiesel obtained were recorded and samples were used for

characterization.

Page 63: LATEEF, FATAI ABIOLA

51

3.8 Characterization of the Crude, Refined Oils, Biodiesel and

Petrodiesel

3.8.1 Saponification Value (SV) Determination

One gram of sample was dissolved in 20ml of the alcoholic potassium

hydroxide solution and put into a conical flask which was attached to a

reflux condenser and the flask was heated in boiling water for 10 minutes,

shaking frequently. Two drops of phenolphthalein indicator were added into

the hot solution and it was titrated with 0.5M hydrochloric acid, shaking

frequently to colourless end point (titration = a ml). A blank was equally

carried out at the same time (titration = b ml).

Saponification value = (b – a) x 28.05

Weight in g of sample

The procedure was repeated two more times and the average titre

values calculated.

3.8.2 Acid Value (AV) Determination

One gram of sample was dissolved and mixed with 25ml diethylether

with 25ml ethanol and 2 drops of phenolphthalein indicator and was

carefully neutralized with 0.1M sodium hydroxide, shaking constantly until a

pink colour which persists for 15 seconds was obtained.

Acid value = Titration (ml) x 5.61

weight of sample used (g)

The procedure was repeated two more times, and the average titre

values calculated.

3.8.3 Iodine Value (IV) Determination

One gram of our sample was dissolved in 5ml carbon tetrachloride

(CCl4) and 5ml Wijs’ solution in a 250ml stoppered conical flask and

allowed to stand in the dark cupboard for 30 minutes. 10ml of the potassium

iodine solution and 25ml of distilled water mixed and titrated with 0.1M

Page 64: LATEEF, FATAI ABIOLA

52

thiosulphate solution using 2ml starch as indicator. Titration of the solution

was continued until the blue black colouration due to iodine was discharged

(titration = a ml). The procedure was repeated twice and the average titre

was calculated. Blank at the same time commenced with 10ml of carbon

tetrachloride (titration = b ml)

Iodine Value = (b – a) x 1.269

weight (in g) of sample

3.8.4 Specific Gravity Determination

Twenty five milli-litre density bottle (Wb) was thoroughly washed

with detergent, water and petroleum ether, dried and weighed (Wb). The

density bottle was filled with water, corked and weighed again (W1). The

water in the density bottle was discharged, cleaned with the diethyl ether.

After drying the bottle, it was filled with oil/biodiesel/petrodiesel samples

and the weight (W2) was taken.

Specific gravity = W2 – Wb

W1 – Wb

3.8.5 Peroxide Value Determination

One gram of biodiesel sample was weighed into a 100ml conical flask

containing 20ml of solvent mixture (2:1 volume of glacial acetic acid and

chloroform), 20ml of 50% potassium iodide (KI) solution and 1.0g of

potassium iodide (KI) crystals and the whole mixture agitated. The mixture

was then placed in a boiling water at 100oC for 30 seconds. About 5 drops of

starch solution was added to the mixture which turned the yellow colour of

the mixture to black and then titrated against 0.1M sodium thiosulphate

(Na2S2O3) until the black colour turns white (colourless). This procedure was

repeated two more times, and the average titre value was calculated (Ibitoye,

2006).

Page 65: LATEEF, FATAI ABIOLA

53

Peroxide Value (P.V) = T x M x 100

Sample weight (g) mmol peroxide/kg sample

where T = titre value of Na2S2O3

M = Molarity of Na2S2O3

3.8.6 Calorific (heating/combustion) Value Determination using Bomb

Calorimeter (Model: XRY – 1A)

Castor, palm kernel and groundnut oils biodiesel sample were

characterized for their combustion values – and compared to conventional

diesel fuel (Petrodiesel).

Procedure:

The outer canister of the bomb calorimeter was filled with water. The

inner canister was filled with 3 litres of distilled water. 1g of biodiesel/

petrodiesel sample to be evaluated was measured and placed in a mould

(small metal crucible). 10cm ignition thread (wire) connected to the

electrodes of the oxygen bomb, was placed and allowed to keep in touch

with the sample. The bomb was filled in with oxygen at a pressure of 2.8 –

3.0 mPa and then transferred into the inner canister (filled with 300cm3 of

distilled water). The necessary wires were connected and the temperature

sensor was placed into inner canister.

The power was switched on and the water inside the inner canister was

stirred for about 2 minutes and the initial temperature of the water was noted

and denoted To. The Bomb calorimeter was fired and the final temperature

(Tf) was recorded when the time got to 31 minutes. Length of the pieces of

unburnt firing wire was measured (l) and the inner lining of the oxygen

bomb and crucible were washed with distilled water into a conical flask. The

wash solution was titrated against 0.0709N Na2S2O3, using 2 drops of methyl

red indicator.

Page 66: LATEEF, FATAI ABIOLA

54

Calorific values of the samples were calculated from the expression:

W = E∆T – Φ – V

m

where w = heat of combustion of sample (calorie/g)

m = mass of sample to be evaluated (g)

E = 13,039.308 calories/g , Benzoic acid standard

T = change in temperature = Tf – To

Φ = 2.3l (where l = length of the unburnt wire)

V = volume of alkali (Na2S2O3 solution) cm3

The combustion value were converted to Joules from the expression

1 calorie = 4.148 Joules

3.9 Investigation of Temperature Dependence of Biodiesel Kinematic

Viscosity and Specific Gravity / Biodiesel-Petrodiesel Blending

The viscosity of fatty acid propyl esters (biodiesel) were measured at

various temperatures. The esters were analyzed for viscosity at temperatures

of 32, 45, 60, 75oC (using samples of alkali catalyzed biodiesel for the

production of biodiesel at 6:1 alcohol to oil molar ratio as in section 3.8 and

allowed to cool to room temperature 30oC) using Ferranti portable

viscometer (Model: VL, VH).

Procedure

The temperatures of the biodiesel samples (castor, palm kernel and

groundnut) and petrodiesel (commercial) were taken each and recorded at

room temperature 30oC. About 150ml of biodiesel/petrodiesel sample was

measured into a 300ml beaker. The outer cylinder was immersed in the

biodiesel/petrodiesel sample, the speeds required were selected and the

motor was switched on. The motor allows the cylinder to rotate until the

reading was stable. The viscometer was raised above the

biodiesel/petrodiesel sample and tilted, to allow the sample to flow from the

Page 67: LATEEF, FATAI ABIOLA

55

annulus back into the container. The viscosity was read on the calibrated dial

at the top of the portable viscometer.

Using the method for the determination of specific gravity described

earlier, biodiesels (fatty acid propyl esters) were found to be temperature

dependent. The same temperature intervals used for the determination of

kinematic viscosity were repeated for the specific gravity. In order to reduce

castor oil kinematic viscosities so as to meet ASTM standards, castor oil

biodiesels were blended with petrodiesel in the following manner:

Petrodiesel (90%) – biodiesel (10%) : B10

Petrodiesel (80%) – biodiesel (20%) : B20

Biodiesel (Propyl esters) : B100

Petrodiesel : P100

3.10 Transesterification Kinetics in a Batch Reactor

Transesterification reactions were performed in a 250ml round bottom

flask (herein referred to as batch reactor) on castor, palm kernel and

groundnut oils at 6:1 alcohol to oil molar ratio in a thermostated water bath

that was capable of varying temperature between 0–120oC. The reaction

temperatures were varied at 32, 45, 60 and 75oC (referred to desired

temperature in the procedure) at various time interval 5, 10, 20, 60 seconds.

(See table B4 in Appendix II for the process parameters).

Procedure:

The batch reactor (250ml round bottom flask was filled with 100ml of

refined oil sample (castor, palm kernel and groundnut oils) was heated to the

desired temperature using optimal catalyst weight–NaOH pellet required

for each oil (0.50g for castor and palm kernel oil and 0.40g for groundnut

oil), and a measured amount of 1-propanol (60ml for CSO and PKO and

50ml for GNO) were mixed and heated to the desired temperature with the

aid of thermostated waterbath. The reacting mixture (1-propanol and sodium

Page 68: LATEEF, FATAI ABIOLA

56

hydroxide (NaOH pellet)) were heated to 65oC in a separate 100ml capacity

container (air-tight) which was then added to the oil originally in the batch

reactor. Mechanical stirrer was inserted and the reacting mixture agitated

(air-tight). At this point, the reaction was assumed to have started and timed.

At various times; 5, 10, 20, 60 seconds, 10ml of the sample was

withdrawn quickly from the reactor with a syringe and then placed in the test

tubes containing 3 drops of 0.1M hydrochloric acid, which was added to

neutralize the catalyst (Darnoko and Cheryan, 2000).

The sample was then centrifuged (Centrifuge-Hetich Universal II,

Speed 0 to 100%). The centrifuge was set at 90% for 20 minutes. Two layers

(phases) were observed after centrifugation – the top layer, consisting of

biodiesel and a semi-solid viscous layer at the bottom. The biodiesel layer

(superior layer) was carefully decanted into a container (50ml capacity).

Volume of biodiesels were recorded as in Table 4.10 This procedure

was repeated for the three oils.

3.11 Infrared Spectroscopy-Biodiesel Product Analysis Procedure

Biodiesel produced at 65oC as outlined in section 3.7 using optimal

catalyst weights were analyzed using Fourier Transform Infrared

spectrophotometer (Model: 310FT-IR). The purpose of this analysis was to

measure how light interacts with fuel components.

Procedure:

A Nicolet Avartar 310FT-IR (Fourier Transform Infrared

Spectrophotometer) equipped with KBr beam splitter. The smart ARKTM

attenuated total reflection sensory was used to collect the data. About 0.4ml

of biodiesel samples were smeared to cover the sampling crystal. Spectra

were collected in 40 seconds.

Page 69: LATEEF, FATAI ABIOLA

57

The data were collected using OMNICTM

spectroscopy software

showing spectrum region, absolute threshold, sensitivity, peak list-positions

as shown in figures 4.10 - 4.15.

Page 70: LATEEF, FATAI ABIOLA

58

CHAPTER FOUR

RESULTS AND DISCUSSION

This is the section in which the data obtained for the project/thesis are

presented, analyzed and discussed. Calculation involved in the

determinations are contained in Appendices I, II, III and IV.

4.1 EXTRACTION OF CRUDE PLANT OILS

The results obtained from the extraction of various plant oils (Castor

seed, palm kernel and groundnut oils) are shown in Table 4.1.

Table 4.1: Yield of oils

Sample Weight of seeds

(Kg)

Weight of extracted

oil (Kg)

Percentage (%)

yield

Castor seed oil 7.80 3.80 48.72

Palm kernel oil 6.56 2.78 42.38

Groundnut oil 8.10 3.76 46.42

The three oil samples showed good yield, with castor seed, having

highest percentage (%) yield.

4.2 PRE-TREATMENT OF OILS

It is a well known fact that crude plant oils contain some free fatty

acids and phospholipids. The three oil samples used for biodiesel production

were pretreated. The pre-treatment (processing) result is shown in Table 4.2

Table 4.2: Percentage Loss on Pre-treatment

Sample Weight of unrefined

oil (kg)

Weight of refined

oil (kg)

Percentage (%)

loss on pre-

treatment

Castor seed oil 3.80 3.58 5.78

Palm kernel oil 2.78 2.61 6.12

Groundnut oil 3.76 3.52 6.38

Page 71: LATEEF, FATAI ABIOLA

59

The free fatty acids and phospholipids are responsible for the

significant losses recorded during pre-treatment. The oil extracted from

groundnut has high acid values which is responsible for its high percentage

loss during pre-treatment with 6.38%, followed by palm kernel oil (6.12%).

4.3 OPTIMAL CATALYST WEIGHT TEST

The graphical relationship between the biodiesel yield (%) and

percentage weight per volume of the catalyst (% wt/v) are depicted in the

Figure 4.1.

From Appendix IV (see Table D1 in Appendix IV for the result of the

optimal catalyst weight test), it is observed that the product volume

(biodiesel yield) increased steadily from 0.10% wt/v of the catalyst (NaOH)

until it peaked at 0.5% wt/v of catalyst and thereafter, a decrease was

witnessed for CSO and PKO. But GNO peaked at 0.4% wt/v but witnessed

decrease thereafter. It is clear therefore that increment in percentage weight

per volume of the catalyst would not yield further volume increase in

biodiesel obtained from castor seed, palm kernel and groundnut oils. The

optimal catalyst weight test helps to confirm that increase in the amount of

catalyst only leads to production of soaps and no biodiesel.

Page 72: LATEEF, FATAI ABIOLA

60

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.810

20

30

40

50

60

70

80

Percentage[%] (wt/v) of catalyst

Pe

rce

nta

ge

[%]

Bio

die

se

l Y

ield

CSOCSO

GNOGNO

PKOPKO

Fig. 4.1: Variation of Percentage (%) Yield of Biodiesel with Percentage

(%) (w/v) of Catalyst for CSO, PKO and GNO

Page 73: LATEEF, FATAI ABIOLA

61

4.4 Properties of Crude, Refined and Biodiesel Obtained

Different properties like saponification value (S.V), iodine value (I.V)

and acid value (A.V) were determined for crude, refined and biodiesel

produced from castor, palm kernel and groundnut oils. These results were

discussed in the following sections.

Table 4.3: Result for the Characterization of Crude, Refined and Biodiesel

Sample S.V

(mg KOH/g)

A.V

(mg NaOH/g)

I.V

g of Iodine/g

Crude CSO 183.70 3.02 12.33

Crude PKO 242.80 3.04 12.69

Crude GNO 189.70 3.78 75.45

Refined CSO 181.00 0.74 11.23

Refined PKO 238.40 0.79 12.38

Refined GNO 182.30 0.98 74.15

CSO biodiesel 168.20 0.44 11.11

PKO biodiesel 204.54 0.48 12.11

GNO biodiesel 180.08 0.52 74.05

CSO – Castor oil, PKO – Palm kernel oil and GNO – Groundnut oil

4.4.1 Saponification Value

Saponification values of propyl esters prepared from castor, palm

kernel and groundnut oils were less than those prepared from crude oils. This

could be due to the presence of phosphatides which were removed during

degumming and refining process. The saponification values of propyl esters

were within the range of 168.20 to 242.80 mgKOH/g. The average molecular

weight of oil can be calculated by multiplying the inverse of saponification

value by 168000 (Titipong, 2006). Therefore, the higher the saponification

value, the lower the molecular weight.

Page 74: LATEEF, FATAI ABIOLA

62

4.4.2 Iodine Value

The iodine value of the conventional diesel fuel was approximately 10

(Gupta et al, 2007). Therefore, the biodiesel had significantly higher degree

of unsaturation than diesel fuel (10.11g of iodine/g for castor biodiesel,

12.11g of iodine/g for palm kernel biodiesel and 74.05g of iodine/g for

groundnut oil biodiesel), though the iodine values of the refined and

unrefined oils did not show any appreciable difference. This implies that

diesel engine utilizing biodiesel is more susceptible to gum formation than

that utilizing conventional diesel fuel due to the higher iodine values of

biodiesel.

4.4.3 Acid Value

The acid value of propyl esters prepared from castor, palm kernel and

groundnut oils were 0.44, 0.48 and 0.52 mg NaOH/g respectively and 3.02,

3.04, 3.78 mg NaOH/g for unrefined oil and 0.74, 0.79, 0.98 mg NaOH/g for

refined oils. The results shows that both refined and unrefined oils could not

be used directly as fuel in diesel engine since their acid value is above the

ASTM requirement while the results of biodiesel produced from these oils

conform to ASTM D 6751 standards.

4.5 Physical Characterization of Castor, Palm Kernel and Groundnut

Oils Biodiesel and Petrodiesel

In assessing the suitability of the castor, palm kernel and groundnut

oils biodiesel produced as alternative diesel fuel, the CSO, PKO, GNO

biodiesel and the commercial grade fossil diesel (petrodiesel) were analyzed

for heat of combustion (heating/calorific value), specific gravity and

kinematic viscosity.

Results obtained are presented in the Table 4.4.

Page 75: LATEEF, FATAI ABIOLA

63

Table 4.4: Physical Characterization of Biodiesel and Petrodiesel

Specific gravity at

28oC

Calorific

value MJ/kg

Kinematic

Viscosity (mm2/s) at

40oC

CSO biodiesel 0.87 37.54 7.21

PKO biodiesel 0.90 38.64 3.56

GNO biodiesel 0.89 37.52 4.12

Petrodiesel 0.83 45.52 2.85

4.5.1 Specific Gravity

The specific gravity recorded for CSO, PKO and GNO biodiesels

were higher than the values obtained for petrodiesel. 1.049 times that of

petrodiesel for CSO, 1.048 times that of petrodiesel for PKO and 1.078 times

that of petrodiesel for GNO.

The specific gravity obtained for the castor, palm kernel and

groundnut oils biodiesel falls within the limit specified for biodiesel fuel in

Europe (EN 14214:086–0.90). Specific gravity (also known as relative

density) refers to the ratio of the density of a fuel to the density of water at

the same temperature. The level of agreement recorded in specific gravities

for the CSO, PKO and GNO biodiesels is an important pointer to suitability

of the biodiesel fuel substitute as important fuel performance indicators such

as heating values, fuel storage are correlated with specific gravity (Yuan et al,

2004; Ajav and Akingbehin, 2002).

4.5.2 Calorific Value

The petrodiesel presented the highest calorific value – 45.52 MJ/kg,

which are far higher than the biodiesel produced from castor, palm kernel

and groundnut oils. Calorific value is the quantity of heat energy, which is

emitted by fuel at the time of combustion under set conditions of experiment.

It was found that overall, that castor, palm kernel and groundnut oils propyl

Page 76: LATEEF, FATAI ABIOLA

64

esters (biodiesel) behaved comparably to diesel fuel (petrodiesel) in terms of

rate of heat release.

4.5.3 Kinematic Viscosity

Fuel viscosity is regulated by the standards at 40oC. From Table 4.4,

CSO, PKO and GNO biodiesel has higher viscosity than conventional diesel

fuel (petrodiesel) in agreement with reports from several researchers (Alamu

et al, 2007; Ajav and Akingbehin, 2002). Kinematic viscosity is defined as

the resistance to flow of a fluid under gravity. The viscosity of biodiesel

from castor oil is thrice the viscosity of fossil diesel (petrodiesel) and for

groundnut oil biodiesel, viscosity is twice the viscosity of petrodiesel.

The kinematic viscosity for PKO and GNO biodiesel falls within the

specified limits by ASTM D6571 (3.5-5.0) but that of castor oil biodiesel

does not. Knothe and Steidley (2005), reported that castor oil biodiesel, in its

neat form exceeds all kinematic viscosity specifications in biodiesel

standards due to the high content of ricinoleic acid. The reported technical

implication of higher viscosity biodiesel is that it decreases the linkages of

fuel in a plunger pair and in turn it changes the parameters of a fuel supply

process (Lebedevas and Vaisekauskas, 2006).

4.6 Alkali-Catalyzed Batch Production of Biodiesel at 65oC

Table 4.5: Results of Alkali-Catalyzed Batch Production of Biodiesel at 65oC

Pre-

treated oil

sample

Volume

of oil

(cm3)

Volume of

1-propanol

(CH3CH2CH2OH)

Volume

of

biodiesel

(cm3)

Conversion Percentage

(%)

Conversion

Castor oil 100.00 60.00 89.50 0.895 89.5

Palm

kernel oil

100.00 60.00 87.00 0.870 87.0

Groundnut

oil

100.00 50.00 81.80 0.818 81.8

Page 77: LATEEF, FATAI ABIOLA

65

The biodiesel was produced by transesterifying refined oil samples

(castor oil, palm kernel and groundnut oils) at 6:1 1-propanol to oil molar

ratio in the presence of their various optimal catalyst weight earlier

calculated (0.50 wt/v for castor and palm kernel oils, and 0.40 wt/v for

groundnut oil) at 65oC for 60 minutes. The process for the production of

biodiesel using an alkali (sodium hydroxide – NaOH) catalyzed method in a

batch reactor was composed of glycerol separation steps (after each

transesterification step) and an ester purification step.

As can be seen from table 4.5 , the process proved to be successful for

the production of biodiesel from a high quality feedstock (refined oils).

Castor oil gave the highest conversion (89.5%) followed by palm kernel

87.0% and groundnut 81.8%. The result proved to be consistent within the

limit of experimental errors. The problems of saponification was greatly

reduced because the feedstock (castor, palm kernel and groundnut oils) used

were pre-treated. The pre-treatment greatly reduced the free fatty acid which

reacts with sodium hydroxide to form soap which could have in turn reduced

the conversion of oil, alcohol (1-propanol) and catalyst (NaOH) into

biodiesel.

Page 78: LATEEF, FATAI ABIOLA

66

4.7 Viscosity of Biodiesel at Different Temperatures

Table 4.6: Variation of Kinematic Viscosity of Biodiesel with Temperature

Temperature Kinematic Viscosity

T

(oC)

T

(K)

1/T

(K-1

)

1/T

(K-1

)

V (mm2/s) ln V

CSO PKO GNO CSO PKO GNO

32 303 0.0033 3.30x10-3

8.12 4.05 5.13 2.09 1.40 1.64

45 318 0.0031 3.14x10-3

7.14 3.54 4.08 1.97 1.26 1.41

60 333 0.0030 3.00x10-3

7.02 3.50 4.01 1.95 1.25 1.39

75 348 0.0029 2.87x10-3

6.89 3.46 3.64 1.93 1.24 1.29

The graphs obtained (Figures 4.2 – 4.4) using Andrade equation shows

relationship between kinematic viscosity V and temperature T (K). In

studying the kinematic viscosity dependence on temperature, variation in

viscosity were analyzed using Andrade Equation (V=AexpB/T

) to establish

the relationship between the viscosities of the sample, with temperature. The

result showed that as the temperature increases, kinematic viscosity

decreases.

From the graphs, Andrade equation was simplified to obtain constants

A and B.

Table 4.7: Simplification of Andrade Equation of Biodiesel Sample

Sample A B

CSO biodiesel 1.76 466.67

PKO biodiesel 1.05 500.00

GNO biodiesel 1.35 599.10

Page 79: LATEEF, FATAI ABIOLA

67

1/T (K-1

)

Fig. 4.2: Graph of lnV CSO against 1/T (K-1

)

1.70

1.60

2.00

1.90

ln V

2.20

1.80

2.10

0.0027 0.0029 0.0027 0.0030 0.0031 0.0032 0.0033 0.0034

Page 80: LATEEF, FATAI ABIOLA

68

Fig. 4.3: Graph of lnV PKO against 1/T (K-1

)

1.20

1.10

1.50

1.40

ln V

1.70

1.30

1.60

0.0027 0.0029 0.0027 0.0030 0.0031 0.0032 0.0033 0.0034

Page 81: LATEEF, FATAI ABIOLA

69

1/T (K-1

)

1.20

1.10

1.50

1.40

ln V

1.70

1.30

1.60

0.0027 0.0029 0.0027 0.0030 0.0031 0.0032 0.0033 0.0034

Fig. 4.4: Graph of lnV GNO against 1/T (K-1

)

Page 82: LATEEF, FATAI ABIOLA

70

A is evaluated from the intercept of the graph of lnV against 1/T (K-1

)

and B is evaluated from the slope of the graph as follows.

B = lnV

(1/T) (K-1

)

Similarly, using the same biodiesel samples, the specific gravities are

recorded at room temperature as shown in the table 4.8.

Table 4:8: Variation of Specific Gravity of Biodiesel Samples with

Temperature

Temperature Specific Gravity

(oC) CSO PKO GNO

32 0.86 0.88 0.82

45 0.84 0.87 0.81

60 0.81 0.85 0.80

75 0.80 0.84 0.80

The graph of specific gravity against temperature (oC) is shown in

Figure 4.5.

The specific gravities of biodiesel fuel displays a linear specific

gravity temperature relationship. Though there is less significant decrease in

specific gravity with respect to biodiesel produced as temperature increases.

Viscosity and specific gravity of biodiesel samples were found to be

temperature dependent. As the temperature increases from 32 to 75oC, their

specific gravities slightly decreases.

Page 83: LATEEF, FATAI ABIOLA

71

Sp

ecif

ic g

rav

ity

T (oC)

Fig. 4.5: Graph of Specific Gravity of CSO, PKO and GNO at different

temperatures

GNO

CSO

PKO

0.80

0.78

0.86

0.84

0.90

0.82

0.88

40 30 50 60 70 80

PKO

CSO

GNO

Page 84: LATEEF, FATAI ABIOLA

72

The viscosities of castor oil biodiesel esters were in the range of 6.89 –

8.12 mm 2

/s (see table 4.6) which were higher than that of European

biodiesel standards (EN 14214:3.5 – 5.0) and American Society of Testing

Material Standards (ASTM D6751. 1.9 – 6.0). Knothe and Steidley (2005)

reported that the introduction of an OH group significantly increases

viscosity and this is of significance for production of castor oil-based

biodiesel, a fuel that in its neat form exceeds all kinematic viscosity

specifications in biodiesel standards due to the high content of ricinoleic acid

in castor oil. In order to solve this problem, biotechnological interventions

for improving castor for biofuels have evolved. In castor oil, though beyond

the limit of this experimental work, transgenic technology approach has been

proposed for reduction of the toxic protein ricin and conversion of ricinoleic

(12-hydroxyoleic acid) rich castor oil to oleic rich oil (Sujatha, 2009).

However, the limit values of viscosity can be met through

transesterification followed by dilution or blending with conventional diesel

fuel and vegetable oil (Sujatha, 2009). Meanwhile palm kernel and

groundnut biodiesel esters conform to ASTM D 6751 standards.

4.8 CASTOR OIL BIODIESEL/PETRODIESEL BLENDING

In order to obtain acceptable kinematic viscosity result for castor oil,

the biodiesels produced were blended with petrodiesel in the following

manner.

Petrodiesel (90%) – biodiesel (10%) : B10

Petrodiesel (80%) – biodiesel (20%) : B20

Biodiesel (propyl esters) : B100

Results obtained are shown in Table 4.9.

Page 85: LATEEF, FATAI ABIOLA

73

Table 4.9: Comparison of Kinematic viscosity and Specific Gravity of

Castor oil Biodiesel and Petrodiesel Blend with Unblended Biodiesel

Properties B10 B20 B100 P100

Specific gravity (28 oC) 0.85 0.86 0.87 0.83

Kinematic viscosity mm2/s (40

oC) 4.24 4.98 7.21 2.85

The properties of the B100, B20 and B10 mixtures are comparable to

those of petroleum diesel P100 and acceptable within what is specified for

biodiesel in the ASTM D6751 standards (with the exception of viscosity of

B100).

4.9 Kinetic Studies at Various Temperatures

This section discusses the effects of temperature and reaction time on

biodiesel yield and conversion. As evident from Appendix III, volume of

biodiesel and percentage (%) biodiesel conversion increases with respect to

time and temperature. Figures 4.6 - 4.9 shows the progress of the biodiesel

conversion for castor, palm kernel and groundnut oils.

With the reaction time increased to 20mins, an improved biodiesel

conversion of 66% was achieved for castor oil (CSO) biodiesel at 32oC. This

trend continued with reaction time up to 60mins with conversion of 89%.

Similar results were achieved for PKO and GNO and the same trend of the

graph obtained. This implies that within the time range of 5–60 mins, CSO,

PKO and GNO biodiesels yield increased with reaction time. It was also

observed that biodiesel conversion of oils were higher as temperature was

increased to 45, 60 and 75oC for the biodiesels (CSO, PKO and GNO).

The graphs of percentage conversion of oils against time were shown

in Figures 4.6 – 4.9.

Page 86: LATEEF, FATAI ABIOLA

74

0 10 20 30 40 50 6030

40

50

60

70

80

90

100

Time [Min]

Pe

rce

nta

ge

[%]c

on

ve

rsio

n a

t 3

0[d

eg

]

GNOGNOPKOPKO

CSOCSO

0 10 20 30 40 50 6040

50

60

70

80

90

100

Time [Min]

Pe

rce

nta

ge

[%]c

onve

rsio

n a

t 4

5[d

eg

]

GNOGNOPKOPKO

CSOCSO

Fig 4.6: Graph of Percentage (%) Conversion of biodiesel against time (min)

Kinetic result at 32oC

Fig 4.7: Graph of Percentage (%) Conversion of biodiesel against time (min)

Kinetic result at 45oC

Page 87: LATEEF, FATAI ABIOLA

75

0 10 20 30 40 50 6040

50

60

70

80

90

100

Time [Min]

Pe

rce

nta

ge

[%]c

on

ve

rsio

n a

t 6

0[d

eg

.]

CSOCSO

PKOPKOGNOGNO

0 10 20 30 40 50 6040

50

60

70

80

90

100

Time [Min]

Pe

rce

nta

ge

[%]c

on

ve

rsio

n a

t 7

5[d

eg

.]

GNOGNO

CSOCSO

PKOPKO

Fig 4.9: Graph of Percentage (%) Conversion of biodiesel against time (min)

Kinetic result at 75oC

Fig 4.8: Graph of Percentage (%) Conversion of biodiesel against time (min)

Kinetic result at 60oC

Page 88: LATEEF, FATAI ABIOLA

76

4.10 Biodiesel Storage/Oxidative Stability Measurement

In the oxidative stability test, the biodiesels produced at 65oC using

alkali-transesterification methods (see section 3.7) were exposed to light and

air. The peroxide value were determined at one month interval each (herein

referred to as run)

The most significant and undesirable change in liquid fuel with time is

the formation of solids, also termed filtrate sediments. During long-term

storage oxidation due to contact with air (autooxidation) presents a

legitimate concern with respect to maintaining fuel biodiesel quality (Ferrari

et al, 2005). The progress of the oxidation was monitored by measuring the

peroxide value or the fraction of the biodiesel/petrodiesel that has been

converted to a peroxide molecule.

Table 4.10 shows oxidative stability results of the biodiesel from

castor, palm kernel, groundnut oils and conventional diesel (Petrodiesel)

samples evaluated through their peroxide values. The results were taken over

a period of three months (a month interval for each run) at room temperature

28–30oC. Gerpen et al (2004) reported that biodiesel have no conversion for

a period of time, due to presence of natural antioxidant in the biodiesel oil

sample (three months – called the induction period) but then oxidize quickly.

Table 4.10: Peroxide Values of Castor, Palm kernel, Groundnut Oils and

Petrodiesel

Peroxide Value (mmol peroxide/kg sample)

Sample Run I Run II Run III

CSO Biodiesel 80 80 81

PKO Biodiesel 20 40 50

GNO Biodiesel 10 50 60

Petrodiesel Neutral Neutral Neutral

CSO – Castor oil, PKO – Palm kernel oil and GNO – Groundnut oil

Page 89: LATEEF, FATAI ABIOLA

77

The results obtained showed that conventional diesel (Petrodiesel)

showed greater stability than the biodiesel produced from castor, palm kernel

and groundnut oils – as it showed no result (trace) when peroxide value was

determined. This is because petroleum based diesel fuels (Petrodiesel) are

treated with a wide range of additives to improve lubricity, oxidative

stability, corrosion resistance and many other properties unlike biodiesel

which contain more or less unsaturated fatty acids in its compositions which

are susceptible to oxidation reactions, accelerated by exposition to oxygen,

being able to change to polymerize compounds (Meher et al, 2004).

Next to the petrodiesel which showed greater stability is castor oil

biodiesel, followed by palm kernel and groundnut oils diesel. For castor oil

biodiesel, the peroxide values remained constant for run I and II – 80mmol

peroxide/kg sample and increases slightly at run III – 81mmol peroxide/kg

sample. Sujatha (2009) reported that the presence of hydroxyl group and

double bonds impacts unique chemical and physical properties that make

castor oil a vital raw material and stabilizes the oil against oxidation. Castor

oil has a good shelf life when compared to other vegetable oils and it does

not turn rancid when subjected to excessive heat.

The peroxide value results taken for the samples shows that castor oil

biodiesel can withstand oxidation after two months of production unlike

palm kernel oil biodiesel whose peroxide value increases sharply from

20mmol peroxide/kg sample (run I) to 40mmol peroxide/kg sample (run II)

and then in the third run increases to 50mmol peroxide/kg sample. The

results from groundnut oil biodiesel show the worst oxidative stability as its

peroxide value increase sharply from 10mmol peroxide/kg sample to

50mmol peroxide/kg sample in the second run and the 60mmol peroxide/kg

sample in the third run. Gerpen et al (2004) reported that, oxidation is an

autocatalytic process so that when it starts, it progresses at ever-increasing

rate and this were exhibited by PKO and GNO biodiesel, but the presence of

Page 90: LATEEF, FATAI ABIOLA

78

hydroxyl group and double bonds in castor oil stabilizes the oil against

oxidation.

4.11 Infrared Spectroscopy-Biodiesel Product Analysis Results

The infrared spectra of a B100 biodiesel (Fatty Acid Propyl Esters

FAPE) from castor, palm kernel and groundnut oils are shown in Figures

4.10-4.15. The strong ester peaks near 1750 cm-1

(the C=O vibration) and

around 1170-1200 cm– 1

(the C–O vibration) were clear and are the basis for

the procedure. There is no interference in the spectra for CSO, PKO and

GNO.

The spectrum, region, Absolute threshold, sensitivity, peak list

positions are shown in Figures 4.10– 4.15.

Page 91: LATEEF, FATAI ABIOLA

79

Fig. 4.10: Percentage Transmittance versus Wavenumber (cm-1

) for CSO

79

Page 92: LATEEF, FATAI ABIOLA

80

Fig. 4.11: Esters, Olefins and Aliphatic Hydrocarbons Percentage Transmittance versus

Wavenumber (cm-1

) for CSO

80

Page 93: LATEEF, FATAI ABIOLA

81

Fig. 4.12: Percentage Transmittance versus Wavenumber (cm-1

) for PKO

81

Page 94: LATEEF, FATAI ABIOLA

82

Fig. 4.13: Esters, Olefins and Aliphatic Hydrocarbons Percentage Transmittance versus

Wavenumber (cm-1

) for PKO

82

Page 95: LATEEF, FATAI ABIOLA

83

Fig. 4.14: Percentage Transmittance versus Wavenumber (cm-1

) for GNO

94

94

83

Page 96: LATEEF, FATAI ABIOLA

84

Fig. 4.15: Esters, Olefins and Aliphatic Hydrocarbons Percentage Transmittance versus

Wavenumber (cm-1

) for GNO

84

Page 97: LATEEF, FATAI ABIOLA

85

CHAPTER FIVE

CONCLUSION AND RECOMMENDATIONS

5.1 CONCLUSION

The population growth, ever increasing use of transport fuels, rising

prices of fossil fuel, climate change and environmental pollution demands

use of renewable energy sources for a more sustainable energy solution. Vast

scope exists for exploitation of castor, palm kernel and groundnut oils as

bioenergy crops (although there are still some technological challenges to

overcome especially for castor oil biodiesel).

The following conclusions could be drawn from the present study:

i. Biodiesels has been produced in a batch reactor using alkali-catalyzed

method.

ii. In order to obtain biodiesel, transesterification process has been

studied. Transesterification time controls the yield of product while

purification is fundamental in order to fulfill the characteristics of

propyl esters (biodiesel) as fuel.

iii. Temperature has noticeable effect on the transesterification process.

iv. Specific gravity of the biodiesel samples were found to be higher than

that of petrodiesel and that specific gravity of biodiesel samples were

temperature dependent. Also, specific gravities of castor oil biodiesel

blend (B10 and B20) were lower than that of unblended biodiesel.

v. Calorific values of the biodiesel samples were lower than that of

petrodiesel combusted under the same conditions of experiment.

vi. Biodiesel yield increases with reaction time up to 60 minutes.

vii. The viscosities of biodiesel samples decrease as temperature increases.

Both palm kernel and groundnut oils biodiesel conforms to kinematic

viscosities ASTM D6751 and EN 14214 specification standards.

viii. Castor oil biodiesel has the highest storage stability.

Page 98: LATEEF, FATAI ABIOLA

86

ix. The infrared analysis produces primary absorbances that give

fundamental knowledge of the type of the chemical group present in

the biodiesels.

Lastly, relying on fossil fuel alone is no longer realistic due to global

depletion of the non-renewable energy sources, its attendant negative

environmental impact. The race for energy security in the face of imminent

oil shortage is already gathering momentum. Countries in Asia, Europe,

South American and many US state governments are not waiting for their

fossil fuel to dry up completely before searching for alternative, and only

countries that don’t value their own security and that of their citizens would

stand aloof.

5.2 RECOMMENDATION

High viscosity limits the widespread use of castor oil as alternative to

be used in diesel engine as this report showed that castor oil biodiesel

viscosity exceeds both European/American and all other specification

standards. Likewise, in castor seed, research effort should be proposed for

reduction of the toxic protein ricin and conversion of ricinoleic acid rich

castor oil to oleic rich oil.

In future, a lot of work can be done to reduce the cost of biodiesel if

we consider non-edible oils, used frying oils instead of edible oils. Non

edible oils such as Neem, Karanja, Jatropha etc are easily available in

Nigeria and very cheap compared to edible oils. With the mushrooming of

fast food centres and restaurants in Nigeria, it is expected that considerable

amount of used frying oils will be discarded. This can be used for making

biodiesel.

All potential feedstocks for biofuel (biodiesel) production are in

abundance in the country (Nigeria). Nigeria, with her expansive arable land

mass, can be one of the world’s leading exporters of biodiesel, if the

Page 99: LATEEF, FATAI ABIOLA

87

government puts a premium on energy security like many countries (such as

US and some European countries) are now doing. However, there are fears

that since biodiesel relies on primary agricultural products, a substantial

growth in the biodiesel industry could make the prices of vegetable oil

unaffordable to the common man. Hence, our approach to renewable energy

sources should be gradual.

Lastly, government should provide funding, enabling environment and

an enticing package of incentives. This involves providing comprehensive

policy support and funding for research in the area of renewable energy

source (biofuels).

Page 100: LATEEF, FATAI ABIOLA

88

REFERENCES

1. Ajav, E. A. and Akingbehin, A. O. (2002). A Study of Some Fuel

Properties of Local Ethanol Blended with Diesel Fuel. Agricultural

Engineering International: the CIGR Journal of Scientific Research

and Development. Manuscript EE 04 004 Vol. VI.

2. Alais, C. and Linden, G. (1999). Food Biochemistry. Aspen

Publication, Gaithersburg, Maryland. pp. 67 – 70.

3. Alamu, O. J. (2007b). Development and Evaluation of Palm Kernel

Oil Biodiesel Fuel. The Pacific Journal of Science and Technology.

Vol 8, No. 2. Retrieved 20 May 2008 from - http://www.akamai

university.us/PJST.htm

4. Alamu, O.J.; Akintola, T. A.; Enweremandu, C.C. and Adeleke, A. E.

(2008). Characterization of Palm-Kernel Oil Biodiesel Produced

Through NaOH – Catalyzed Transesterification Process. Scientific

Research and Essay Vol. 3(7). pp. 308 –311, Retrieved 20th May 2008

from - http:// www.academicjournals.org/SRE.

5. Alamu, O.J.; Waheed, M.A.; Jekayinfa, S.O. and Akintola, T.A.

(2007a). Optimal Transesterification Duration for Biodiesel

Production from Nigerian Palm Kernel Oil. Agricultural Engineering

International. The CIGR Ejournal. Manuscript EE 07018, Vol. IX.

6. Asiedu, J. J. (1989). Processing Tropical Crops. A Technological

Approach. Macmillan Education Ltd, London. pp. 124 – 171.

7. Atkins, P. and Paula, D. J. (2002). Atkins’ Physical Chemistry. 7th

Edition, Indian. pp. 908.

8. Bamgboye, A. I. and Hansen, A. C. (2008). Prediction of Cetane

Number of Biodiesel Fuel from the Fatty Acid Methylester

Composition. Institute of Agrophysics, 22, 21 – 29. Retrieved 20th

May

2008 from - www.international-agrophysics.org

Page 101: LATEEF, FATAI ABIOLA

89

9. Baroutian, S.; Aroua, M.K.; AbdulRaman A. and Sulaiman, N.M.

(2008). Prediction of Palm Oil. Based Methyl Ester Biodiesel Density

Using Artificial Neural Network. Journal of Applied Science 8 (10):

1938 – 1993.

10. Brandley, M. (2007). Biodiesel (FAME) Analysis by FT-IR.

Application Note: 51258, Thermo Fisher Scientific, Madison, USA.

11. Canakci, M, and Gerpen, J. V. (2001). Biodiesel Production from Oils

and Fats with High Free Fatty Acids. Journal of American Society of

Agricultural Engineers. Vol. 44 (6):1429 – 1436.

12. Chhetri, A.B.; Tango, M.S.; Budge, S.M.; Watts, K.C. and Islam, M.R.

(2008). Non-Edible Plant Oils as New Sources for Biodiesel

Production. International Journal of Molecular Sciences. Vol. 9, pp.

169–180. Retrieved 2nd

December 2008 from - http://www.

mdpi.org/ijms

13. Darnoko, D. and Cheryan, M. (2000). Kinetics of Palm Oil

Transesterification in a Batch Reactor. Journal of American Oil

Chemists’ Society (JAOCS). Retrieved 2nd

May 2008 from -

http://www.fineprint.com

14. Darnoko, D. and Cheryan, M. (2000b). Continuous Production of

Palm Methyl Esters. Journal of American Oil Chemists’ Society

(JAOCS), Vol. 77, No. 12. pp. 1269 – 1272.

15. David, P. (1976). The Chemical Analysis of Food. 7th Edition,

Churchill Livingstone. Edinburgh, London. pp. 489 – 516.

16. Dunn, R. O. and Bagby, M. O. (1995) Low-Temperature Properties of

Triglyceride – Based Diesel Fuels: Transesterified Methyl Esters and

Petroleum Middle Distillate Ester Blends. Journal of American Oil

Chemists Society, Vol. 72, No. 8, pp. 895 – 904.

Page 102: LATEEF, FATAI ABIOLA

90

17. Eckhardt, J. and Kawaguchi, N. (1997). Oil Processing Retrieved.

June 15th

2008 from http://www.oilprocessing/11A:Edibleand

EssentialOil Processing.htm.

18. Energy Information Administration (EIA) (2007). World Proved

Reserved of Oil and Natural Gas. Most Recent Estimates. Retrieved

20th

May 2008 from - www.eid.doe.gov

19. El Diwani, G. and El Rafie, S. (2008). Modification of Thermal and

Oxidative Properties of Biodiesel Produced from Vegetable Oils. Inter.

J. of Environ. Sci. Tech., 5(3), 391-400.

20. Ferrari, A. R; Oliveira, S. D. V. and Scabio, A. (2005). Oxidative

Stability of Biodiesel from Soybean Oil Fatty Acid Ethyl Esters.

Scientia Agricola, Vol. 62, No. 3. Piracicaba.

21. Fischer, J.C. (1998). Biodiesel in Europe: Biodiesel Processing

Technologies. Paper Presented at International Liquid Biofuels

Congress, July 19-22, Curitiba, Parana, Brazil.

22. Fogler, S. H. (2006). Elements of Chemical Reaction Engineering. 4th

Edition. Pearson Education International, United States. pp. 10 – 40.

23. Fukuda H.; Kondo, A. and Noda, H. (2001) Biodiesel Fuel Production

by Transesterification of Oils. Journal of Bioscience and

Bioengineering. Vol. 92, No. 5, pp. 405 – 416.

24. Gerpen J. V. (2005). Biodiesel Processing and Production. Fuel

Processing Technology 85 1097–1107. Retrieved 2nd

December 2008

from - www.elsevier.com/locate/ fuproc.

25. Gerpen, J. V.; Shanks, B. and Pruszko, R. (2004). Biodiesel

Production Technology. National Renewable Energy Laboratory

(NREL) Colorado. Sub-Contractor Report. pp. 1 – 74.

26. Gupta, A.; Sharma, S. K. and Toor, A. P. (2007) Production of

Biodiesel from Waste Soybean Oil, Journal of Petrotech Society. pp.

40 – 45.

Page 103: LATEEF, FATAI ABIOLA

91

27. Ibitoye, A. A. (2006). Laboratory Manual on Basic Method in Plant

Analysis. Concept IT, Akure, Nigeria. pp. 20 – 22.

28. Jacobs, M. B. (1999). The Chemical Analysis of Foods and Food

Production. 3rd ed.: CBS Publisher, New Delhi, India. pp. 15 – 16.

29. Jitputti, J.; Kitiyanan, B.; Bunyakiat, K.; Rangsunvigit, P. and

Jenvanitpanjakul, P. (2004). Transesterification of Palm Kernel Oil

and Coconut Oil by Difference Solid Catalysts. The Joint International

Conference on “Sustainable Energy and Environment, 1 – 3 December,

2004, Hua, Thailand.

30. Joslyn, A. M. (1970). Methods in Food Analysis, 2nd edition,

Academic Press, New York. San Francisco, London. pp. 226 – 264.

31. Khan, A. K.(2002). Research into Biodiesel Kinetics and Catalyst

Development. Unpublished M.Sc. Thesis, Department of Chemical

Engineering, University of Queensland Brisbane, Queensland,

Australia.

32. Knothe G. (2006). Analyzing Biodiesel; Standards and Other Methods.

Journal of American Oil Chemists Society JOACS, Vol. 83, No. 10.

33. Knothe, G. (1999). Rapid Monitoring of Transesterification and

Assessing Biodiesel Fuel Quality by Near-Infrared Spectrostrocopy

Using a Fiber-Optic Probe. Journal of American Oil Chemists Society

(JAOCS), Vol. 76, No. 7.

34. Knothe, G. (2001) Analytical Methods Used in the Production and

Fuel Quality Assessment of Biodiesel. Journal of American Society of

Agricultural Engineers. Vol. 44 (2). pp. 193 – 200.

35. Knothe, G. and Steidley, K. R. (2005) Kinematic Viscosity of

Biodiesel Fuel Components and Related Compounds. Influence of

Compound Structure and Comparison to Petrodiesel Fuel Components.

Elsevier. Fuel 84 (2005) 1059 – 1065. Retrieved 12th July 2008 from -

www.sciencedirect.com

Page 104: LATEEF, FATAI ABIOLA

92

36. Krause, V. M. and Hunscher, A. M. (1999 ). Food, Nutrition and Diet

Therapy. 5th Edition, W. S. Saunders Company, Philadelphia, London.

pp. 56 – 60.

37. Krisnangkura K.; Yimsuwan, T. and Pairintra, R. (2005). An

Empirical Approach in Predicting Biodiesel Viscosity at Various

Temperature. Elsevier, pp. 2.

38. Kyu-Wan L. Yu, J. X.; Mei, J. H.; Kim, Y.W.; Yan, L. and Chun K.

W. (2007) A Kinetic Study on the Transesterification of Glycery

Monooleate and Soybean Used Frying Oil to Biodiesel, Journal of

Industrial Engineering Chemistry, Vol. 13, No.5. 799-807.

39. Lebedevas, S. and Vaicekausas, A. (2006). Research into the

Application of Biodiesel in the Transport Sector of Lithuania.

Transport. Vol. XXI, No. 2. pp. 80 – 87.

40. Leevijit, T.; Wisutmethangoon, W.; Prateepchaikul, G.; Tongurai, C.

and Allen, M. (2004). A Second Other Kinetics of Palm Oil

Transesterification. The Joint International Conference on

“Sustainable Energy and Environment (SEE) (3 – 025(O)) 1–3

December 2004, Hua Hin, Thailand. pp 277 – 281.

41. Leevijit, T.; Wisutmethangoon, W.; Prateepchaikul, G.; Tongurai, C.

and Allen, M. (2004b). Transesterification of Palm oil in Series of

Continuous Stirred Tank Reactors. The Joint International Conference

on Sustainable Energy and Environment (SEE) (3 – 024(P)) 1–3

December 2004, Hua Hin, Thailand. pp 272 – 276.

42. Lewis, J. M. (1987). Physical Properties of Foods and Food

Processing Systems. Ellis Horwood Series in Food Science and

Technology.

43. Ma. F., and Hanna, M. A. (1999). Biodiesel Production: A Review.

Bioresource Technology, 70, pp. 1 -15. Published by Elsevier Science

B. V.

Page 105: LATEEF, FATAI ABIOLA

93

44. McCabe, L. W.; Smith, C. J. and Harriot, P. (2005). Unit Operations

of Chemical Engineering (7th Ed.) New York: McGraw-Hill

International Edition, pp. 45 – 54.

45. Meher, C. L, Sagar, V. D. and Nack, S. N. (2004). Technical Aspects

of Biodiesel Production by Transesterification – A Review. Centre for

Rural Development and Technological, Indian Institute of Technology,

Delhi, New Delhi, India. pp. 1 – 21.

46. Meyer, H. L. (2004). Food Chemistry. CBS Publisher, New Delhi,

India. pp. 13 – 25.

47. Murray, R.S..P. (1977). Principles of Organic Chemistry. 2nd ed.

Heinemann Educational Books, London. pp. 246 – 248.

48. Narvaez, P. C.; Castaneda, L. Z.; Rineon, S. M.; Sanchez, F.; and

Torres, J. (2006). Continuous Process for the Production of Fatty Acid

Methyl Esters in a Falling Liquid-Liquid Film Reactor. Universidad

Nacional de Colombia.

49. Nigerian National Petroleum Corporation Newsletter (2009).

Pipeliners: An In-House Second Quarterly Newsletter of the Pipelines

and Products Marketing Company Limited. pp. 16-17.

50. Noureddini, H. and Zhu, D. (1997). Kinetics of Transesterification of

Soybean Oil: Biocatalysis Article. Journal of American Oil Chemists’

Society JOACS 74, 1457 – 1463.

51. Onimawo, I. A. and Akubor, P. I. (2005). Food Chemistry. Ambik

Press Ltd. Benin City, Edo State, Nigeria. pp. 57 – 80.

52. Pimentel, D. and Patrik, T. W. (2005). Ethanol Production Using Corn,

Switchgrass and Wood; Biodiesel Production Using Soybean and

Sunflower. Natural Resources Research. Vol. 14, No. 1,

DO1:10.1007/S/1053 – 005 – 4678 – 8.

Page 106: LATEEF, FATAI ABIOLA

94

53. Refaat, A. A, Attia, N. K.; Sibak, H. A.; ElSheltawy, S. T. and

ElDiwani, G. I. (2008) Production Optimization and Quality

Assessment of Biodiesel from Waste Vegetable Oil. International

Journal of Environmental Science and Technology, Vol. 5 (1), 75 – 82.

54. Ritz, P. and Nash, L. (2004). Catching the Wave IR Applications for

Biofuels. PAC. pp. 1-36.

55. Royon, D.; Daz, M.; Ellenrieder, G. and Locatelli, S. (2006).

Enzymatic Production of Biodiesel from Cotton Seed Oil using t-

butanol as a Solvent. Journal of Bioresource Technology. 98 648-653.

Elsevier. Retrieved 18th

August, 2008 from www.sciencedirect.com.

56. Schuchardt U.; Sercheli, R. and Vargas, R. M. (1998)

Transesterification of Vegetable Oils: A Review. Journal of Brazilian

Chemical Society. Vol. 9, No.1. 119-210.

57. Shah S., Sharma, S. and Gupta, M. N. (2004) Biodiesel Preparation by

Lipase-Catalyzed Transesterification of Jatropha Oil. Journal of

American Chemical Society, Energy and Fuels, 18, 154 – 159.

58. Sharma, Y.C; Singh, B. and Updahyay, S.N. (2008). Advancements in

Development and Characterization of Biodiesel: A Review. Fuel 87,

2355-2372. Available online at www.sciencedirect.com

59. Sivansankar, B. (2008). Food Processing and Preservation. Prentice

Hall of India Private Ltd., New Delhi. pp. 45 – 50.

60. Sujatha, M. (2009). Biotechnological Interventions for Improving

Jatropha and Castor for Biofuels. Petrotech, New Delhi, India. (P09 –

869).

61. Tapasvi, D.; Wiesenborn, D. and Gustafson, C. (2005) Process Model

for Biodiesel Production from Various Feedstock. Journal of

American Society of Agricultural Engineers, Vol. 48(6): 2215-2221.

Page 107: LATEEF, FATAI ABIOLA

95

62. Titipong I. (2006). Biodiesel Production from Fryer Grease.

Unpublished M.Sc. Thesis, Submitted to the College of Graduate

Studies and Research, Department of Chemical Engineering,

University of Saskatchewan, Saskatoon, Saskatchewan.

63. Turner, T. L. (2005). Modelling and Simulation of Reaction Kinetics

for Biodiesel Production. Unpublished M.Sc Thesis, Submitted to the

Graduate Faculty of North Carolina State University, Department of

Mechanical Engineering, North Carolina State University, USA.

64. USEIA (2009). Energy Information Administration World Proved

Reserves of Oil and Natural Gas Most Recent Estimates.

www.eia.doe.gov Visited March 3, 2009.

65. Wawrzyniak, R.; Wasiak, W. and Frackowiak, M. (2005).

Determination of Methyl Esters in Diesel Oils by Gas

Chromatography-Validation of the Method. Chem. Paper 59 (65) 449

– 452; Presented at the 32nd International Conference of the Slovak

Society of Chemical Engineering, Tatranske Matliare.

66. Wenzel, B.; Tait, M.; Modenes, A. and Kroumov, A. (2006).

Modelling Chemical Kinetics of Soybean Oil Transesterification

Process for Biodiesel Production: An Analysis of Molar Ratio

Between Alcohol and Soybean Oil Temperature Changes on the

Process Conversion Rate; Bioautomation, 5, 13 – 22.

67. Wikipedia, the free Encyclopedia, Castor Oil. Retrieved on August

23rd 2008 from: http://www.en.wikipedia.org/wiki/castoroil

68. Wikipedia, the Free Encyclopedia. Biodiesel: A Historical

Background. Retrieved on August 23rd 2008 from

http://www.en.wikipedia.org/ wiki/biodiesel.

Page 108: LATEEF, FATAI ABIOLA

96

69. Yuan, Y. A. and Zhang, Q. (2004). The Specific Gravity of Biodiesel

Fuels and their Blends with Diesel Fuel. Agricultural Engineering

International: The CIGR Journal of Scientific Research and

Development. Manuscript EE 04 004, Vol. VI.

70. Zhang Y.; Dube, M A.; McLean, D.D. and Kates, M. (2003) Biodiesel

Production from Waste Cooking Oil: Process Design and

Technological Assessment. Bioresource Technology 89, 1-16.

www.sciencedirect.com

71. Zhao, X; El-Zahab, B.; Broznahan, R. (2007). An Organic Soluble

Lipase for Water – Free Synthesis of Biodiesel. Journal of Applied

Biochemistry Biotechnology, Vol. 143: 236 – 243.

Page 109: LATEEF, FATAI ABIOLA

97

APPENDIX I

SAPONIFICATION VALUES (S.V)

These values were calculated from our expression in the procedure

S.V = (b – a) x 28.05

Weight (in g) of sample

For example,

GNO biodiesel = 28.05 (24.00 – 17.58)

1

= 180.08mg KOH/g

IODINE VALUES (I.V)

The iodine values were calculated from the expression as outlines in

the procedure.

I.V = (b – a) x 1.269

Weight (in g) of sample

For example,

Crude PKO = (49.00 – 39.00) x 1.269

1

= 10 x 1.269

= 12.69g of iodine/g

ACID VALUES (A.V)

The acid values were estimated from the expression in the procedure.

AV = Titration (ml) x 5.61

Weight of sample used

For example,

Refined GNO = 0.28 X 5.61

1

= 1.57mg NaOH/g

Page 110: LATEEF, FATAI ABIOLA

98

PEROXIDE VALUE (P.V):

Peroxide value were calculated using the formula

Peroxide value (P.V) = T x M x 1000

Sample weight (g) mmol peroxide/kg sample

where T – titre value of Na2S2O3

M – Molarity of Na2S2O3

For example, castor oil biodiesel

Initial burette reading = 38.50

Final burette = 38.20

P.V = (38.50 – 38.20) x 0.1 x 1000

1

= 30 mmol peroxide/kg sample

Viscosities:

Dynamic viscosity (cP) = 102Mf.Ir

Kinematic viscosity (mm2/s) = 10

2Mf.Ir/p (g/cm

3)

where Mf - Multiplication factor, peculiar for each cylinder

Ir – Instrument reading (Poise)

For example CSO kinematic viscosity is calculated as

Dynamic viscosity µ (cP) = 102 (0.0087) 39.5

Kinematic viscosity V (mm2/s) = µ (Dynamic viscosity)

ρ (sample density)

= 102 (0.0087) (39.5)

0.96

= 35.78 mm2/s

SPECIFIC GRAVITY (S.G)

S.G were calculated from the expression

S.G = W2 – Wb

W1 – Wb

where W2 = weight of density bottle + sample

W1 = weight of density bottle + water

Page 111: LATEEF, FATAI ABIOLA

99

For example, S.G of Petrodiesel;

Volume of density bottle = 25ml

weight of Density bottle = 24.18g

weight of Density bottle + water = 49.07g

weight of Density bottle + Petrodiesel = 45.07g

S.G of Petrodiesel = 45.07 – 24.18 = 20.89

49.23 – 24.18 25.85

= 0.83

CALORIFIC/COMBUSTION VALUES:

Calorific values were calculated from the expression

W = E∆T – Φ – V

m

where E = Benzoic acid standard = 13,039.308 cal/g

∆T = Temperature difference = (Tf – To)

W = Calorific value of the sample (calorie/g)

m = weight of the sample (g)

Φ = 2.3L (where L = Length of the unburnt wire)

V = titre value of the Na2CO3 solution

The combustion value were converted to Joules, 1 calorie = 4.148 Joules

For example, the calorific value of PKO biodiesel in Table 4.6 was

obtained thus:

W = 13.039.308 (0.75) – 16.21 – 5.32 cal/g

1.048

= 38.62 MJ/kg

Page 112: LATEEF, FATAI ABIOLA

100

APPENDIX II

Percentage (%) yield of the oil = Weight of extracted oil x 100%

Weight of oil seeds

The percentage weight per volume (% wt/v) of the catalyst was

calculated based on the formula.

% wt/v = weight of catalyst x 100%

volume of oil

and that of percentage yield of biodiesel

% yield of biodiesel = Volume of (CSO, PKO, GNO biodiesel) x 100%

Volume of oil

% loss on pretreatment = Weight of unrefined oil – Weight of refined oil x 100%

Weight of unrefined oil

Conversion of oils

The conversions of oils used in Table 4.7 were based on the

stoichiometric balance, that 100cm3 of oil gives approximately 100cm

3 of

methyl esters.

For example

The conversion of CSO is calculated as

89.50 = 0.895

100

= 85.5%

Average Fatty Acids Compositions

Oil sample CSO PKO GNO Relative Molecular Mass

of Fatty Acids

Capric _ 3.5 _ 172

Caprylic _ 4.0 _ 144

Lauric _ 48.0 _ 200

Myristic _ 16.0 _ 228

Palmitic 0.5 8.0 11.6 256

Stearic 0.5 3.0 3.1 284

Oleic 2.0 15.0 48.5 282

Linolenic 1.0 _ 31.4 278

Linoleic 1.0 2.0 _ 278

Ricinoleic 95.0 - - 155

Dihydroxystearic 0.3 - - 302

Page 113: LATEEF, FATAI ABIOLA

101

PALM KERNEL OIL

Relative molecular mass (RMM)

RMM = MA x Abundance (%) + MB x Abundance (%) + …

100 100

Capric = CH3(CH2)8COOH

= 12 + 3 + (12 + 2)8 + 12 + 32 + 1

= 15 + 112 + 12 + 33 = 172

Lauric = CH3(CH2)10COOH

= 15 + (12 + 2)10 + 12 + 32 + 1

= 15 + 140 + 12 + 33 = 200

Myristic = CH3(CH2)12COOH

= 15 + 168 + 12 + 32 + 1 = 228

Palmitic = CH3(CH2)14COOH

= 15 + 196 + 12 + 33 = 256

Stearic = CH3(CH2)16COOH

= 15 + (12 + 2)16 + 12 + 32 + 2

= 15 + 224 + 12 + 33 = 284

Oleic acid = CH3(CH2)7CH=CH(CH2)7COOH

= 15 + (12 + 2)7 + 12 + 1 + 12 + 1(12 + 2)7 + 12 + 32 + 2

= 282

Linoleic acid = CH3(CH2)4CH=CHCH2COOH

= CH(CH2)7COOH

= 15+(12 + 2)4+12+1+12+1+14+13+13+(14)7+12+32+1

= 278

RMM =

172 x 4 + 200 x 48 + 228 x16 +256 x 8 +284 x 3 + 282 x 15 + 278 x 2

100 100 100 100 100 100 100

= 6.88 + 96 + 36.48 + 20.48 + 8.52 + 42.30 + 5.56

RMM = 216.22

Page 114: LATEEF, FATAI ABIOLA

102

Equation

3 fatty acid + CH2OH Triglyceride + 3H2O

CHOH

CH2OH

3 x 216.26 + 92 = X + 54

740.66 = X + 54

= 686.66 MM of Triglyceride

MM = Molecular mass

Now, for biodiesel preparation

The equation for biodiesel is

Triglyceride + 1-Propanol Alkylesters + glycerol

683.21 + 180

Palm kernel oil

Density = mass/volume

0.9168 = 686.66/vol

Volume = 686.66

0.9168

Volume = 748.97ml or cm3

For alcohol (1-propanol)

0.8 = 180/volume

volume = 180/0.8

= 225ml

748.97 : 225

For every 100ml of oil 100 : X

Therefore by cross multiplication

X = 22500 = 30.04 ≈ 30.00

748.97

Page 115: LATEEF, FATAI ABIOLA

103

for 1 : 6 oil-propanol basis

oil Propanol

100 30ml x 2 = 60ml

1 : 6

CASTOR OIL

1. Ricinoleic acid

CH3(CH2)5CHOHCH2CH=CH

12+3+(12+2)5+12+1+16+1+12+2+12+1+12+1

85 + 30 + 27 + 13 = 155

2. Oleic acid

CH3(CH2)7CH=CH(CH2)7COOH

12+3+(12+2)7+12+1+12+1+(12+2)7+12+32+1

15 + 98 + 13 + 13 + 98 + 12 + 33 = 282

3. Linoleic acid

CH3(CH2)4CH=CHCH2CH=CH(CH2)7COOH

2800

4. Linoleic acid: 278

5. Stearic acid: 284

6. Palmitic acid: 256

7. Dihydrostearic acid: 302

8. Others: 1

Relative molecular mass

155 x 95 + 282 x 6 + 280 x 6 +280 x 8 +278 x 1 + 284 x 1 + 256 x 1 +

100 100 100 100 100 100 100

302 x 0.5 = 147.25 + 16.92 + 14 + 2.78 + 2.84 + 2.56 + 1.51 = 187.86

100

Density of CSO = 46.20/50 = 0.9240 gcm –3

Page 116: LATEEF, FATAI ABIOLA

104

Castor oil relative molecular mass

3 fatty acid + CH2OH Triglyceride + 3H2O

CHOH

CH2OH

3 x 181.86 + 92 X + 54

563.58 + 92 X + 54

X = 693.58 MM of Triglyceride

MM = Molecular mass

Now, for biodiesel production

Triglyceride + Propanol Alkylesters + glycerol

For castor

Density = mass of Triglyceride/ volume

0.9240 = 693.58/volume

volume = 693.58/0.9240

= 750.12ml or cm3

For Alcohol (1-propanol)

0.812 = 180/Volume

Volume = 180/0.812

Volume = 221.67

750.12ml : 225

For every 100ml, 100 : X

X = 22500 = 29.99 ≈ 30ml or cm3

750.12

for 1 : 6 oil-propanol basis

oil Propanol

100 30ml x 2 = 60ml

1 : 6

For every 100ml of oil, 60ml of 1-propanol is required

Page 117: LATEEF, FATAI ABIOLA

105

Groundnut Oil

Palmitic: CH3(CH2)14COOH

256

Stearic: CH3(CH2)16COOH

284

Oleic: CH3(CH2)7CH=CH3(CH2)7COOH

282

Linolenic: CH3CH2CH=CHCH2CH=CHCH2CH=CH(CH2)7COOH

278

Relative molecular mass =

256 x 11.6 + 284 x 3.1 + 282 x 48.5 +278 x 31.4

100 100 100 100

= 29.696 + 8.804 + 136.77 + 87.292

Equation

3 fatty acid + CH2OH Triglyceride + 3H2O

CHOH

CH2OH

3 x 262.562 92 X + 54

X = 825.68 MM of Triglyceride

MM = Molecular mass

Density = mass/volume

0.927 = 825.68

Volume

Volume = 825.68

0.927

= 890.70ml or cm3

Page 118: LATEEF, FATAI ABIOLA

106

For alcohol (1-Propanol)

0.8 = 180/Volume

Volume = 225ml

890.70 : 225

For every 100ml, 100 : X

X = 22500/890.70

X = 25.26

Oil : Propanol

100 : 25ml x 2 = 50ml

1 : 3x2

1 : 6

Oil Alcohol ratio

This implies that for every 100ml of oil, 50ml of 1-propanol is required

Mobile phase = Acetonitrile : Acetone (59:41)

Flow rate = 1.0 ml/min

Detector = 2.15µm (UV)

Column = C8 (4.2 x 100mm, 3.5µ)

Amount injected = 20 µl (microlitre)

Standard concentration of fatty acid (%)

Lauric = 0.15

Palmitic = 0.25

Oleic = 0.63

Myristic = 0.17

Stearic = 0.07

Linoleic = 0.63

Arachidic = 0.19

Page 119: LATEEF, FATAI ABIOLA

107

Calculation of the relative molecular weight of triglycerides and

oil/alcohol molar ratio

Table 1

Sample CSO PKO GNO

Av. Molar wt of fatty acid 187.86 216.22 262.56

Density (cm3) 0.924 0.917 0.927

Relative molecular weight of

TG

693.58 686.66 825.68

Oil/Alcohol molar 1:6 ratio

for every 100ml of oil

60 60 50

Table 2

Samples CSO PKO GNO 1-Propanol

Average molar weight of

fatty acids

187.86 216.22 262.56

Density (cm3) 0.924 0.917 0.927

Relative molecular

weight of Triglycerides

693.58 686.66 825.68

Molar volume (m/s)

cm3/mol

750.63 748.81 890.70

Table 3

Samples

CSO PKO GNO

Molar volume (cm3/mol) 750.63 748.81 890.70

Molar reaction volume

(cm3)

750.63 748.81 890.70

Oil/Alcohol ratio for 100

cm3 of oil at

30 30 25

Oil/Alcohol molar ratio

at 1:6

for every 100 cm3 of oil

at 6:1

60 60 50

Density of 1-propanol = 0.812kg/m3

Page 120: LATEEF, FATAI ABIOLA

108

Table B2

Transesterification Process Parameters for Optimal Catalyst Weight Test of

Biodiesel at 1:3 Oil/Alcohol Molar Ratio

Experimental conditions Values for all the experiment

Castor, palm kernel and

groundnut oils quantity (cm3)

50.0

1-propanol quantity (cm3) 15.0 for CSO and PKO and 12.5 for GNO

Reaction Temperature (oC) 55

Weight of catalyst (NaOH)g Varies with reaction batches 1 – 8

Transesterification duration

(minutes)

60.0

Table B3

Alkali-catalyzed Batch Production of Biodiesel Parameter at 1:6 Oil/Alcohol

Molar Ratio

Experimental conditions Values for all the experiment

Castor, palm kernel and

groundnut oils quantity (cm3)

100.0

1-propanol quantity (cm3) 60.0 for CSO and PKO and 5.0 for GNO

Reaction Temperature (oC) 65

Weight of catalyst (NaOH)g 0.50 for CSO and PKO, 0.40 for GNO

Transesterification duration

(mins)

60

Page 121: LATEEF, FATAI ABIOLA

109

Table B4

Kinetics Studies Parameters for Biodiesel Production at 1:6 Oil/Alcohol

Molar Ratio

Experimental conditions Values for all the experiment

Castor, palm kernel and

groundnut oils quantity (cm3)

50.0

1-propanol quantity (cm3) 30.0 for CSO and PKO and 25.0 for GNO

Reaction Temperature (oC) Varies (30 – 75)

Weight of catalyst (NaOH)g 0.50 for CSO and PKO, 0.40 for GNO

Transesterification duration (min) Varies with experimental batches 1 - 4

Page 122: LATEEF, FATAI ABIOLA

110

APPENDIX III

Kinetic Studies Result at Various Temperatures

at 30oC

Time

(min)

Volume of biodiesel

(cm3)

Conversion of oils % Conversion

CSO PKO GNO CSO PKO GNO CSO PKO GNO

5 3.6 3.9 3.2 0.42 0.46 0.38 42.0 46.0 38.0

10 4.2 5.6 4.8 0.49 0.66 0.57 49.0 66.0 57.0

20 5.6 6.4 6.0 0.66 0.75 0.71 66.0 75.0 71.0

60 7.6 7.9 7.8 0.89 0.93 0.93 89.0 93.0 93.0

at 45oC

Time

(min)

Volume of biodiesel

(cm3)

Conversion of oils % Conversion

CSO PKO GNO CSO PKO GNO CSO PKO GNO

5 3.8 4.0 3.5 0.44 0.47 0.42 44.0 47.0 42.0

10 4.5 5.8 5.1 0.53 0.68 0.61 53.0 68.0 61.0

20 5.9 6.7 6.2 0.69 0.79 0.74 69.0 79.0 74.0

60 7.6 8.1 8.0 0.92 0.95 0.95 92.0 95.0 95.0

at 60oC

Time

(min)

Volume of biodiesel

(cm3)

Conversion of oils % Conversion

CSO PKO GNO CSO PKO GNO CSO PKO GNO

5 4.1 4.2 3.8 0.48 0.49 0.45 48.0 49.0 45.0

10 4.7 6.1 5.4 0.55 0.72 0.64 55.0 72.0 64.0

20 6.3 6.9 6.4 0.74 0.81 0.76 74.0 81.0 76.0

60 8.1 8.0 8.0 0.95 0.94 0.95 95.0 94.0 95.0

at 75oC

Time

(min)

Volume of biodiesel

(cm3)

Conversion of oils % Conversion

CSO PKO GNO CSO PKO GNO CSO PKO GNO

5 4.2 4.4 3.9 0.49 0.52 0.46 49.0 52.0 46.0

10 4.9 6.2 5.7 0.58 0.73 0.68 58.0 73.0 68.0

20 6.4 7.1 6.5 0.75 0.84 0.77 75.0 84.0 77.0

60 8.2 8.2 8.3 0.92 0.96 0.99 96.0 96.0 99.0

Page 123: LATEEF, FATAI ABIOLA

111

Infrared Spectra for Castor Seed Oil

111

Page 124: LATEEF, FATAI ABIOLA

112

Infrared Spectra for Palm Kernel Oil

112

Page 125: LATEEF, FATAI ABIOLA

113

Infrared Spectra for Groundnut Oil

113

Page 126: LATEEF, FATAI ABIOLA

114

APPENDIX IV

Table D1

Result of the Optimal Catalyst Weight Test for Maximum Biodiesel Yield at 1:3 Oil/Alcohol Ratio

The biodiesel yield obtained for the various catalyst weights are as shown below.

Volume of

oil (cm3)

Volume of C3H2OH

(cm3)

CSO, PKO,

GNO wt of

catalyst (g)

% (wt/v) of

catalyst

Vol. of Biodiesel % yield of Biodiesel

CSO PKO GNO CSO PKO GNO

50 15.00 for CSO &

PKO and 12.50 GNO

0.05 0.10 4.1 7.0 1.5 8.2 14.0 3.0

50 15.00 for CSO &

PKO and 12.50 GNO

0.10 0.20 6.0 18.0 19.5 12.0 36.0 39.0

50 15.00 for CSO &

PKO and 12.50 GNO

0.15 0.30 19.0 33.0 35.0 36.0 64.0 70.0

50 15.00 for CSO &

PKO and 12.50 GNO

0.20 0.40 30.0 38.0 39.6 60.0 76.0 79.0

50 15.00 for CSO &

PKO and 12.50 GNO

0.25 0.50 33.0 39.5 34.0 66.0 79.0 68.0

50 15.00 for CSO &

PKO and 12.50 GNO

0.30 0.60 12.5 37.5 38.5 65.0 75.0 57.0

50 15.00 for CSO &

PKO and 12.50 GNO

0.35 0.70 31.5 36.0 25.0 62.5 72.0 50.0

50 15.00 for CSO &

PKO and 12.50 GNO

0.40 0.80 29.5 25.0 15.0 59.0 50 30.0

114


Recommended