Download as pdf or txt
Download as pdf or txt
You are on page 1of 305

SYNTHESIS OF BIODIESEL FROM RUBBER SEED OIL FOR

INTERNAL COMPRESSION IGNITION ENGINE

Samuel Erhigare Onoji

A thesis submitted to the Faculty of Engineering and the Built

Environment, University of the Witwatersrand, Johannesburg, in

fulfillment of the requirements for the degree of Doctor of Philosophy.

Johannesburg, South Africa.

2nd December, 2017


DECLARATION

I declare that this thesis is my own, unaided work. It is being submitted for the Degree of Doctor

of Philosophy in the University of the Witwatersrand, Johannesburg. It has not been submitted

before for any degree or examination in any other University.

(Signature of candidate)

2nd.............Day of........December, ..........2017

i
ABSTRACT

Biodiesel has been identified as a good complement and plausible replacement of fossil diesel

because of the overwhelming characteristic properties similar to fossil diesel in addition to its

good lubricity, biodegradability, non-toxicity and eco-friendliness when used in diesel engines.

The production of biodiesel from edible vegetable oils competes with food sources, thereby

resulting in high cost of food and biodiesel. Studies have shown that rubber seed contains 35–

45 % oil, which portrays a better competitor to other non-edible oil bearing plants in biodiesel

production. In this study, non-edible vegetable oils from underutilized Nigerian NIG800 clonal

rubber seeds were extracted from 0.5 mm kernel particle size using n-hexane as solvent to

obtain a yield of 43 wt.% over an extraction time of 1 h. The oil was characterized for fatty

acids by using gas chromatography-mass spectrometry (GC-MS), and for structural properties

by Fourier transform-infrared (FT-IR) and nuclear magnetic resonance (NMR) analyses.

The optimization of the process conditions of the vegetable oil extraction was evaluated using

response surface methodology (RSM) and artificial neural network (ANN) techniques both of

which, were based on a statistically designed experimentation via the Box-Behnken design

(BBD). A three-level, three-factor BBD was employed using rubber seed powder (X1), volume

of n-hexane (X2) and extraction time (X3) as process variables. The RSM model predicted

optimal oil yield of 42.98 wt. % at conditions of X1 (60 g), X2 (250 mL) and X3 (45 min) and

experimentally validated as 42.64 wt. %. The ANN model predicted optimal oil yield of 43 wt.

% at conditions of X1 (40 g), X2 (202 mL) and X3 (49.99 min) and validated as 42.96 wt. %.

Both models were effective in describing the parametric effect of the considered operating

variables on the extraction of oil from the rubber seeds. On further examinations of the

potentials of the vegetable oil, the kinetics of thermo-oxidative degradation of the oil was

investigated. The kinetics produced a first-order reaction, with activation energy of 13.07

kJ/mol within the temperature range of 100–250 oC. In a bid to attain enhanced yield of

ii
biodiesel produced via heterogeneous catalysis, coupled with the carbonaceous potentials of

the pericarp and mesocarp of rubber seed shell casing as a suitable catalytic material, the rubber

seed shells (RSS) were used to develop a heterogeneous catalyst. RSS was washed 3–4 times

with hot distilled water, dried at 110 oC for 5 h, ground to powder, and calcined at 800 oC at a

heating rate of 10 oC/min as a catalyst and analyzed for thermal, structural, and textural

properties using thermogravimetric analyzer, x-ray diffractometer, and nitrogen

adsorption/desorption analyzer, respectively. The catalyst was further analyzed for elemental

compositions and surface morphology by x-ray fluorescence and scanning electron

microscopy, respectively. The catalyst was then applied in biodiesel production from rubber

seed oil. A central composite design (CCD) was employed together with RSM and ANN to

obtain optimal conditions of the process variables consisting of reaction time, methanol/oil

ratio, and catalyst loading on biodiesel yield. The optimum conditions obtained using RSM

were as follows: reaction time (60 min), methanol/oil ratio (0.20 vol/vol), and catalyst loading

(2.5 g) with biodiesel yield of 83.11% which was validated experimentally as 83.06±0.013%.

Whereas, those obtained via ANN were reaction time (56.7 min), methanol/oil ratio (0.21

vol/vol), and catalyst loading (2.2 g) with a biodiesel yield of 85.07%, which was validated

experimentally as 85.03±0.013%. The characterized biodiesel complied with ASTM D 6751

and EN 14214 biodiesel standards and was used in modern diesel test engine without technical

modifications. Though the produced biodiesel has a lower energy content compared with

conventional diesel fuel, in all the cases of blends considered, the optimal engine speed for

higher performance and lower emissions was observed at 2500 rpm. In this study, the B20

blend has best engine performance with a lower emission profile, and was closely followed by

B50 blend.

iii
DEDICATION

To my lovely wife Mrs. Augustina Onoji, and my children Faith, Emmanuel, Deborah, and

Mishael for their inestimable support, prayers and being my source of strength.

The work is also dedicated to my post-octogenarian Father, Mr. Iboyotete Onoji for his

confidence in me and encouragement throughout my years of study. I will not fail to dedicate

the work also to the memories of my mother, Mrs. Igogoma Onoji who has gone to be with the

Lord.

iv
ACKNOWLEDGMENTS

First and foremost, I appreciate God Almighty for the strength, knowledge and the grace He

gave me to start and complete the research successfully.

My deep appreciation goes to Professor Sunny E. Iyuke for his thorough supervision,

encouragement, guidance and valuable support he gave me while the project lasted. The three

years under his supervision motivated me to develop a scientific rigour and effectiveness.

My profound gratitude goes to my co-supervisors, Professor Michael O. Daramola of the

University of the Witwatersrand, Johannesburg, South Africa, and Professor Anselm I. Igbafe

of Afe Babalola University, Ado-Ekiti, Nigeria, for their close supervision, expertise and

valuable contributions to this study.

I acknowledged immensely the contributions of Dr. Julius I. Osayi of the Nanotechnology and

Petroleum Research Group of the University of the Witwatersrand for the role he played

throughout the duration of the research. Other members of the Nanotechnology and Petroleum

Research Group are also appreciated for their immense contributions during our weekly

research group meetings, and to Mr. Bruce Mothibeli for his laboratory assistance and materials

provided for the practical work.

I would like to acknowledge the contribution of Dr. Diakanua B. Nkazi and Family for the

hospitality showed to me during my stay in Johannesburg, and to Dr. Tunde F. Adepoju of

Landmark University, Omu-Aran, Nigeria, for his encouragement and laboratory assistance.

I want to warmly appreciate Dr. Kevin I. Idehen and Dr. Reuben M. Umoh, both of the

Petroleum Training Institute, Effurun, Nigeria for their valuable suggestions, insightful advice

and support.

v
I sincerely appreciate His lordship, Bishop Dr. & Rev. (Mrs.) Curtis Fianu, Bishop of New

Warri Bishopric Headquarters, Church of God Mission International Inc., Praise Centre,

Effurun, Nigeria for their encouragement and prayers, and to all Pastors, Ordained Ministers,

and Members of the Church for their support and goodwill.

Finally, I truly acknowledge the wonderful support of my beloved wife, Mrs. Augustina Onoji

and my lovely children, Faith, Emmanuel, Deborah, and Mishael for their prayers, patience and

sacrifice that strengthened me throughout the period of my study.

vi
LIST OF REFERRED PUBLICATIONS

Onoji, S.E., Iyuke, S.E., Igbafe, A.I., & Nkazi, D. B. (2016). Rubber seed oil: A potential

renewable source of biodiesel for sustainable development in sub-Saharan Africa. Energy

Conversion and Management, 110, 125–134.

Onoji, S.E., Iyuke, S.E., & Igbafe, A. I. (2016). Hevea brasiliensis (Rubber seed) oil:

Extraction, characterization, and kinetics of thermo-oxidative degradation using classical

chemical methods. Energy & Fuels, 30 (12), 10555–10567.

Onoji, S.E., Iyuke, S.E., Igbafe, A.I., & Daramola, M. O. (2017). Hevea brasiliensis (Rubber

seed) oil: Modeling and optimization of extraction process parameters using response surface

methodology and artificial neural network techniques. Biofuels, 1–15.

doi:10.1080/17597269.2017.1338122.

Onoji, S.E., Iyuke, S.E., Igbafe, A.I., & Daramola, M. O. (2017). Transesterification of rubber

seed oil to biodiesel over a calcined waste rubber seed shell catalyst: Modeling and

optimization of process variables. Energy & Fuels, 31 (6), 6109–6119.

vii
ORAL PRESENTATIONS AT CONFERENCES

Onoji, S.E., Iyuke, S.E., Igbafe, A.I., & Daramola, M. O. (2017, May). Synthesis of Rubber

Seed (Hevea brasiliensis) Oil Methyl Esters over a Calcined Waste Rubber Seed Shell Catalyst:

Modeling and Optimization of Process Variables. Paper presented at the 13th International

African Scholar Academic Conference on Delving into Issues and Resources for sub-Saharan

Africa Growth in the Millennium Era: A Multi-Disciplinary Approach. Universite de Sciences

et Technologies de Côte d'Ivoire.

viii
TABLE OF CONTENTS

Contents Page

DECLARATION...................................................................................................................i

ABSTRACT...........................................................................................................................ii

DEDICATION.......................................................................................................................iv

ACKNOWLEDGMENTS......................................................................................................v

LIST OF REFERRED PUBLICATIONS.............................................................................vii

PRESENTATIONS AT CONFERENCES..........................................................................viii

CONTENTS..........................................................................................................................ix

LIST OF FIGURES............................................................................................................xvii

LIST OF TABLES..............................................................................................................xxii

NOMENCLATURE..........................................................................................................xxiv

CHAPTER 1: INTRODUCTION...........................................................................................1

1.1 Research background and motivation................................................................................1

1.2 Hypothesis........................................................................................................................3

1.3 Justification of study.........................................................................................................3

1.4 Research aims and objectives............................................................................................5

1.5 Thesis outline....................................................................................................................7

CHAPTER 2: LITERATURE REVIEW................................................................................9

2.0 Introduction......................................................................................................................,9

2.1 The rubber tree..................................................................................................................9

2.2 Rubber seed capacity in sub-Saharan Africa...................................................................13

2.3 Rubber seed collection and processing............................................................................15

ix
2.4 Biomass valorization processes.......................................................................................15

2.5 Modeling and optimization approach to industrial research............................................18

2.6 Qualitative and quantitative analytical procedures..........................................................21

2.6.1 Thermogravimetric analysis.........................................................................................22

2.6.2 Fourier transform infrared spectroscopy......................................................................22

2.6.3 Scanning electron microscopy......................................................................................23

2.6.4 Brunauer- Emmett-Teller (BET) analysis....................................................................23

2.6.5 Nuclear magnetic resonance.........................................................................................24

2.6.6 Gas chromatography-mass spectroscopy.....................................................................24

2.6.7 X-ray diffraction and X-ray fluorescence.....................................................................25

2.7 Rubber seed analysis.......................................................................................................25

2.8 Rubber seed oil extraction methods.................................................................................29

2.8.1 Conventional extraction methods (mechanical and chemical) .....................................30

2.8.2 Gas-assisted mechanical extraction method.................................................................32

2.8.3 Microwave-assisted extraction (MAE) method............................................................32

2.8.4 Ultrasonic-assisted extraction method..........................................................................32

2.8.5 Aqueous enzymatic extraction method.........................................................................33

2.8.6 Supercritical fluid extraction (SFE) method.................................................................33

2.9 Physico-chemical properties of rubber seed oil...............................................................33

2.10 Thermo-oxidative analysis of rubber seed oil................................................................37

2.11 Industrial applications of rubber seed and oil................................................................39

2.12 Biofuels policies and sustainability in Africa................................................................40

2.12.1 Nigerian biofuel policy (A case study in Africa) ........................................................43

2.13 Catalysis........................................................................................................................45

2.13.1 Catalyst deactivation..................................................................................................47

x
2.13.2 Physical adsorption and chemisorption in chemical reaction catalysis......................47

2.14 Gas-solid adsorption systems........................................................................................49

2.14.1 Langmuir model equation (linearized).......................................................................51

2.14.2 BET model equation...................................................................................................51

2.14.3 Freundlich model equation.........................................................................................52

2.14.4 The t-plot....................................................................................................................53

2.14.4.1 Harkins-Jura model.................................................................................................54

2.14.4.2 Halsey equation.......................................................................................................55

2.14.5 Polanyi's potential theory...........................................................................................55

2.14.6 Dubinin-Radushkevich equation................................................................................56

2.14.7 Kelvin equation..........................................................................................................57

2.14.8 Thermodynamics of gas adsorption............................................................................58

2.15 Biodiesel and its origin..................................................................................................59

2.15.1 Biodiesel feedstock....................................................................................................63

2.16 Catalysis in biodiesel production...................................................................................64

2.16.1 Acid-catalyzed esterification reaction........................................................................64

2.16.2 Transesterification of oils using homogeneous catalyst............................................65

2.16.2.1 Alkali-catalyzed transesterification reaction...........................................................66

2.16.2.2 Two-step transesterification reaction......................................................................67

2.16.2.3 Mechanisms for homogeneous catalysis.................................................................67

2.16.3 Transesterification of oils using heterogeneous catalysts..........................................67

2.16.3.1 Heterogeneous acid transesterification reaction......................................................68

2.16.3.2 Heterogeneous base transesterification reaction......................................................69

2.16.4 Enzyme-catalyzed transesterification reaction..........................................................71

2.16.5 Heterogeneous catalyst from natural sources............................................................71

xi
2.17 Transesterification of rubber seed oil............................................................................72

2.18 Transesterification methods for rubber seed oil.............................................................73

2.18.1 The conventional method...........................................................................................73

2.18.2 The in-situ method......................................................................................................73

2.18.3 The enzymatic method...............................................................................................74

2.18.4 The supercritical method............................................................................................74

2.18.5 The co-solvent assisted method..................................................................................74

2.18.6 The ultrasonic-assisted method..................................................................................74

2.18.7 The microwave-assisted method................................................................................75

2.19 Factors affecting the transesterification of rubber seed oil.............................................75

2.19.1 FFA content of rubber seed oil....................................................................................75

2.19.2 Alcohol/oil molar ratio...............................................................................................76

2.19.3 Catalyst concentration................................................................................................76

2.19.4 Reaction temperature and time...................................................................................77

2.19.5 Agitation speed..........................................................................................................77

2.20 Thermodynamics of internal combustion ignition engines............................................78

2.20.1 Overview of reciprocating engines.............................................................................78

2.20.2 Air-standard assumptions...........................................................................................80

2.20.3 Diesel cycle-The ideal cycle for compression ignition engines..................................81

2.20.4 Mathematical models for internal compression ignition engines...............................84

2.20.5 Models for gas mixture properties..............................................................................87

2.20.6 First and second laws of thermodynamics for internal combustion engines..............88

2.20.7 Second law of thermodynamics analysis (availability).............................................91

2.21 Performance evaluation of diesel engines....................................................................95

2.21.1 Brake mean effective pressure...................................................................................96

xii
2.21.2 Brake power...............................................................................................................96

2.21.3 Brake specific fuel consumption................................................................................97

2.21.4 Brake thermal efficiency............................................................................................97

2.22 Performance and emissions evaluation of diesel engine using diesel-biodiesel blends of

rubber seed oil.......................................................................................................................97

CHAPTER 3: METHODOLOGY AND EXPERIMENTAL PROCEDURES...................100

3.0 Introduction……………………………………………...............................................100

3.1 Materials and chemicals................................................................................................100

3.2 Experimental procedures for seed oil extraction………………………………………101

3.2.1 Rubber seed preparation……….……………………………………………………101

3.2.2 Extraction of the seed oil……………………………………………………………102

3.3 Characterization of the seed oil……………………………………………………….103

3.4 Spectroscopy analyses of the seed oil…………...…………………………………….104

3.4.1 Fatty acid profile of the seed oil………….….............................................................104

3.4.2 Fourier transform infrared spectroscopy....................................................................105

3.4.3 Nuclear magnetic resonance spectroscopy...………….…………………………….105

3.5 Kinetics of thermal oxidative degradation of the seed oil..............................................106

3.6 Optimization study of the oil extraction process using RSM and ANN techniques.......107

3.6.1 Modeling of seed oil extraction process variables via RSM……...............................108

3.6.2 Modeling of seed oil extraction process variables via ANN………….......................111

3.7 Catalyst preparation and characterization......................................................................113

3.8 Biodiesel production from the extracted seed oil and analysis………….......................115

3.8.1 Acid-catalyzed esterification of rubber seed oil..........................................................115

3.8.2 Biodiesel production from the esterified seed oil........................................................116

xiii
3.9 Characterization of the produced biodiesel…………...................................................118

3.10 Reusability test for the catalyst used in biodiesel production…...................................119

3.11 Statistical analysis of the biodiesel production by RSM..............................................119

3.12 Analysis of variance (ANOVA) study for biodiesel production..................................120

3.13 Modeling of biodiesel production process variables……............................................120

3.14 Data validation for RSM and ANN models.................................................................121

3.15 Combustion of biodiesel-diesel blends in internal compression ignition engine.........121

3.15.1 Biodiesel-diesel blending analysis via ultrasonication technique.............................122

3.15.2 Methodology and experimental setup for combustion study of biodiesel blends......123

CHAPTER 4: RESULTS AND DISCUSSION.................................................................129

4.0 Introduction……………………………………….......................................................129

4.1 Extraction of seed oil and determination of the best seed particle size...........................129

4.2 Physical characterization of the seed and seed oil……………......................................130

4.3 Characterization of the seed oil…….............................................................................131

4.3.1Colour, pH, density and Specific gravity.....................................................................132

4.3.2 Iodine, peroxide, saponification and acid values........................................................132

4.3.3 Kinematic viscosity....................................................................................................133

4.3.4 Cold flow properties...................................................................................................134

4.3.4.1 Pour point................................................................................................................134

4.3.4.2 Cloud point..............................................................................................................134

4.3.4.3 Cold filter plugging point........................................................................................135

4.3.5 Aniline point..............................................................................................................135

4.3.6 Boiling, freezing, flash and fire points........................................................................135

4.3.7 Refractive index.........................................................................................................136

xiv
4.3.8 Other parameters........................................................................................................136

4.3.8.1 Cetane number........................................................................................................136

4.3.8.2 Diesel index.............................................................................................................137

4.3.8.3 Higher heating value...............................................................................................137

4.3.8.4 Average molecular weight of rubber seed oil..........................................................137

4.3.9 Fatty acid profile of the seed oil..................................................................................140

4.3.10 FT-IR and NMR spectroscopy of the seed oil...........................................................142

4.3.10.1 FT-IR analysis of the seed oil................................................................................142

4.3.10.2 1H NMR analysis of the seed oil…………………................................................145

4.3.10.3 13C NMR analysis of the seed oil...........................................................................147

4.4 Kinetic study of thermo-oxidative degradation of rubber seed oil.................................148

4.5 Parametric effects and optimization of oil extraction using RSM………......................154

4.5.1 Full quadratic regression model.................................................................................154

4.5.2 Optimization of oil extraction process variables based on RSM.................................157

4.5.3 Analysis of variance (ANOVA) study for seed oil extraction.....................................158

4.5.4 Parametric effects of interactive factors on the seed oil yield...................................161

4.6 Parametric effects and optimization of oil extraction using ANN model……............164

4.7 Comparison of RSM and ANN models for the seed oil extraction..............................169

4.8 Catalyst preparation and characterization………………………….............................170

4.8.1 Thermo-gravimetric analysis of the rubber seed shell catalyst….............................170

4.8.2 XRF elemental analysis and basic property of the rubber seed shell catalyst….......171

4.8.3 XRD analysis of the rubber seed shell catalyst...........................................................172

4.8.4 SEM analysis of the rubber seed shell catalyst catalyst...............................................174

4.8.5 Surface properties of the rubber seed shell catalyst....................................................175

4.9 Biodiesel production from the extracted seed o………………………….....................180

xv
4.9.1 Characterization of the rubber seed oil biodiesel......................................................180

4.9.2 Reusability of the catalyst for biodiesel production..................................................182

4.10 Parametric effect and optimization of biodiesel production process variables..........183

4.10.1 Modeling and process variables optimization by RSM…………….......................183

4.10.2 Analysis of variance (ANOVA) for biodiesel production………...........................188

4.10.3 Parametric effects of interactive factors on biodiesel yield using RSM…................191

4.10.4 Optimization and interactive effects of process variables on biodiesel yield by ANN

model..................................................................................................................................193

4.10.5 Comparison of RSM and ANN models for the biodiesel production......................197

4.11 Performance and emissions analysis of test engine using biodiesel-diesel blends.......198

4.11.1 Air/Fuel ratio............................................................................................................198

4.11.2 Brake mean effective pressure..................................................................................199

4.11.3 Brake thermal efficiency..........................................................................................200

4.11.4 Brake specific fuel consumption..............................................................................201

4.11.5 Engine brake power..................................................................................................202

4.11.6 Engine torque...........................................................................................................203

4.11.7 Exhaust gas temperature...........................................................................................204

4.11.8 CO emissions...........................................................................................................205

4.11.9 CO2 emissions..........................................................................................................206

4.11.10 NOx emissions........................................................................................................207

4.11.11 Total hydrocarbons.................................................................................................208

4.11.12 Smoke opacity........................................................................................................209

CHAPTER 5: CONCLUSIONS AND RECOMMENDATIONS.......................................211

5.1 Conclusion....................................................................................................................211

xvi
5.2 Recommendations.........................................................................................................212

5.3 Contributions to knowledge..........................................................................................213

REFERENCES...................................................................................................................214

APPENDIX.........................................................................................................................240

Appendix A.........................................................................................................................240

Appendix B.........................................................................................................................244

Appendix C.........................................................................................................................246

Appendix D.........................................................................................................................248

Appendix E.........................................................................................................................251

Appendix F..........................................................................................................................259

Appendix G.........................................................................................................................262

Appendix H.........................................................................................................................269

Appendix I…………………………………………..........................................................272

LIST OF FIGURES

Figure Page

1.1 Systematic approach employed in this research.................................................................6

2.1 Rubber trees plantation....................................................................................................11

2.2 Rubber seed.....................................................................................................................12

2.3 Top ten leading producers of rubber seeds in 2014..........................................................12

2.4 Overview of main biomass valorizations.........................................................................18

2.5 Multilayer structure of an ANN model............................................................................20

2.6 General flowchart of Genetic Algorithm.........................................................................21

xvii
2.7 Flowchart of rubber seed oil processing..........................................................................30

2.8 Representation of the action of a catalyst.........................................................................47

2.9 Adsorption isotherm classified by Brunauer...................................................................50

2.10 A typical structure of triglyceride molecule...................................................................63

2.11 Scheme for stepwise transesterification of triglyceride to methyl esters........................66

2.12 Schematic diagram for a reciprocating engine cylinder.................................................79

2.13 P–v (a) and T–s (b) diagrams for the ideal Diesel cycle................................................82

3.1 Flowchart of rubber seed powder preparation………………………………………..102

3.2 Schematic of 3-3-1 developed ANN topology..............................................................111

3.3 Levenberg-Marquardt training flowchart for ANN model............................................113

3.4 Experimental setup for biodiesel-diesel blending analysis...........................................123

3.5 Experimental TD202 test engine for performance and combustion analyses located at

Petroleum Training Institute, Effurun, Nigeria....................................................................124

3.6 A schematic layout of TD202 diesel test engine............................................................128

4.1 Effects of kernel particle size on the rubber seed oil yield..............................................130

4.2 Gas chromatogram of the rubber seed oil with identified peaks.....................................141

4.3 FT-IR spectrum of the rubber seed oil...........................................................................143

4.4 1H NMR spectrum of the rubber seed oil.......................................................................147

4.5 13C NMR spectrum of the rubber seed oil......................................................................148

4.6 Overall process of oil degradation.................................................................................150

4.7 Acyclic polymer formations from oleic acid during thermal oxidation of oils...............151

4.8 ln (Ct/Co) vs. time (min) for iodine values at 100, 150, 200 and 250 oC........................152

4.9 Arrhenius plot for kinetics of thermo-oxidative degradation of rubber seed oil..........153

4.10 Correlation of predicted and experimental oil yields of rubber seed oil extraction......155

xviii
4.11 (a) Interactive effects between solvent volume and rubber seed powder weight on oil

yield for RSM response surface (3D) plot...........................................................................162

4.11 (b) Interactive effects between extraction time and rubber seed powder weight on oil

yield for RSM response surface (3D) plot...........................................................................163

4.11 (c) Interactive effects between extraction time and solvent volume on oil yield for RSM

response surface (3D) plot...................................................................................................164

4.12 Correlation of predicted yield and experimental yield of RSO by ANN model...........165

4.13 (a) Interactive effects of rubber seed powder weight and solvent volume on oil yield for

ANN response surface (3D) and contour (2D) plots............................................................166

4.13 (b) Interactive effects of rubber seed powder weight and extraction time on oil yield for

ANN response surface (3D) and contour (2D) plots............................................................167

4.13 (c) Interactive effects of extraction time and solvent volume on oil yield for ANN

response surface (3D) and contour (2D) plots.....................................................................168

4.14 Percentage contribution of process variables to seed oil extraction via ANN model...169

4.15 TGA profile of the rubber seed shell (RSS) catalyst....................................................170

4.16 XRF analysis of the RSS: (a) raw and (b) calcined at 800 oC.......................................172

4.17 XRD diffractograms of the RSS: (a) raw, and (b) calcined at 800 oC...........................174

4.18 SEM images for (a) raw, and (b) calcined RSS catalyst at magnification 2450x..........175

4.19 Nitrogen adsorption/desorption isotherms for the raw and calcined RSS....................176

4.20 Effects of reuse of different catalyst loadings (2, 2.2, 2.5, 3, and 3.5 g) on biodiesel yield

(%) at optimum conditions..................................................................................................182

4.21 Effects of Ca2+ ion leaching from reused catalyst on biodiesel yield (%) at optimum

conditions............................................................................................................................183

4.22 Correlation of predicted vs. actual biodiesel yields (%) for the RSM model................186

xix
4.23 (a) RSM 3D plot for the interactive effect between reaction time and methanol/oil ratio

on biodiesel (RSOME) yield...............................................................................................191

4.23 (b) RSM 3D plot for the interactive effect between reaction time and catalyst amount

on biodiesel (RSOME) yield...............................................................................................192

4.23 (c) RSM 3D plot for the interactive effect between methanol/oil ratio and catalyst

amount on biodiesel (RSOME) yield..................................................................................193

4.24 Correlation of predicted vs. actual biodiesel yields for ANN model............................194

4.25 ANN Surface plot for the interactive effect of methanol/oil ratio with reaction time on

biodiesel (RSOME) yield....................................................................................................195

4.26 ANN Surface plot for the interactive effect of catalyst amount with reaction time on

biodiesel (RSOME) yield....................................................................................................196

4.27 ANN Surface plot for the interactive effect of methanol/oil ratio with catalyst amount

on biodiesel (RSOME) yield...............................................................................................196

4.28 Percentage contribution of process variables to biodiesel production via ANN

model..................................................................................................................................197

4.29 Variation of air/fuel ratio with speed for diesel-biodiesel blends.................................199

4.30 Variation of brake thermal efficiency with speed for diesel-biodiesel blends..............200

4.31 Variation of brake specific fuel consumption with speed for diesel-biodiesel

blends..................................................................................................................................201

4.32 Variation of engine power with speed for diesel-biodiesel blends...............................203

4.33 Variation of engine torque with speed for diesel-biodiesel blends...............................204

4.34 Variation of exhaust gas temperature with speed for diesel-biodiesel blends..............205

4.35 Variation of carbon monoxide emissions with speed for diesel-biodiesel blends........206

4.36 Variation of carbon dioxide emissions with speed for diesel-biodiesel blends............207

4.37 Variation of nitrogen oxides emissions with speed for diesel-biodiesel blends...........208

xx
4.38 Variation of total hydrocarbons emissions with speed for diesel-biodiesel blends......209

4.39 Variation of smoke opacity with speed for diesel-biodiesel blends.............................210

C.1 Preparation of milled kernel for oil extraction ..............................................................246

C.2 Oil soxhlet extractor used for the study.........................................................................246

C.3 Rotary evaporator used in the study..............................................................................247

C.4 Extracted rubber seed oil: (a) filtered, (b) crude............................................................247

E.1 Calcination of rubber seed shell (RSS) catalyst for biodiesel production.....................259

F.1 Esterified oil separation into two layers.........................................................................260

F.2 Gravity settling of biodiesel, catalyst and glycerol phases….........................................260

F.3 Processed rubber seed oil methyl ester (biodiesel) for analysis.....................................261

F.4 Regression line (−), upper (−), and lower (−) prediction intervals and experimental data

(∎) of oxidation stability determination for biodiesel without additive...............................261

G.1 13C NMR spectra of rubber seed oil (12–28 ppm) ........................................................265

G.2 13C NMR spectra of rubber seed oil (28.4–34.8 ppm) ..................................................266

G.3 13C NMR spectra of rubber seed oil (40–84 ppm) ........................................................266

G.4 13C NMR spectra of rubber seed oil (126.4–132.0 ppm) ..............................................267

G.5 13C NMR spectra of rubber seed oil (165–184 ppm) ....................................................267
13
G.6 Double bond peaks in the C NMR spectrum for pure oleic acid, linoleic acid, and

linolenic acid.......................................................................................................................268

H.1 Nitrogen adsorption/desorption log plot for calcined RSS............................................270

H.2 Nitrogen adsorption/desorption log plot for raw RSS...................................................270

H.3 BJH adsorption pore-size distribution log plot for calcined RSS..................................271

xxi
LIST OF TABLES

Table Page

2.1 Estimated annual rubber seed oil and biodiesel production capacity for sub-Saharan

Africa (SSA) countries in 2013.............................................................................................14

2.2 Typical levels of cellulose, hemicellulose, and lignin in biomass....................................17

2.3 Proximate analysis of rubber seed....................................................................................27

2.4 Amino acid profile of rubber seed...................................................................................28

2.5 Proximate analysis of RSS and RSK...............................................................................28

2.6 Ultimate analysis of RSS and RSK..................................................................................28

2.7 Extractives, holocellulose, hemicelluloses and cellulose contents of RSS and RSK.......29

2.8 Calorific value and oil yield of RSS and RSK..................................................................29

2.9 Physico-chemical properties of rubber seed oil................................................................35

2.10 Fatty acids profile of rubber seed oil..............................................................................36

2.11 Fatty acid compositions of selected non-edible oils (%)................................................37

2.12 Africa countries which import and export energy..........................................................42

2.13 SSA countries and their bioenergy policies....................................................................43

2.14: Comparison between physisorption and chemisorption processes................................49

2.15 Methods of viscosity reduction in vegetable oils for biodiesel production....................62

3.1 Experimental range and levels of independent process variables for seed oil extraction

via Box-Behnken design.....................................................................................................108

3.2 Experimental range and levels of independent process variables for biodiesel production

via central composite design...............................................................................................117

3.3 Technical specifications for TD 202 test diesel engine..................................................124

4.1 Physical characterization of rubber seed and the seed oil...............................................131

4.2 Physico-chemical properties and other characteristics of rubber seed oil......................139

xxii
4.3 Fatty acids profile of rubber seed oil..............................................................................140

4.4 Wavenumbers and their functional groups for FT-IR analysis of rubber seed oil..........144

4.5 1H NMR spectrum analysis of rubber seed oil...............................................................146

4.6 13C NMR analysis of carbon type present in the rubber seed oil.....................................148

4.7 Effect of heating on peroxide, iodine, and refractive index values of rubber seed oil.....149

4.8 Kinetic parameters for degradation of rubber seed oil...................................................153

4.9 Box–Behnken design for the rubber seed oil extraction process optimization by RSM and

ANN....................................................................................................................................156

4.10 Analysis of variance (ANOVA) for response surface quadratic model.......................159

4.11 RSM regression analysis for rubber seed oil extraction...............................................161

4.12 Summary report of N2 adsorption/desorption analysis for raw RSS............................178

4.13 Summary report of N2 adsorption/desorption analysis for calcined RSS (800 oC) ...179

4.14 Fuel properties of biodiesel from rubber seed oil.........................................................181

4.15 Central composite design for optimization of biodiesel production process by RSM and

ANN....................................................................................................................................185

4.16 RSM regression coefficients analysis for the oil biodiesel production.........................187

4.17 Analysis of variance (ANOVA) results for the fitted second-order polynomial

model..................................................................................................................................190

4.18 Properties of the test fuels............................................................................................198

D.1 Variation of oil yield with seed kernel particle size.......................................................249

E.1 Variation of peroxide, iodine, and refractive index values of rubber seed oil with

temperature and time...........................................................................................................258

F.1 Esterification parameters of rubber seed oil..................................................................259

G.1 Infrared group absorption frequencies..........................................................................262

G.2 Assignment of signals of 1H NMR spectra for vegetable oils (hazel nut and walnut)....265

xxiii
H.1 XRF spectroscopic elemental analysis of rubber seed shells catalyst...........................269

H.2 Biodiesel yields for fresh and reused catalyst loadings at optimum conditions.............271

I.1 Variation of engine speed with fuel and air parameters for B00 (diesel).........................273

I.2 Variation of engine speed with fuel and air parameters for B10.....................................273

I.3 Variation of engine speed with fuel and air parameters for B20.....................................274

I.4 Variation of engine speed with fuel and air parameters for B30.....................................274

I.5 Variation of engine speed with fuel and air parameters for B50.....................................275

I.6 Variation of engine speed with fuel and air parameters for B100 (biodiesel).................275

I.7 Variation of brake thermal efficiency with engine speed................................................276

I.8 Variation of brake specific fuel consumption with engine speed....................................276

I.9 Variation of engine brake power with engine speed.......................................................276

I.10 Variation of engine torque with engine speed...............................................................276

I.11 Variation of exhaust gas temperature with engine speed..............................................277

I.12 Variation of carbon monoxide emissions with engine speed........................................277

I.13 Variation of carbon dioxide emissions with engine speed............................................277

I.14 Variation of oxides of nitrogen emissions with engine speed.......................................277

I.15 Variation of total hydrocarbons emissions with engine speed......................................278

I.16 Variation of smoke opacity emissions with engine speed.............................................278

NOMENCLATURE

Description Symbol

Adsorbate molecular area Am

Adsorbed volume Va

Adsorptive energy En

Adsorption enthalpy change ∆Hads

xxiv
Adsorption entropy change ∆Sads

Air-box differential pressure ∆P

Avogadro's number N

Brake thermal efficiency ηth

Brunauer-Emmett-Teller constant C

Carbon dioxide, Carbon monoxide CO2, CO

Clearance volume Vc

Coefficient of discharge Cd

Compression ratio r

Concentration Ct, Co

Conference of the Parties COP

Connecting rod l

Crank angle θ

Crank radius α

Critical pressure, Critical temperature Pc, Tc

Cutoff ratio rc

Engine speed N

Exergy ai

Fuel f

Hectares, Hour ha, h

Internal energy, Velocity U

Irreversibility I

Langmuir constant b

Langmuir specific surface area SL

Mass m

xxv
Mass flowrate of air 𝒎̇𝒂

Minutes min

Molecular mass M

Monolayer volume Vm

Number of moles, variables, and order of reaction n

Orifice diameter d

Oxides of nitrogen NOx

Pressure P

Process actual variable X

Process coded variable x

Quantity of heat Q

Rate constant k

Saturation pressure Po

Specific heat at constant pressure Cp

Specific heat at constant volume Cv

Specific heat ratio (Cp/Cv) γ

Specific volume v

Standard deviation of data σ

Statistical thickness t

Temperature, Torque T

Time t

Universal gas constant R

Volume of biodiesel blends (%) Bxx

Work W

xxvi
CHAPTER 1: INTRODUCTION

1.1 Research background and motivation

Industrial advancement in recent times has accelerated the need for sustained energy

requirement across the globe. The depletion of non-renewable fossil materials (oil, gas, and

coal) that arises from their continued usage, their price instability, energy insecurity, and the

concern for the degraded environment stimulated global research for alternative sources of

energy (Aldhaidhawi et al., 2017; Onoji et al., 2016a). Renewable energies from biomass,

hydropower, solar, tidal waves, wind and geothermal have been proposed and considered as

cheap sources of sustainable energies (Dwivedi et al., 2013; Thanh et al., 2012). The biomass

source option, which appeared the cheapest or easiest to acquire will most certainly result to

fast economic and sustainable growth of most developing countries in Africa (Onoji et al.,

2016a; Thanh et al., 2012). Vegetable oils from biomass are seen as source materials for

renewable energy production. The suitability of vegetable oils as source of biodiesel production

has been reported in the literature by several researchers in recent times. Biodiesel is a mono-

alkyl fatty acid methyl ester (FAME) produced by the transesterification of vegetable oils,

animal fats and other sources of oils/fats with primary alcohols (methanol, in this study) in the

presence of suitable catalysts (homogeneous, heterogeneous, and enzymes, etc) (Roschat et al.,

2012; Guo et al., 2010).

The choice of biodiesel is based on its biodegradability, non-toxic, lower emissions, sulphur-

free, low levels of polycyclic aromatic hydrocarbons (PAHs) and n-PAHs (Mofijur et al., 2014;

Portet-Koltalo & Machour, 2013; Jaya & Ethirajulu, 2011; Kumar & Purushothaman, 2010).

In recent times, most of the biodiesel production routes are from edible oil plants such as palm

oil, soybean, rapeseed, sunflower, and these compete with food/pharmaceutical and cosmetic

uses, resulting to food-fuel crisis and high price of biodiesel generated thereof (Onoji et al.,

1
2016a; Ahmad et al., 2013; Asuquo et al., 2012; Ma & Hanna, 2012). To avert a food-fuel

crisis, non-edible vegetable oils considered as low-grade oils that are not suitable for food uses,

could be an attractive preference to edible oils for industrial applications (Dharma et al., 2016).

One of the few versatile bioenergy crops with non-edible oil that has attracted the attention of

researchers in recent times is the rubber tree (Hevea brasiliensis) (Ng et al., 2013).

Hevea brasiliensis is economically cultivated for the production of latex as a source of natural

rubber for the production of various rubber products in use globally, while the seeds are

underutilized (Takase et al., 2015; Atabani et al., 2013; Ng et al., 2013). However, the oil from

the seed is the second most valuable product after the latex (Banković-Ilić et al., 2012). With

the concept of sustainable development and the Paris Agreement on climate change (UNFCCC,

2015), the production of synthetic rubber would be reduced when the use of fossil materials is

phased out in the dawn of next century. The cultivation of rubber tree to generate latex for

natural rubber production, and the utilization of its seeds to produce non-edible oil for biodiesel

production will boost the economy of most Africa countries that are poverty-ridden.

Homogeneous base catalysts such as NaOH, KOH and methoxides of sodium and potassium

are most suitable for industrial transesterification of edible vegetable oils that are generally low

in free fatty acids (FFAs). However, associated problems such as difficulty in catalyst-products

separation, initial catalyst cost, soap formation, low-grade by-product glycerol, reactor

corrosion and large volume of washing water requirements induced further research for

environmental benign heterogeneous catalyst for biodiesel synthesis from non-edible oils

though high in FFAs (Roschat et al., 2012; Zabeti et al., 2009). In addition, the use of

homogeneous catalysts in transesterification of non-edible oils to biodiesel is hampered by the

high FFAs of the oils that react with the catalyst resulting to low biodiesel yield. Uprety et al.

2
(2016) reported that heterogeneous catalysts are easily separated from reaction mixtures

through simple filtration, and reused severally producing high-grade glycerol. Against this

background, this study investigated the conversion of non-edible rubber seed oil to biodiesel

over a catalyst derived from waste rubber seed shells (RSS). The results obtained from the

study are promising and it is expected that this will further reduce the cost of biodiesel produced

from rubber seed oil and make it more competitive to petro-diesel.

1.2 Hypothesis

 Industrial processes go faster to completion when homogeneous, heterogeneous, or

enzymes catalysts are employed;

 Homogeneous catalysts are not suitable for the transesterification of non-edible rubber

seed oil with high FFAs;

 Heterogeneous catalyst proposed in this research will attempt to solve the problems and

will provide a better product output when employed for the transesterification reaction.

 When reactions take place in solid-liquid-vapour phase, mass transfer limitations are

avoided, and the reaction may be enhanced if solid surface area becomes large;

 The structural and textural properties of the catalyst may offer insight to anticipated

behaviour.

1.3 Justification of the study

During the Climate Change Summit (COP 21) held in Paris, from November 30 to December

12, 2015 to address the threat posed to humanity from climate change, world leaders, Policy

makers, and scientists agreed that man is the major cause of climate change that results to

increased global warming. A universal climate change agreement termed 'Paris Agreement'

was reached and agreed that clean renewable energy sources would replace fossil materials for

3
energy generation in the dawn of next century (UNFCCC, 2015). The agreement will cause a

reduction in the production of synthetic rubber from petrochemicals derived from fossil

materials. The rubber tree (Hevea brasiliensis) from which natural rubber is sourced will play

a significant role in the future energy mix. The underutilized rubber seeds will provide non-

edible oil for biodiesel production to avert food-fuel crisis and enhance poverty-alleviation

programme especially in the third world countries. The rubber seed shells (RSS) generated

during oil extraction posed waste disposal problems to rubber seed oil millers. Synthesis of

solid catalyst from waste RSS for use in biodiesel production in this study will further reduce

the cost of biodiesel and make it more affordable.

Pursuant to government directives in 2005 for the diversification of Nigeria’s economy through

the mainstreaming of an Automotive Biomass Programme, the Nigerian National Petroleum

Corporation (NNPC) was mandated to create a conducive environment for the take-off of a

domestic bioenergy industry. The underlying objectives for government interest in a national

biofuel programme are revenue diversification, reduced dependence on imported refined

petroleum-derived products, improved agricultural productivity, job creation, meeting energy

needs, and environmental benefits (Ishola et al., 2013; Abila, 2012). The Nigeria biofuel policy

did not address the potential food-fuel conflicts that could arise from the use of food crops as

biofuel feedstock (Ohimain, 2013). In Nigeria, about 18 million hectares (ha) of land are

suitable for industrial cultivation of rubber trees in the southern region (Onoji et al., 2016a).

The complete life cycle and environmental benefits of rubber trees in Nigeria is well

investigated and defined compared to Jatropha curcas, a similar non-edible plant (Ishola et al.,

2013), but detailed study on the benefits of rubber tree as a versatile bioenergy crop is limited

as it was not mentioned as a plausible feedstock in the Nigeria's biofuels policy.

The outcomes of this research, as documented in this thesis, will encourage the continued

cultivation of rubber trees for latex production, and harvesting the seed to produce oil for

4
biodiesel synthesis. It will also serve as a pool of information for the relevant Government

Ministries, Departments, and Agencies that will serve as advocates for the implementation of

the National Biofuels Policy across Nigeria's business strata with respect to biodiesel

production from rubber seed oil and other African nations where rubber trees are cultivated.

1.4 Research aims and objectives

The main objectives of this research study were as follows:

 Extract, characterize, and determine the kinetics of thermo-oxidative degradation of

rubber seed oil;

 Develop a mathematical model, and optimize the extraction of rubber seed oil using the

developed model;

 Synthesize and characterize a solid catalyst from waste rubber seed shells for

transesterification of rubber seed oil to biodiesel;

 Develop a mathematical model, and optimize the biodiesel production from rubber seed

Oil using the developed model;

 Characterize the produced biodiesel and compare the properties with biodiesel

Standards (ASTM D 6751 and EN 14214);

 Study the internal combustion of diesel-biodiesel blends using a test engine;

 Analyse the level of pollutants and emissions arising from the use of diesel, biodiesel,

and diesel-biodiesel blends to ascertain their levels of environmental friendliness.

Figure 1.1 depicts the systematic approach employed in this research. The approach involves

catalyst synthesis from waste rubber seed shells (RSS) calcined at 800 oC based on thermo-

gravimetric analysis. The catalyst was characterized for structural and textural properties.

Rubber seed oil (RSO) extracted from rubber seeds was esterified with methanol using 6% v/v

sulphuric acid as catalyst. The esterified oil was then transesterified to biodiesel over a calcined

5
RSS as base catalyst. The biodiesel produced was blended with petro-diesel, and the blends

were analysed in TD202 internal combustion diesel test engine for performance and emissions

evaluation.

rubber seeds

oil extraction
process

rubber seed oil


(RSO)
esterification
6% v/v H2SO4 process
waste water
rubber seed
shells (RSS)
methanol esterified RSO

catalyst transesterification
synthesis RSS process high grade
furnace Cat. glycerol
operated
@ 800oC
(10 oC/min)

methanol biodiesel

internal
blending process exhaust gases
blend combustion
diesel s
TD 202 (for analysis)
diesel
test engine

Electrical energy

Figure 1.1: Systematic approach employed in this study

6
1.5 Thesis outline

The layout of the thesis is as follows:

This thesis investigated the conversion of non-edible and high free fatty acids rubber seed oil

to biodiesel over a catalyst synthesized from waste rubber seed shells. The following chapters

discuss the issues related to the synthesis and characterization of the catalyst, biodiesel

production and evaluation of the performance and emission profiles of diesel-biodiesel blends

using a diesel test engine.

Chapter 1 is the introductory chapter of the thesis and it provides information on the research

background and motivation, hypothesis, and justification of the study. The chapter further

discusses the aims and objectives, conceptual approach used to achieve the main goal of the

research, and finally the thesis outline.

Chapter 2 presents the literature review and background information on rubber tree (Hevea

brasiliensis). The chapter further highlights the benefits of biodiesel policies in sub-Saharan

Africa (SSA); rubber seed capacity in SSA; industrial applications of rubber seeds; modeling

and optimization in research. Others include oil extraction methods; kinetics of oil degradation;

transesterification methods; catalysis; optimization of oil extraction and biodiesel production;

and finally, the thermodynamics of the internal combustion of diesel-biodiesel blends for the

evaluation of engine performance and emission analysis.

Chapter 3 contains details of the research methodology, experimental methods, and analytical

procedures used in this research. The chapter presents materials and seed preparation; seed oil

extraction; characterization of seed oil; oil degradation kinetics and optimization of oil

extraction process. It also includes catalyst synthesis and characterization; biodiesel production

and characterization; modeling and optimization of biodiesel production process; and engine

performance and emission analysis of biodiesel-diesel blends.

7
Chapter 4 presents the experimental results and discussion on oil extraction, catalyst synthesis,

biodiesel production, and their analyses. The chapter also discusses the performance of the

blends in the internal compression ignition engine, and emission analysis. Results of previous

researchers were compared with the findings of this research.

Chapter 5 concludes this research by providing a summary of the main findings,

recommendations for future work, and contributions to knowledge.

8
CHAPTER 2: LITERATURE REVIEW

2.0 Introduction

The literature review presented in this chapter is a part of a review article, and experimental

work on this research published in Energy Conversion and Management, 2016 (issue 110, pp.

125-134), and Energy & Fuels, 2016 (issue 30, number 12, pp. 10555-10567). The cover and

abstract pages of the published papers are provided in Appendix A.

2.1 The rubber tree

The rubber tree (Hevea brasiliensis) shown in Figure 2.1, also known as the Para rubber

belongs to the plant family Euphorbiaceae and it is the most economically important member

of the genus Hevea that originated from Amazon region of Brazil in South America (Onoji et

al., 2016a; Reshad et al., 2015; Kittigowittana et al., 2013). Hevea exhibits morphological

variability, with ten species (H. brasiliensis, H. benthamiana, H. camorun, H. guianensis, H.

microphylla, H. nitida, H. pauciflora, H. rigidifolia, H. camargoana, and H. spruceana)

identified that range from large forest trees to shrubs with latex in their parts (Venkatachalam

et al., 2013). Hevea brasiliensis produces about 99% of the world's natural rubber (Aravind et

al., 2015; Atabani et al., 2013). It was commercialized for latex production to boost the

economy of Brazil in the 1800s. Several researchers investigated the properties of rubber latex

(natural rubber) as it was an important product during the industrial revolution era. Notable

among them was Charles Mariѐ de La Condamine who brought rubber seeds to Acadèmic

Royale des Sciences of France in 1736 and published the first scientific properties of rubber

(Charles, 2002).

Joseph Priestley, an English scientist discovered in 1770 the erasing nature of rubber on pencil

mark on papers, earning it the name ‘rubber’. Vulcanization of rubber with sulphur to improve

9
its elastic and resistance properties was discovered by Charles Goodyear and Hancock in 1839

(Charles, 2002).

In 1876, Sir Henry Wickham, a British explorer took 70,000-Para rubber tree seeds out of

Brazil to the Kew botanical gardens in London, England for research where different resistant

varieties of the seeds were developed. The developed species were sent to the British colonies

of Southeast Asia countries of Malaysia, Sri Lanka, and Singapore in 1876 for planting (Reshad

et al., 2015). The most successful application of latex rubber was achieved in the 1880s when

rubber was found to be the basic material for pneumatic tyres for automotive industry.

Obviously, the well-organized plantations in the Southeast Asia countries resulted in a

significant increase in productivity. The competition from Southeast Asia brought down the

decades of rubber latex boom to an end for much of northern Brazil in the early 1900s. In the

later part of the 19th century, Indonesia, Thailand, Malaysia, China, India, and Vietnam joined

the league of natural rubber producers. Rubber seed (Figure 2.2) was introduced into Nigeria

in 1895 from the Kews botanical gardens in London, and the first rubber plantation established

at Sapele, Delta State in 1903 (Omo-Ikerodah et al., 2009). Rubber tree has 6-9 years of

gestation period before tapping for latex, though the seeds are available between 4-6 years

(Onoji et al., 2016a; Eka et al., 2010). A recent statistical analysis put the global harvested areas

of rubber trees plantations at 11,096,400 ha (FAOSTAT, 2014), thus eliminating competition

for land space with vegetable crops (Zhu et al., 2014). Rubber trees have been reported to grow

to a height of 34 to 40 m in the wild (Takase et al., 2015; Ashraful et al., 2014; Kumar &

Sharma, 2011). It requires warm humid climate with temperatures ranging from 20 to 35 oC, a

fairly-distributed rainfall of about 1800 to 2000 mm throughout the year (Ashraful et al., 2014;

Omo-Ikerodah et al., 2009) and growth is most rapid at altitudes below 200 m (Reshad et al.,

2015). The analysis depicted in Figure 2.3 identified Southeast Asia countries of Indonesia,

Thailand and Malaysia as major areas where rubber trees are cultivated; Nigeria and Côte

10
d'Ivoire are Africa dominant. Other Africa countries where rubber trees are cultivated include

Democratic Republic of Congo, Cameroon, Republic of Congo, Central African Republic,

Ghana, Liberia, Gabon and Guinea (Onoji et al., 2016a). Seeds from rubber tree considered as

wastes over the years except for seed propagations (Kittigowittana et al., 2013), is the second

most valuable product after the latex as source of non-edible oil (Banković-Ilić et al., 2012).

Researchers in Asia (Zhu et al., 2014) investigated the factors influencing rubber seed yield

through in-situ counting of rubber seeds in Xishuangbanna, China, and reported age, size and

clonal types as dominant. Other factors include powdering mildew disease, abnormal leaf

disease, phytophora disease, and weather (Iyayi et al., 2008). A typical rubber plantation has

about 350-500 trees/ha with each tree yielding about 800 seeds and an average of 70 to 500 kg

seed/ha on annual basis (Abdulkadir et al., 2014; Ebewele et al., 2010a; Eka et al., 2010).

Figure 2.1: Rubber trees plantation (Adapted from Onoji et al., 2016a)

11
Figure 2.2: Rubber seed (Adapted from Onoji et al., 2016a)

Cotê d'Ivoire
Philippines 2%
2%
Burma Others
2% 7%
Nigeria
3%
India
Indonesia
4%
33%

Vietnam
5%

China
6%

Malaysia
10%
Thailand
26%

Figure 2.3: Top ten leading producers of rubber seeds in 2014 (Adapted from FAOSTAT,

2014)

12
2.2 Rubber seed capacity in sub-Saharan Africa

In consideration of the nature of non-edible oil plants such as rubber, jatropha, castor, linseed,

moringa oleifera, cotton, neem, and tobacco for energy production to avert food-fuel crisis,

rubber tree has fewer investigation on its potential in many sub-Saharan Africa (SSA) countries

where it is cultivated for latex production. Onoji et al. (2016a) provided a holistic comparative

data relevant to rubber seed as a significant source route of biodiesel production for the

sustainable development of SSA countries. Using statistical data from Food and Agriculture

Organization (FAO) of the United Nations for 2013, Onoji et al. (2016a) reported the hectares

(ha) of rubber plantations cultivated in SSA as follows: Nigeria (345,000 ha), Côte d'Voire

(135,000 ha), Liberia (76,000 ha), and Cameroon (55,000 ha). Others are Ghana (26,800 ha),

Gabon (15,000 ha), Guinea (10,800 ha), Democratic Republic of Congo (50,500 ha), Republic

of Congo (2,400 ha) and Central African Republic (CAR) (1,250 ha). With a plantation density

of 350 trees/ha, 150 kg/ha and a projected 717,750 ha, the researchers reported that SSA could

produce 17,947 ton (RSO) and 16,691 ton (biodiesel). Table 2.1 presents the details of their

findings. The cultivation of rubber trees will sustain the production of rubber products from

latex. It will also provide seed oils for biodiesel that will create additional financial benefits to

the plantation farmers and reinvigorates the economies of the local communities in sub-Saharan

Africa.

13
Table 2.1: Estimated annual rubber seed oil and biodiesel production capacity for SSA countries in 2013 (Onoji et al., 2016a)

Country Harvested Trees/ha Trees Seed (ton/ha), Seed (ton) RSO (ton) Biodiesel (ton)
area (ha), (million) B A×B = RStotal = RStotal × 0.1667 = RSO× 0.93
A

Nigeria 345,000 350 120.75 0.15 51,750 8,626.725 8,022.854

Côte d'Voire 135,000 350 47.25 0.15 20,250 3,375.675 3,139.378

Liberia 76,000 350 26.6 0.15 11,400 1,900.380 1,767.353

Cameroon 55,000 350 19.25 0.15 8,250 1,375.275 1,279.006

D.R. Congo 50,500 350 17.675 0.15 7,575 1,262.753 1,174.360

Ghana 26,800 350 9.38 0.15 4,020 670.134 623.225

Gabon 15,000 350 5.25 0.15 2,250 375.075 348.820

Guinea 10,800 350 3.78 0.15 1,620 270.054 251.150

Rep. Congo 2,400 350 0.84 0.15 360 60.012 55.811

CAR 1,250 350 0.4375 0.15 187.5 31.256 29.068

Total 717,750 251.2125 107,662.5 17,947.339 16,691.025

14
2.3 Rubber seed collection and processing

The matured rubber seeds dehisce from the trees with a loud sound similar to that of a military

rifle during the short period of dry season between the months of July and October in Nigeria

(Onoji et al., 2016a), although it varies from one geographical region to another (Iyayi et al.,

2008; Igeleke & Omorusi, 2007). Overgrown weeds make seed collection a difficult task as

most seeds hide under the cover of weeds. During the period of rainfall, the seed pods will not

dehisce to release the seeds as they pick up moisture that facilitates its deterioration (Iyayi et

al., 2008). Sunshine is therefore a critical factor required to dehisce the seeds for a good yield.

Igeleke & Omorusi (2007) listed excessive moisture, fungi, and pests such as insects and

rodents as major causes of post-harvest deterioration of rubber seeds, and suggested that the

seeds be dried to a moisture level of 7% or less, before storage. Aniamaka & Uriah (1990)

suggested that the seeds should be collected as soon as they dehisce from the trees, and

collection period repeated at least every fourth day to avoid deterioration. Seed that is not

properly stored after harvesting could be affected by chemical reactions that may increase the

free fatty acids content of the seed and reduce its potency for biofuels production. Drying of

seeds at 70 oC or less will halt biochemical processes and reduce cyanide content; chemical

treatment with 2% primiphos methyl and aluminum phosphide, and storage in polypropylene

bags were suggested methods to extend the shelf life of post-harvested seeds in bulk storage

(Igeleke & Omorusi, 2007). The seeds enclosed in a pod of 3-ellipsoidal capsules are mottled

brown with glossy surfaces. They are oval-shaped, 2–5 cm long and average weight of 2–6 g

(Onoji et al., 2016a).

2.4 Biomass valorization processes

Biomass covers organic matter that are traceable to plant or animal origin, either as raw (wood,

energy crops, and agricultural residues, etc.) or processed (effluents, food processing residues,

15
and green wastes, etc.). Biomass materials are generally composed of cellulose, hemicellulose,

lignin, lipids, proteins, simple sugars, and starches, depending on its origins (Claude et al.,

2016). Among these compounds, cellulose, hemicellulose, and lignin are the three main

constituents as shown in Table 2.2 (Zhang et al., 2010). The renewable aspect that is attributed

to biomass can be explained by the carbon cycle: the carbon dioxide emitted during its use is

compensated for by the carbon stock accumulated during its growing stage. Therefore, the

biomass can only be considered as a clean and renewable energy source if obtained in a

sustainable manner (Kumar et al., 2009). Bioenergy is a term that refers to all the processes

whether industrial or not, that can produce energy from biomass. A general view of the

worldwide energy consumption in 2013 depicts renewable energy accounting for about 19%

(Claude et al., 2016). Biomass can be converted into energy or byproducts via versatile

transformation technologies such as physico-chemical or thermo-chemical as shown in Figure

2.4. It is expected that at the dawn of the 22nd century when the use of fossil fuels would be

limited, biodiesel that currently accounts for 0.15% of the share of renewable energy will

become a major source of energy from biomass.

16
Table 2.2: Typical levels of cellulose, hemicellulose, and lignin in biomass (Zhang et al.,

2010)

Component Dry weight (%) Description

Cellulose 40–60 A high-molecular weight (106 or more) linear chain of


glucose linked by β-glycosidic linkage. This chain is stable
and resistant to chemical attack.

Hemicellulose 20–40 Consists of short, highly branched chains of sugars (five-


carbon sugars such as σ-xylose and ι-galactose, and six-
carbon sugars such as σ-galactose, σ-glucose, and σ-
mannose) and aromic acid. Lower molecular weight than
cellulose. Relatively easy to be hydrolysed into basic sugars.

Lignin 10–25 A biopolymer rich in three-dimensional, highly branched


polyphenolic constituents that provide structural integrity to
plants. Amorphous with no exact structure. More difficult to
be dehydrated than cellulose and hemicelluloses.

17
Thermochemical
processes
Heat
Combustion Combustion
Electricity
Biomass Bio-oils
Pyrolysis Bio-chars

Reforming Bio-fuels
- Energy crops
(biodiesel, bio-
(rubber, jatropha,
cotton ...) Gasification Syngas ethanol)

- Agricultural and Bio-materials


urban wastes Physicochemical
(bio-based
transformations
polymers)
- Food and industrial
processing residues
Oil extraction

Syngas
Fermentation

Heat
Methanization Combustion
Reforming
Electricity

Figure 2.4: Overview of main biomass valorizations (Adapted and modified from Claude

et al., 2016)

2.5 Modeling and optimization approach to industrial research

Modeling and optimization of process variables to increase the efficiency of a process is one

of the most important stages in a biochemical process (Baş & Boyaci, 2007a). The traditional

method of optimization is expensive, complex and time consuming due to large amount of

experiments required. It considers one variable at a time, while other factors are kept constant

and the interactive effects of the variables on the response of the system are neglected (Ma et

al., 2016; Kostić et al., 2013; Baş & Boyaci, 2007a). The disadvantages of the traditional

method prompted researchers to look for alternatives that will describe the entire system

comprehensively. In the present day, industrial research uses response surface methodology

18
(RSM), and artificial intelligence tools such as artificial neural network (ANN) and genetic

algorithm (GA), fuzzy logic (FL), ant algorithm (AA) and particle swarm optimization (PSO)

to effectively optimize a process (Gueguim Kana et al., 2012).

Response surface methodology (RSM) is a collection of statistical and mathematical techniques

useful for developing, improving, and optimizing processes in which responses of interest are

influenced by several process variables and the objective is to optimize these responses (Kostić

et al., 2013; Baş & Boyaci, 2007b). RSM defines the effect of these variables, alone and/or

their interactive effects on the response of the process. RSM offers a large amount of

information from a small number of experiments to explain the behavior of the system (Baş &

Boyaci, 2007b). RSM generates a second-order polynomial predictive model that relates the

response of a process to the independent variables considered. On the other hand, ANN is a

modeling tool inspired by biological neural networks which consists of interconnected group

of artificial neurons (input, hidden and output), and processes information using a connectionist

approach to computation (Kostić et al., 2013). In an ANN, a neuron sums the weighted inputs

from several connections and then applies a transfer function (e.g. sigmoid) to the sum. The

resulting value is propagated through outgoing connections to other neurons. Figure 2.5 depicts

a typical multilayer structure of an ANN model (Gueguim Kana et al., 2012). The ANN model

describes processes that exhibit nonlinearities better than RSM, since it does not require

theoretical knowledge or experience during the training process (Zahedi & Azarpour, 2011). It

has the inherent ability to learn and generalize data by producing reasonable outputs, for inputs

that were not encountered during the training process, and update easily when new set of data

are submitted to it. The limitation in process optimization by ANN model is overcome by

coupling a cost-effective and less time-consuming GA as described in Figure 2.6 (Chuahy &

Kokjohn, 2017). Design of experiments (DOEs) such as central composite design (CCD), Box–

19
Behnken design (BBD), and Taguchi experimental design (TED) employing RSM and ANN–

GA optimization techniques are commercially available for research. Dhawane et al. (2015)

employed CCD approach using RSM to study the influence of process variables (methanol/oil

ratio, catalyst loading, temperature) at a reaction time of 1 h on the yield of biodiesel produced

from Hevea brasiliensis oil. They reported an optimal yield of 89.3% at the following

conditions: methanol/oil ratio (15:1), catalyst loading (3.5 wt.%), and reaction temperature (60
o
C). A BBD-RSM was used by Dwivedi & Sharma (2015) to study the optimization of biodiesel

synthesized from pongamia oil and reported a yield of 98.4% at optimal process conditions of

MeOH/oil ratio (11.06:1), KOH catalyst loading (1.43% w/w), reaction time (81.43 min), and

reaction temperature (56.6 oC). Taguchi optimization method was used to investigate the

influence of process variables (MeOH/oil ratio, reaction temperature, and catalyst amount) at

constant reaction time (3 h) on biodiesel yield from Kapok (Ceiba pentandra) oil (Norazahar

et al., 2012). They reported 98% maximum yield at conditions of 55 oC, 2 wt % KOH catalyst,

and 8:1 methanol/oil ratio.

Hidden layer
Output layer
Input layer

Input variable 1

Input variable 2
Output

Input variable 3

Input variable 4

Figure 2.5: Multilayer structure of an ANN model (Adapted and modified from Gueguim

Kana et al., 2012)

20
Initial DOE

Design-Expert® Software
Factor Coding: Actual
% RSO
Design points above predicted value
Design points below predicted value
43
44.1124
28.64
50
X1 = A: Rubber seed powder weight (g)
X2 = B: Solvent volume (ml) 45
Actual Factor

R S O
C: Extraction time (min) = 50.00 40

35

%
30

25

200.00
40.00
212.50 45.00
225.00 50.00

237.50 55.00
Solvent volume (ml) Rubber seed powder weight (g)
250.00 60.00

1st response surface

Search surface for optimal using GA

Restrict DOE space


based on results

Design-Expert® Software
Factor Coding: Actual
% RSO
Design points above predicted value
Design points below predicted value
43
44.1124
28.64
50
X1 = A: Rubber seed powder weight (g)
X2 = B: Solvent volume (ml) 45
Actual Factor
R S O

C: Extraction time (min) = 50.00 40

35
%

30

25

200.00
40.00
212.50 45.00
225.00 50.00

237.50 55.00
Solvent volume (ml) Rubber seed powder weight (g)
250.00 60.00

Refined response surface

No Yes
Is optimal Search surface with GA and
in DOE within find optimal point
interval?

Figure 2.6: General flowchart of Genetic Algorithm (Adapted and modified from Chuahy

& Kokjohn, 2017)

2.6 Qualitative and quantitative analytical procedures

Various spectroscopy methods of analysis have been developed to quantify and/ or quantify

the relative proportions of different materials present in compounds such as lipids and solids

(Guillèn & Ruiz, 2003). These methods are relatively fast, and provide great deal of

information on the material under consideration. Some of the spectroscopy methods include

Fourier transform infrared and scanning electron microscopy. Others include Brunauer-

Emmett-Teller analysis, nuclear magnetic resonance, gas chromatography-mass spectroscopy,

21
x-ray diffraction, and x-ray fluorescence. These methods of analyses and information

obtainable from them are discussed below.

2.6.1 Thermogravimetric analysis (TGA)

The application of thermal techniques of analysis has gained much attractiveness in recent

years. Thermal techniques of investigation are methods used to characterize a system, an

element of a compound (Coats & Redfern, 1963). Thermogravimetry analysis has received

immense patronage in the understanding of solid fuel substrate degradation to release energy

(Islam et al., 2016). The TGA traces the thermal degradation patterns of fresh biomass, char

and hydro-char under air and inert environment for subsequent kinetic studies (Molintas &

Gupta, 2011). The patterns revealed the response of heating rates, the inherent biomass

properties and severity that define the char combustion profiles in the form of thermographs

(Ceylan & Goldfarb, 2015). The degradation of biomass by pyrolysis occurs over wide

temperature ranges. For instance, holocellulose was reported to decompose between 200 and

380 oC, and lignin component degrades over wide temperatures (180–600 oC) to char

(Slopiecka et al., 2012). A typical TGA studies reveal three stages of differential weight loss

of biomass combustion that include dehydration, devolatilization and char oxidation (Islam et

al., 2016). The two leading methods include; differential thermal analysis and

thermogravimetric analysis, where changes in the system weight are measured as a function of

temperature (Coats & Redfern, 1963).

2.6.2 Fourier transform infrared spectroscopy (FT-IR)

One of the most common spectroscopic techniques applied by organic and inorganic chemists

is the Fourier infrared spectroscopy. FT-IR is used to study surface chemistry of materials

or substance and equipped with the ability of conveying information about the

22
molecules in a specimen with their respective concentrations (Smith, 2011), and where

all infrared frequencies in a specimen are detected simultaneously (Bohre, 2013). FT-

IR as an analytical technique has been widely used to monitor the quality of biodiesel/diesel

blends, including checking for fuel adulteration (Valente et al., 2016). It has been reported for

use in the analysis of free fatty acids (Ismail et al., 1993), anisidine determination (Man &

Setiowaty, 1999), and monitoring of the transesterification of oils with methanol to biodiesel

(Knothe, 1999). Infrared spectroscopic analysis is also used to identify the chemical

functional groups present in a specimen undergoing analysis (Bohre, 2013).

2.6.3 Scanning electron microscopy (SEM)

Scanning electron microscopy is a technique used to investigate morphology of a

designated sample by scanning with a beam of electron (Khan, 2014). Interactions

between electrons and atoms in a sample creates signals that are detected, and convey

information about the sample’s composition, morphology and surface topology. The

sample to be analysed should be conductive (Hosmani, 2014). A non-conductive

material to be analysed is usually coated with a conductive material such as gold,

gold/palladium alloy, or platinum prior to analysis.

2.6.4 Brunauer- Emmett-Teller (BET) Analysis

The Brunauer-Emmett-Teller explains the theory behind the physical adsorption of gas

molecules on a solid surface. It serves as a basis for an analytical technique to measure

the specific area of a solid and other textural properties such as pore size and pore

volume that play important role in its catalytic behaviour (Lowell et al., 2004). Since

the surface area of a solid cannot be calculated directly from particle size information,

23
a method to do this is to determine its surface area at an atomic level. The amount of

gas adsorbed is a function of the total amount of exposed surface. It is also a function

of temperature, gas pressure and the strength of interaction between gas and solid

(Lowell et al., 2004). Nitrogen gas is commonly used for BET analysis because of high

purity, availability, and relatively strong interaction with most solid surfaces.

2.6.5 Nuclear magnetic resonance (NMR)

Proton (1H) and carbon-13 (13C) NMR spectroscopy is a technique that is useful to determine

the proportion of different acyl groups present in oils and fats, and other liquids in a very short

time (Onoji et al., 2016b). The area of the signals of the 1H NMR spectra is proportional to the

number of hydrogen atoms that produce the signals. This method has many advantages in

relation to the classical methods because it does not involve chemical modification of the

sample.

2.6.6 Gas chromatography-mass spectroscopy (GC-MS)

Gas chromatography is a process of separating component (s) from the given sample by using

a gaseous mobile phase. It involves a sample being vapourized and injected onto the head of

the chromatographic column (Karger et al., 1973). The sample is transported through the

column by the flow of inert gas (Helium or nitrogen) or hydrogen gas. The components are

recorded as a sequence of peaks as they leave the column. The peaks can be read and analysed

to determine the exact components of the mixture. The components are recorded as a sequence

of peaks as they leave the column. The peaks can be read and analysed to determine the exact

components of the mixture (Skooge at al., 2007). The number of peaks determines the number

of components in a sample. The amount of a given component in a sample is determined by

the area under the peaks. The identity of components can be determined by the given retention

24
times. Detecting the mass of the individual compound allows for conclusive of the sample

(Willard et al., 1981). This is confirmed by the mass spectrometer coupled to it to obtain the

compound’s molecular weight.

2.6.7 X-ray diffraction and X-ray fluorescence of solid materials

The powder X-ray diffraction is a non-destructive technique used to identify crystalline phases

and orientation of materials. It determines the structural properties and atomic

orientation/arrangement of the material. The diffractometer will create the IRD pattern of the

sample, measure d- spacing and obtain integrated intensities. The data are compared with

known standards in the Joint Committee on Powder Diffraction Standards (JCPDS) site, which

are for random orientations (Cullity, 1978).

XRF is a non-destructive technique used for the determination of elemental compositions of a

material. An X-ray source is used to irradiate the sample and to cause the elements in the sample

to emit (or fluoresce) their characteristics X-rays. A detector system (Wavelength dispersive)

is used to measure the position and intensity of peaks of the emitted X-rays for qualitative and

quantitative measurements of the elements and their amounts.

2.7 Rubber seed analysis

Eka et al. (2010) analysed rubber seed to determine its potential use as proteins, fats,

carbohydrate and amino acids as food supplements for livestock. They reported that well

processed rubber seed could be used as food claiming that good storage and heat treatment can

reduce the levels of poisonous hydrogen cyanide present in the seed. The results of the

proximate analysis and amino acids profile of rubber seed are presented in Table 2.3 and Table

2.4, respectively. The use of rubber seed shell (RSS) and rubber seed kernel (RSK) as source

materials for biofuel has been reported in literature. Hassan et al. (2014) investigated RSS and

25
RSK before oil extraction as raw materials for bio-oil production. They determined their

proximate (Table 2.5) and ultimate (Table 2.6) compositions, extractives, cellulose,

holocellulose and hemicelluloses content (Table 2.7) and, calorific value and oil yield (Table

2.8). The ultimate analysis from their report shows that carbon and oxygen are the major

components of RSS and RSK while other elements are in minority. The proximate analysis

indicates that volatile matters in both samples are a major component. High amount of volatile

matter indicates that the material could liquefy easily. The knowledge of the number of

extractives (Waxes, fats, resins, gum, sugars, starches, sterols, flavonoids, etc.) helps to

estimate exact amount of fermentable sugar present in biomass, and should be extracted before

processing the lignocelluloses material as a biofuel feedstock (Sasmal et al., 2012). The report

also shows that cellulose contents of RSS and RSK are significantly different. Kim et al. (2012)

reported that cellulose content of approximately 40% would be useful for biofuel production.

From the analysis presented by Hassan et al. (2014), RSK is a more suitable material for biofuel

production than RSS. The results presented in Table 2.8 shows that the calorific value of both

samples is in the range 23–28 MJ/kg, and the highest oil yield of 33.1% was obtained from

RSK. The TGA results presented show that major degradation of RSS and RSK was observed

at 300 and 350 oC, respectively. The differential thermo-gravimetric analysis (DTG) showed

two major weight losses for both samples. The first major loss occurred within 250–350 oC and

the second around 400 oC. Weight losses resulting from hemicelluloses and cellulose occurred

at the region between 250–350 oC, while the lignin degradation starts at 500 oC. Sun & Jiang

(2010) investigated the use of RSS as a raw material for the production of activated carbon by

the physical activation with steam. Their results of the proximate analysis (Table 2.5) obtained

for RSS were very close to those reported by Hassan et al. (2014). Other results presented by

Sun & Jiang (2010) show that RSS is a good precursor for activated carbon. The optimal

activation conditions, reported are pyrolyzed temperature (880 oC), steam flowrate (6 kg/h),

26
and residence time (60 min). The characteristics of the activated carbon from RSS with a yield

(30.5%) reported are specific surface area (SBET) 948 m2/g, total volume (0.988 m2/kg), iodine

number of adsorbent (qiodine) 1.326 g/g, amount of methylene blue adsorption of adsorbent (9

mb) 265 mg/g, and hardness 94.7%. The results of TGA show two stages of decomposition as

follows: stage one (273–385 oC) and stage two (385–800 oC). Little weight loss was observed

at temperature below 270 oC because the rubber seed had been dried under 105 oC for 3 h. The

TGA results from both researchers are very similar as presented above. The DSC analysis

presented by Sun & Jiang (2010) shows a little drop at a temperature of 100 oC, because the

sample needs heat to evaporate the remaining water. The reported decomposition temperatures

for hemicelluloses and cellulose are 180 and 240 oC, respectively. When temperature was about

320 oC, DSC arrived at valley, where hemicellulose and cellulose decomposed and produced

abundant organic compounds (wood tar, ketone and methanol) which would vapourized by

adsorbing large heat. At 500 oC, DSC curves climbed up steeply and reached at a peak, because

at this step lignin pyrolyzed and emitted abundant heat. At 800 oC, DSC curve reached a base

line, pyrolysis reaction was over and the residue reported as graphite carbon and ash (Sun &

Jiang, 2010).

Table 2.3: Proximate analysis of rubber seed (Eka et al., 2010)

Parameter Value

Moisture, % 3.99 ± 0.01

Protein, g/100g 17.41 ± 0.01

Fat, g/100g 68.53 ± 0.04

Ash, g/100g 3.08 ± 0.01

Total Carbohydrate, % (by difference) 6.99

27
Table 2.4: Amino acid profile of rubber seed (Eka et al., 2010)

Non-essential amino acids (g/100g protein) Essential amino acids (g/100g protein)

Aspartic acid 11.18 ± 1.57 Threonine 3.72 ± 0.42

Serine 5.89 ± 0.38 Valine 7.08 ± 0.27

Glutamic acid 16.13 ± 0.34 Methionine 1.37 ± 0.31

Glycine 5.14 ± 0.00 Lysine 4.26 ± 0.41

Histidine 2.95 ± 0.25 Isoleucine 3.28 ± 0.04

Arginine 12.45 ± 0.19 Leucine 6.81 ± 0.08

Alanine 4.71 ± 0.15 Phenylalanine 4.88 ± 0.46

Proline 6.77 ± 0.02

Cysteine 0.78 ± 0.79

Tyrosine 2.99 ± 0.25

Table 2.5: Proximate analysis of RSS and RSK (Hassan et al., 2014; a Sung & Jiang, 2010)

Biomass Moisture Ash Volatile mater Fixed carbon

RSS 14.3, 14.37a 0.1, 0.17a 71.7, 71.76a 13.9, 13.97a

RSK 4.3 0.2 89.4 6.1

Table 2.6: Ultimate analysis of RSS and RSK (Hassan et al., 2014)

Biomass C H N S O*

RSS 48.8 5.9 1.5 0.1 43.7

RSK 64.5 8.2 3.6 0.3 23.4

28
Table 2.7: Extractives, holocellulose, hemicelluloses and cellulose contents of RSS and

RSK (Hassan et al., 2014)


*
Biomass Extractives Holocellulose Hemicelluloses Cellulose

RSS 7.8 92.2 66.4 25.8

RSK 3.6 96.4 26.9 69.5

*calculated by difference

Table 2.8: Calorific value and oil yield of RSS and RSK (Hassan et al., 2014)

Calorific value (MJ/kg) Oil yield (%)


Biomass

RSS 23.9 8.7

RSK 27.5 33.1

2.8 Rubber seed oil extraction methods

Industrially, several methods are employed for oil extraction such as mechanical (hydraulic

and screw press), solvent extraction, enzymatic, aqueous, supercritical fluid extraction, and

their combinations (Onoji et al., 2016b). Mechanical type of extraction is widely used and well

adapted to rural communities with moderate initial and operating costs, but oil yield is low

(Subrotoet al., 2015; Willems et al., 2008; Olajide et al., 2007). Modeling and optimizing the

extraction process variables is paramount to obtaining a higher oil yield. Such optimized

parameters could provide vital information to researchers and industrialists. Figure 2.7 depicts

a typical flowchart for rubber seed oil processing with a combination of mechanical and

chemical methods. The seeds are dried, crushed, and milled to reduce particle size. Milled seed

(meal) was conditioned with scorching gas in a heated rotary dryer for 10–20 min at 60–70 oC

(Iyayi et al., 2008). It was then introduced into a mechanical screw press to recover rubber seed

29
oil, cake and foots (wastes). Rubber seed oil was recovered from the foots using a filtration

unit and stored for further use. More rubber seed oil was further extracted from the oily cake

in a solvent extraction unit using n-hexane. The cake was stored in a storage unit, and oil-

hexane mixture vacuum evaporated to recover the oil, while n-hexane solvent recycled.

Seed storage Kernel


Drying unit Crushing storage unit
unit
unit

Solvent Conditioning Milling Seed shell


tank unit unit
Foots
oily oil +
cake foots Filtration Rubber seed oil Rubber seed oil
Cake unit Screw
Solvent unit storage unit
press unit
extraction unit

Rubber seed oil


Rotary
oil + solvent evaporator

Recovered solvent

Figure 2.7: Flowchart of rubber seed oil processing (Adapted from Onoji et al., 2016a)

2.8.1 Conventional extraction method (Mechanical and chemical)

Conventional method uses mechanical and chemical means to extract the oil. The mechanical

type (mainly hydraulic and screw press) is relatively inexpensive after the initial capital costs

(Onoji et al., 2016a; Li et al., 2013). Quality of the oil is high, but it has low yield compared to

other extraction methods (Ebewele et al., 2010a). Mechanical screw press is relatively

inefficient because it leaves 8–14% of the available oil in the de-oiled cake (Ali & Watson,

2014). Singh & Bargale (2000) recommended the use of double-stage compression expellers

to enhance oil recovery. With increased pressure and heated screw press head, oil could easily

be extracted from pre-treated seeds (Ebewele et al., 2010a; Zheng et al., 2003). The chemical

method is widely used in the laboratory using n-hexane or other suitable organic solvents as

30
extractive media. Co-extraction of undesirable components in the oil results to low quality.

Boiling point of solvent, solvent/seed ratio, seed size, moisture content, drying and extraction

time are some of the parameters that affect the yield (Onoji et al., 2016b). Ebewele et al. (2010a)

employed hydraulic press to study the influence of extraction variables-pressure (5–8 MPa),

temperature (40–90 oC), moisture content (7–16%), and particle size (1.16–3.36 mm) on the

oil yield of different rubber species. They reported that optimal yields for the different species

were obtained at 8 MPa, 70 oC, 10%, and 1.16 mm. Sabarish et al. (2016) studied the effects of

drying time, drying temperature, pressing time, and pressure on the yield of rubber seed oil

extracted with hydraulic press using design of experiment and a 24-full factorial design for the

response surface analysis. The analysis of variance (ANOVA) report indicates that the

maximum oil yield was obtained at the following conditions: drying temperature 55 oC, drying

time 1 h, pressure 35 kg/cm2, and pressing time 25 min. Sayyar et al. (2009) in their study on

the extraction of oil from jatropha seed using n-hexane solvent reported variation in oil yields

(46–47.3%) at 8 h extraction time, 0.16 solute/solvent ratio and particle size (0.5–0.75 mm).

Kostić et al. (2013) considered the effect of solvent/seed ratio (3:1, 6.5:1 and 10:1 ml/g),

extraction temperature (20, 45 and 70 oC) and extraction time (5, 10 and 15 min) on hempseed

oil yield using RSM and ANN optimization techniques. They reported optimal yield of 29.56%

based on ANN conditions: solvent/seed ratio (10:1), temperature (70 oC) and extraction time

(10 min). Bokhari et al. (2012) reported similar observations on extraction of rubber seed oil

using RSM. They obtained an optimal yield of 33.56% with n-hexane as solvent at 0.027 g/ml

solute/solvent ratio and 4.5 h extraction time. Reshad et al. (2015) also applied RSM using

central composite design for the extraction of oil from rubber seeds with n-hexane as solvent.

They reported a maximum yield of 49.36% at optimal conditions of 0.08 g/ml solute/solvent

ratio, 1 mm particle size, and 8 h extraction time.

31
2.8.2 Gas-assisted mechanical extraction method

This process involves the dissolution of supercritical CO2 fluid into the milled seed followed

by the screw pressing process (Ebewele et al., 2010a). It allowed for the use of lower pressure,

while maintaining higher yield without compromising quality (Onoji et al., 2016a).

2.8.3 Microwave-assisted extraction (MAE) method

Microwaves are electromagnetic waves; a combination of a vertical magnetic field and an

electrical wave, with wavelengths ranging from 1 mm to 1 m (Ma et al., 2016). Microwave

irradiation follows two mechanisms; viz. dipolar rotation and ionic conduction that generate

electromagnetic field that accelerates the extraction process (Ma et al., 2016; Li et al., 2013).

The process allows shorter time, low solvent consumption, higher yield and much lower energy

requirements compared to the conventional methods (Ali & Watson, 2014). Gimbun et al.

(2012) in their study on rubber seed oil extraction compared the efficacy of MAE with Soxhlet

method. They obtained a higher oil yield of 40% with MAE and 36% with Soxhlet technique.

They further reported a shorter extraction time of 15 min for MAE and 6 h for Soxhlet method.

The efficiency of MAE is due to good interactive effect of microwave with solvent molecules.

2.8.4 Ultrasonic-assisted extraction method

The process involves the formation and collapse of microscopic bubbles that release huge

amount of energy, pressure and mechanical shear. The ultrasound penetrates the seeds' cell

tissues and extracts the oil into the solvent by disrupting the cell walls (Ali & Watson, 2014;

Li et al., 2004). Shorter time and low solvent consumption are some of its advantages (Onoji

et al., 2016a; Ali & Watson, 2014).

32
2.8.5 Aqueous enzymatic extraction method

The high efficiency and specificity of enzymes allow them to be useful extractive media. They

hydrolyze and rupture the cell walls of the seed kernels and release their oil constituent. Since

no solvent is involved, this method is eco-friendly with low initial cost accompanied with the

ease of processing of the de-oiled cake into livestock feeds (Li et al., 2004).

2.8.6 Supercritical fluid extraction (SFE) method

SFE is a separation process where the substances are dissolved in a fluid that is able to modify

its dissolving power under specific condition above their supercritical regions-critical

temperature (Tc) and pressure (Pc) (Rodríguez-Solana et al., 2014). Supercritical fluids have

the ability to penetrate into the micropores of a solid and good transport properties (Raventós

et al., 2002). Among the available supercritical fluid, carbon dioxide is considered the best

because of its inherent advantages of non-toxic, non-flammable, cost-effective, and it has a low

Tc (304.15 K). Efficiency of CO2 can be enhanced by adding a co-solvent such as methanol

that is capable of hydrogen bonding and dipole-dipole interactions with the contents of the seed

(Murga et al., 2000).

2.9 Physico-chemical properties of rubber seed oil (RSO)

The physicochemical properties of rubber seed oil reported by various researchers are available

in literature. The determination of these properties enables the evaluation of the oil as a source

of potential industrial feedstock for biodiesel production and other uses (Onoji et al., 2016a).

The common fatty acids reported for rubber seed oil are palmitic (C16:0), stearic (C18:0), oleic

(C18:1), linoleic (C18:2) and linolenic (C18:3) (Kittigowittana et al., 2013; Widayat & Kiono,

2012; Ramadhas et al., 2005). Other types of acid may be present in smaller proportions

(Ashraful et al., 2014). The reports indicate that rubber seed oil consists mainly of unsaturated

33
fatty acids with polyunsaturated acids in higher proportion. The values reported by researchers

may be different depending on the extraction processes, clonal types, analytical methods used,

and geographical locations of the rubber tree (Onoji et al., 2016a). Tables 2.9 and 2.10 present

the physico-chemical properties and the fatty acid compositions, respectively, of rubber seed

oil found in literature. These properties are comparable to other non-edible vegetable oils found

in literature (Table 2.11).

34
Table 2.9: Physico-chemical properties of rubber seed oil

Parameters Researchers a, b, c This study d


Colour golden yellow, light/dark brown dark brown
Density, g/cm3 @ 25 oC 0.857–0.943 0.886 ± 0.002
Specific gravity @ 15 oC 0.91 0.909 ± 0.002
o
API gravity @ 15 oC NA 24.1 ± 0.282
Ph 6 6 ± 0.141
Oil content (wt.%) 35–60 43 ± 0.141
Iodine value, g I2/100 g oil 113–146 137.02 ± 0.028
Peroxide value, meq. O2/kg oil 1.6–16 10.46 ± 0.098
Saponification value, mg KOH/g oil 183.91–235.28 195.30 ± 0.282
Acid value, mg KOH/g oil 1.68–42.41 18.20 ± 0.141
Free fatty acid (%FFA as oleic acid) 0.84–42.412 9.10 ± 0.07
Kinematic viscosity, mm2/s @ 40 oC 6–66 40.18 ± 0.028
Refractive index @ 20 oC 1.46–1.47 1.4707 ± 0.00028
Pour point, oC –9 to –1.5 –6
Cloud point, oC 3–4 5.5
Cold filter plugging point, oC NA –0.025
Flash point, oC 72–295 240.3
Fire point, oC 298 256
Aniline point, oC (oF) NA 21 (69.8)
Boiling point, oC NA 119
Freezing point, oC NA –18
Cetane number 45–49.73 43.42
Higher heating value (HHV), MJ/kg 36.1–44 39.37
Mean mol. weight of fatty acids (g/mol) NA 286.74
Average mol. weight of RSO (g/mol) NA 898
Diesel index NA 15.71
NA
means not available; values are mean ± standard deviation of duplicate data .

a
Onoji et al. (2016a); b Aravind et al. (2015); c Reshad et al. (2015); d Onoji et al. (2016b)

35
Table 2.10: Fatty acids profile of rubber seed oil

Fatty acid/Systemic name Chemical Composition (%)


formulae
Researchers a This study d
Myristic (C14:0) / Tetradecanoic C14H28O2 2.2 -

Palmitic (C16:0)/ Hexadecanoic C16H32O2 0.23–10.6 13.85

Palmitoleic (C16:1)/ Hexadec-9-enoic C16H30O2 0.23–0.25 -

Stearic (C18:0)/ Octadecanoic C18H36O2 5.69–12 16.82

Oleic (C18:1)/ cis-9-Octadecenoic C18H34O2 12.7–42.08 64.11

Linoleic(C18:2)/ cis-9-cis-12-Octadecadienoic C18H32O2 39.6–52.84 -

α-Linolenic (C18:3)/ cis-9-cis-12-cis-15- C18H30O2 2.38–26 -


Octadecatrienoic

Arachidonic (C20:0)/ Eicosanoic C20H40O2 0.66–0.97 -

Erucic acid (C22:1)/cis-13-Docosenoic C22H42O2 - 5.22

Total saturated 8.78–23.57 30.67

Total monounsaturated 12.93–42.33 69.33

Total polyunsaturated (linoleic, linolenic, etc) 41.98–78.84 Trace

a
Onoji et al. (2016a); d Onoji et al. (2016b)

36
Table 2.11: Fatty acid compositions of selected non-edible oils (%) (Onoji et al., 2016a)

Fatty acid Rubber Jatropha Neem Linseed Tobacco

Myristic acid 2.2 0.3–1.4 0.2–0.26 0.045 0.09–0.17


(C14:0)
Palmitic acid 0.23–10.6 12.6–16.0 16–33 5.85–6.21 8.46–10.96
(C16:0)
Palmitoleic acid 0.23–0.25 0.57–3.5 0.24 0.3 0.2
(C16:1)
Stearic acid 5.69–12 5.97–7.4 9–24 5.47–5.63 2.64–3.34
(C18:0)
Oleic acid 12.7–42.08 34.3–44.7 25–54 20.17–24.05 11.24–14.54
(C18:1)
Linoleic acid 39.6–52.84 31.4–43.2 6–16 13.29–14.93 69.49–75.58
(C18:2)
α-Linolenic acid 2.38–26 0.8–3.4 0.56 46.10–51.12 0.69–4.20
(C18:3)
Arachidonic acid 0.66–0.97 0.17–0.3 1.04 0.2 0.25
(C20:0)
Behenic acid – 0.5–0.7 0.3 0.3 0.12
(C22:0)
Oil content (%) 35–60 20–60 20–30 35–45 35–49

Cxx: y means number of carbon atoms: number of double bonds

2.10 Thermo-oxidative analysis of rubber seed oil

Generally, seed oils deteriorate when handled defectively with the major decomposition

reaction being oxidation in the presence of atmospheric air (oxygen). The oxidative stability of

oils is an important indicator of performance and shelf life, and it depends on the fatty acid

compositions and conditions to which it is subjected (Onoji et al., 2016b; Guillèn and Cabo,

2002). Poor oxidative stability, unpleasant odor, filter clogging tendency, low temperature

fluidity and other conditions that make them unsuitable for long-term storage before use are

some of the characteristics of vegetable oils (Oderinde et al., 2009). Thermal oxidation of seed

oil on prolonged heating occurs through a free radical mechanism. This is characterized by the

initial emergence of a sweetish and unpleasant odor that becomes progressively worse until it

37
attains a characteristic smell of rancid fat, yielding peroxides and unstable hydroperoxides as

primary products (Gouveia de Souza et al., 2004; Guillèn & Cabo, 2002). The primary products

degrade easily to produce secondary products such as aldehydes, ketones, acids and alcohols

(Gouveia de Souza et al., 2004; Santos et al., 2002). Oxidative stability of vegetable oils can

be improved by use of various antioxidant additives available in markets or more importantly

by increasing the saturated fatty acids and oleic acid contents in oil by genetic modification of

oil bearing plants (Borugadda & Goud, 2013). Such thermal oxidative reactions can be

investigated using thermo-analytical methods such as differential scanning calorimetry (DSC),

thermogravimetry analysis (TGA), derivative thermogravimetry (DTG) and Fourier transform-

infrared (FT-IR) spectroscopy (Onoji et al., 2016b; Borugadda & Goud, 2013; Gouveia de

Souza et al., 2004). A good thermal decomposition profile of oil can be obtained at a heating

rate of 10 oC/min under inert atmosphere (Borugadda & Goud, 2013). Reshad et al., (2015)

estimated the thermal properties of rubber seed oil (RSO), rubber seed kernel (RSK), and cake

by TGA and DSC analyses. They reported single-stage decomposition for RSO, while cake

and kernel occurred in two and three stages, respectively. A negligible weight loss (<0.2%) of

RSO was observed at 90 oC due to the presence of moisture (free water). Sudden decrease in

weight of the oil sample started from 250–470 oC which was reported to be due to

decomposition of heavier hydrocarbon molecules to lower molecular hydrocarbons, and non-

hydrocarbon gases such as CO2 and CO. TGA confirms RSO is thermally stable up to

temperature of 290 oC (onset point) higher than kernel (210 oC) and cake (120 oC). These

reports collaborated with the findings of Aravind et al. (2015) who investigated the thermal

degradation of RSO, coconut oil, sunflower oil and SAE20W40 under oxygen environment in

the temperature range 0–500 oC and reported that RSO is thermally stable up to 250 oC and

that gradual degradation occurs after 300 oC. The TGA curve of RSO is similar to that of

SAE20W40 as the degradation is gradual unlike coconut and sunflower oil. However, simple

38
and low-cost classic chemical analysis such as determination of the concentration of peroxide

value, iodine value, acid value and refractive index has been effectively employed to evaluate

the kinetics of oils degradation under thermal conditions that influence the mechanisms of

degradation process (Onoji et al., 2016b; Oderinde et al., 2009).

2.11 Industrial applications of rubber seed and oil

Recent scientific studies show that rubber seed and seed oil have several potential industrial

applications such as biodiesel (Onoji et al., 2016a; Vipin et al., 2016; Reshad et al., 2015),

semi-drying oil (Eka et al., 2010; Ebewele et al., 2010a), paints and coating (Aigbodion &

Pillai, 2000), printing ink (Iyayi et al., 2008; Igeleke & Omorusi, 2007). Others are livestock

feeds and fertilizer (Iyayi et al., 2008; Igeleke & Omorusi, 2007), lubricants (Aravind et al.,

2015), powdered activated carbon (Sung & Jiang, 2010; Okieimen et al., 2005a), cosmetic

(Kittigowittana et al., 2013), liquid soap and hair shampoo (Igeleke & Omorusi, 2007) and as

a carrier for copper fungicide (Ebewele et al., 2010a). Matured fresh rubber seed contains 40-

50% kernels (Banković-Ilić et al., 2012), 37% shell (Atabani et al., 2013) and 20–25wt. %

moisture (Gimbun et al., 2012). The seed contains protein, essential and non-essential amino

acids (Eka et al., 2010), while the dried kernel contains 40-50% oil (Onoji et al., 2016a;

Widyarani et al., 2014). Rubber seed contains toxic substances such as cyanogenic glycoside

(186 mg) and about 638-749 mg HCN per kg of fresh kernel (Eka et al., 2010; Iyayi et al.,

2008). HCN constituent can be reduced on storage for a period not less than 2 months at room

temperature (Iyayi et al., 2008). With the emergence of biodiesel industry in SSA countries,

rubber plantation farmers will earn extra profit by harvesting rubber seeds for biodiesel

production. This will also encourage the cultivation of new plantations that will reinvigorate

the economy of these local communities, which had abandoned their plantations due to low

demand for rubber latex in the past (Onoji et al., 2016a).

39
2.12 Biofuels policies and sustainability in Africa

The African continent is plagued with poverty and unrests because of lack of good governance

and accountability. Africa has about 9.5% of the world proven reserves of crude oil that are

concentrated in four countries (Nigeria, Algeria, Egypt, and Libya) and contributes about 12%

to global oil production (Onoji et al., 2016a; Yang et al., 2014; Amigun et al., 2011). The rest

African countries are net importers of energy (Table 2.12) and that places a heavy economic

burden and reduces energy security and sovereignty. Thus, there is an urgent need to invest in

domestic energy infrastructure driven by the private sector for quick economic recovery and

growth within the continent (Yang et al., 2014).

Biofuels production in Africa could be commercialized to diversify energy and agricultural

activity in best practices, reduce dependence on crude oil and contribute to economic growth

in a sustainable manner.

The advantages of biofuels such as biodiesel, bioethanol, biohydrogen, biogas, biomethanol,

etc. are renewability, biodegradability, lower Sulphur and aromatic content, higher combustion

efficiency, lower emissions and amongst others. However, biodiesel and bioethanol are the two

alternative fuels promoted with potential to reduce dependence on fossil oil derived products

imports (Demirbaş, 2009). The development of biodiesel industry in most African countries is

still in the infancy stage with Jatropha curcas as the main non-edible oil source (Yang et al.,

2014). Advanced countries of US, Germany, France, Spain, Italy, and emerging economies like

Brazil and India, have well-articulated and sustainable biofuels policies that enhance their

development. Availability of adequate arable land for bioenergy crops cultivation is a major

challenge to these nations (Amigun et al., 2008).

40
Onoji et al. (2016a) suggested that African countries should articulate institutional policies and

regulatory frameworks to guard against deforestation and biodiversity challenges. This is due

to the presence of foreign investors that will invest in bioenergy crops cultivation for biodiesel

production. Most of the bioenergy policies in draft forms in African countries targeted food

crops as biofuels feedstock. These policies are counterproductive as most Africans depend

heavily on these crops for food. Available reports in literature support rubber tree seeds as

plausible source of non-edible oil to complement jatropha seed oil for biodiesel production in

sub-Saharan Africa (Onoji et al., 2016a). However, none of the sub-Saharan African countries

where rubber trees exist in large quantity considers it as a feedstock for biodiesel production in

their bioenergy policies shown in Table 2.13. This is a policy oversight considering the multiple

applications of rubber plant in automotive and biofuel industry. A case study of Nigeria is

presented to show the lapses in biofuel policies in Africa and energy insecurity.

41
Table 2.12: Africa countries which import and export energy (Yang et al., 2014)

Major energy exporters a Net energy exporters Importers b

Algeria Angola Benin

Rep. Congo Cameroon Eritrea

Egypt Congo Ethiopia

Gabon Côte d'Voire Ghana

Libya D.R. Congo Kenya

Nigeria Gabon Morocco

South Africa Sudan Mozambique

Namibia

Senegal

Tanzania

Togo

Zambia

Zimbabwe

a
Major energy exporter are in excess of 0.5 quads.
b
Most of the African countries' imports are very small (less than 0.3 quads).

42
Table 2.13: Sub-Saharan African countries and their bioenergy policies (Onoji et al.,

2016a)

Country Biodiesel feedstock Presence of policies

D.R. Congo rubber*, oil palm, melon, sesame draft energy policy

Cameroon rubber*, oil palm, melon, None

Rep. Congo rubber*, oil palm, melon policy under development

CAR rubber*, melon, sesame None

Nigeria rubber*, jatropha, oil palm, sesame energy policy, biofuels strategy

Ghana rubber*, jatropha, oil palm energy policy, renewable energy

Côte d'Ivoire rubber*, sesame, tobacco policy under development

Liberia rubber*, pongamia pinnata, oil palm draft biofuel policy

Gabon rubber*, oil palm, pongamia pinnata None

Guinea rubber*, oil palm, sesame None

* means available but wasted, and are not listed as feedstock in biofuels policies.

2.12.1 Nigerian biofuel policy (A case study in Africa)

With estimated reserves of 37.2 billion barrels, a daily production output of 2.2 million barrels

of crude oil and gas reserves of about 187 trillion cubic feet, Nigeria is still facing energy

challenges with about 49% of the population lacking access to electricity. Energy output rated

at about 4000 MW is grossly inadequate, and demand is expected to increase over the next 20

years (Ishola et al., 2013). The declining capacity for energy generation in Nigeria has deeply

entrenched the dependence on renewable biomass fuels for cooking (Abila, 2012). The

adoption of biofuel policy in Nigeria in 2007 holds a diversity of opportunities and potentials

to project Nigeria into a bioeconomy and increase her energy generation (Abila, 2012).

Demirbaş et al. (2009) classified the importance of the biofuel industry as economic,

environmental, and energy security. The biofuel industry could attract investment in new

43
technology, boost rural development through industrialization, and create employment leading

to poverty reduction. It could also reduce dependence on revenue from oil, recycle carbon

dioxide, reduce rural-urban drift of unemployed youths (Wohlgemutgh, 1999), and increase

the income of smallholder farmers integrated into the biofuel supply chain (Malik, 2009).

Ohimain (2013) however reviewed the Nigerian biofuel policy and incentives released in 2007

and observed that:

 the policy narrowly classified biofuel to include only bioethanol and biodiesel

neglecting other biofuels and energy carriers obtainable from biomass such as biohydrogen,

biogas, biomethanol, biocrude, biobutanol, dimethyl ether, and other products of Fisher-

Tropsch (FT) synthesis- FT diesel, FT gasoline, and chemicals;

 The policy classified the biofuel enterprise as an agro-allied industry, yet the policy is

driven by the petroleum sector;

 The policy inadvertently refers to food crops as biofuel feedstock and did not address

the potential food-fuel crisis that could arise from the use of food crops as biofuel feedstock;

 The policy did not adequately address issues pertaining to technology transfer, biomass

collection, waste management (environmental impact) and transportation of biofuel feedstock;

 The policy did not encourage expansion of feedstock sources by investing in research

and development on second and third generation biofuel feedstock that do not compete with

food;

 That modern biomass conversion technology issues were not addressed by the policy.

The researcher suggested upgrade of the policy to accommodate the observed lapses. Ishola et

al. (2013) reported similar observations in US$3.86 billion investment in nineteen ethanol bio-

refineries in Nigeria with unsatisfactory results. This could be attributed to limited availability

of food crops as biofuel feedstock. Many of the food crops are staple food in Nigeria and their

44
use as biofuel feedstock is counterproductive. The report however, did not highlight any

expenditure on biodiesel programme. The adoption of rubber seed as a source of non-edible oil

feedstock will drive the Nigerian biofuel policy to fruition. Nigeria has the capacity to produce

over 8,000 ton of biodiesel/annum from rubber seed oil and this figure could increase with

proper management (Onoji et al., 2016a). Despite the potentials and availability of rubber seeds

in Nigeria, there is no trial production of biodiesel from rubber seed oil except on research

scale, which is even limited.

2.13 Catalysis

A catalyst is a substance that affects the rate of a reaction (positive or negative) without been

substantially affected by the process. Catalysts are neither reactants nor product, and have been

used by man for over 2000 years in several industrial applications (Levenspiel, 2005). Catalysis

is the occurrence, study, and use of catalysts and catalytic processes that involve the constant

search for new ways of increasing product yield and selectivity (Fogler, 2008). There are two

broad classes of catalysts: (1) Enzymes - operate at close to ambient temperature with

biochemical systems, (2) Man-made catalysts - operate at high temperature (homogeneous and

heterogeneous). Homogeneous catalysis concerns processes in which a catalyst is in solution

with at least one of the reactants. A heterogeneous catalytic process involves more than one

phase; usually the catalyst is a solid and the reactants and products are in liquid or gaseous

form. A heterogeneous catalytic reaction occurs at or very near the fluid-solid. Because a

catalytic reaction occurs at the fluid -solid interface, a large interfacial area is essential in

attaining significant reaction rate. In many catalysts, this area is providing by an inner porous

structure needed for the high rate of reaction (Fogler, 2008). A typical silica-alumina cracking

catalyst has a pore volume of 0.6 cm3/g, average pore radius of 4mm, and a corresponding

surface area of 300 m2/g (Fogler, 2008). In some cases, a catalyst may contain some minute

45
particles of an active material dispersed over a less active substance called a support. The active

material is frequently a pure metal or a metal alloy. Catalysts can also have small amounts of

active ingredient added called promoters, which increase the activity of the solid catalysts.

Although a catalyst can easily speed up the rate of a reaction in several folds, when varieties of

reactions are encountered, the catalyst selectivity enables it to affect a single reaction, leaving

the rest unaffected. Thus, in the presence of an appropriate catalyst, products containing

predominantly the material(s) desired could be obtained from a given feed. The following are

some general observations made by Levenspiel (2005) on catalyst:

 The selection of a catalyst to promote a reaction is not well understood. In practice,

extensive trial and error may be needed to produce a satisfactory catalyst;

 Duplication of the chemical constitution of a good catalyst is no guarantee that the solid

produced will have any catalytic activity;

 Various theories have been proposed to explain the action of catalyst, and it is thought

that reactant molecules are changed, energized, or affected to form intermediates in the

regions close to the catalyst surface;

 In terms of the transition-state theory, the catalyst reduces the potential energy barrier

over which the reactants must pass to form products (Figure 2.8);

 Though a catalyst may speed up a reaction, it never determines the equilibrium or

endpoint of a reaction. Thus, with or without a catalyst, the equilibrium constant for a

reaction is always the same.

46
Without catalyst, the complex
has high potential energy
Energy of reactants
resulting in low rate of reaction

Different
With catalyst, the lower energy
reaction paths
barrier allows higher rate of
reaction
Initial
State
Final
State
Reactants Complex Products
Reaction Path

Figure 2.8: Representation of the action of a catalyst (Adapted from Levenspiel,

2005)

2.13.1 Catalyst deactivation

Most catalysts do not maintain their activities at the same levels for indefinite periods. They

are subjected to deactivation, which causes a decline in catalyst's activity as time progresses.

Catalyst deactivation may be caused by: (1) Aging -This is a gradual change in surface crystal

structure; (2) Poisoning-This is the irreversible deposition of substances on the active site; or

(3) fouling or coking- This is the deposition of reversible carbonaceous or other material on the

entire surface.

2.13.2 Physical adsorption and chemisorption in chemical reaction catalysis

For a catalytic reaction to occur, at least one, or frequently all of the reactants must be attached

to the surface of the catalyst (Wu, 2004). This attachment known as adsorption, takes place by

two different processes: physical adsorption (physisorption) and chemisorption

(chemisorption) (Mechrafi, 2002). Physical adsorption mainly caused by van der Waals forces

is the result of low energy exchange between the adsorbate and the adsorbent on the solid
47
surface (Desjardins, 1997). The adsorbate is relatively free to move on the surface, as the

attraction is not fixed to a specific site. Physisorption is relatively reversible and capable of

multilayer adsorption (Madani, 2004). The process is similar to condensation, with small

exothermic heat of adsorption (4.2–63 kJ/mol) (Fogler, 2008). In chemisorption, the adsorbed

atoms or molecules are held to the surface by valence forces of the same type as those that

occur between bonded atoms in molecules (Fogler, 2008). In this case, the adsorbed molecules

are not free to move on the solid surface; thus, the adsorption process depends on the functional

groups present on the adsorbent and not on the surface area. As a result, the electronic structure

of the chemisorbed molecule is perturbed significantly, causing it to be extremely reactive. Due

to the stronger molecular electronic structure that results from the formation of chemical bonds

between adsorbate and adsorbent, chemisorption is an irreversible process. It occurs at a higher

temperature that requires more energy. Like physical adsorption, chemisorption is exothermic

process, but the heats of adsorption are generally of the same magnitude as the heat of a

chemical reaction (i.e., 40–400 kJ/mol). Table 2.14 presents the comparisons between

physisorption and chemisorption.

48
Table 2.14: Comparison between physisorption and chemisorption processes

Physisorption Chemisorption

Adsorbates are molecules Adsorbates are atoms or radicals

Multilayer adsorption process Monolayer adsorption process

Attraction is a result of van der Waals forces Attraction is a result of chemical bonds

Reversible process Irreversible process

Molecules are adsorbed on available sites Molecules are adsorbed on active sites only

Adsorption temperature must be below the It occurs at any temperature

boiling point of the Adsorbate

Heat of Adsorption is less than 50 kJ/mol Heat of Adsorption can be more than 100

kJ/mol (Christmann, 2012)

2.14 Gas-solid adsorption systems

The ability of the surface of a solid to adsorb gases and vapour is considered as its most

characteristic property. Surface area and porosity are important properties in the field of

catalyst design and heterogeneous catalysis (Storck et al., 1998). The physisorption method as

routine tool is the most widely used technique for the characterization of micropores (diameter

< 2 nm), mesopores (diameter 2–50 nm) and macropores (diameter> 50 nm) of new materials

for catalysts design. Gas-solid adsorption has been carried out extensively in the past few years

due to the comparative ease of experimental work (Shanavas et al., 2011). The results of

physisorption experiments are usually presented in graphical forms as isotherms. The

Langmuir equation is the simplest equation for the gas-solid adsorption systems upon which

other isotherms were derived. The earliest reports on gas-solid adsorption were classified by

Brunauer-Emmett-Teller (BET) based on their derived models type I, II, III, IV and V

49
isotherms (Figure 2.9), although Freundlich isotherm was reported as the oldest (Iyer & Kunji,

1992).

The Langmuir model and the BET isotherms are used to obtain the monolayer values from

which the specific surface areas are calculated. The monolayer values and the molecular areas

estimated by Langmuir and BET isotherms are inaccurate, resulting to errors in surface areas

estimations. BET admitted that their equation be strictly applied in the range of relative

pressures between 0.05 and 0.35 for adsorption process, since the equation did not fit the

experimental data for all relative pressures (De Boer et al., 1966). The Langmuir equation has

only limited applicability, for it is valid for type I isotherm of Brunauer's classification.

Figure 2.9: Adsorption isotherm classified by Brunauer (Adapted from Storck et al., 1998)

50
2.14.1 Langmuir model equation (linearized)

The linear form of the Langmuir equation is described as:

P P 1
 
V Vm bVm (2.1)

Where: 'b' is the Langmuir constant related to the affinity of binding sites, P, Po, and Vm are

P
as defined earlier. A plot of versus P yields a straight line from Langmuir Vm and b can be
V

evaluated. The quantity Vm obtained from the Langmuir equation is the monolayer value, which

is used to determine the Langmuir specific surface area (SL) of the adsorbent by using Equation

(2.2) (Shanavas et al., 2011).

𝑆𝐿 = 𝑉𝑚 × 𝑁 × 𝐴𝑚 (2.2)

Where: N is the Avogadro's number; and Am is the adsorbate molecular area (0.162 nm2 for

nitrogen).

The Langmuir equation derived from the adsorption on the nonporous surface adsorption, gives

the type I isotherm. In physical adsorption, however, the type I isotherm is obtained from

microporous adsorbents such as active carbons or zeolites. In such microporous adsorbents, the

type I isotherm can be explained by filling of the adsorbate molecules into micropores observed

as initial sharp rise of the isotherm. The flat region of the isotherm appears when the micropores

are completely filled by adsorbate. It is reasonable to conclude that the monolayer volume (Vm)

of microporous adsorbent obtained from the Langmuir plot corresponds to the microspore

volume. The surface area of microporous adsorbent estimated from Vim occasionally gives the

abnormally high value.

2.14.2 BET model equation

The assumptions of the BET equation are:

 Estimates monolayer capacity;

51
 Adsorption occurs layer by layer;

 Molecules in the first layer interact with the solid, and molecules in the following

layers behave as in bulk liquid;

 Heat of adsorption for the first layer is higher than successive layers;

 Heat of adsorption for second and successive layers equals the heat of liquefaction;

 Lateral interactions between adsorbed molecules are ignored.

x  C 1 1
   x 
V 1  x   Vm C  Vm C
(2.3)

where: 'x' is the relative pressure (P/Po); P is the partial pressure; Po is the saturation pressure;

Vm is the maximum amount of gas adsorbed/unit mass of solid at high pressure conditions

required to form a complete monolayer over the entire surface of the solid, and C is the BET

x
constant related to the heat of adsorption. A plot of versus x yields a straight line
V 1  x 

from which Vm and C are calculated. The volume of the monolayer (Vm) is converted to a

specific surface area (SBET) by assuming that each molecule in the monolayer occupies a given

surface (0.162 nm2 for nitrogen).

2.14.3 Freundlich model equation

The Freundlich isotherm represents the isothermal variation of adsorption of a quantity of gas

adsorbed by unit mass of solid adsorbent with pressure. The isotherm is obeyed when the

adsorbate forms a monomolecular layer on the surface of the adsorbent. It was determined

experimentally that the extent of adsorption varies directly with pressure until saturation

pressure Po is reached. Beyond that point, the rate of adsorption saturates even after applying

higher pressure. The Freundlich isotherm indicates that:

x
 At low pressure, the extent of adsorption varies linearly with pressure,  P1 ;
m

52
x
 At high pressure, it becomes independent of pressure,  P0
m

Thus, Freundlich adsorption isotherm fails at higher pressure.

Mathematically, the model equation is expressed as:

1
x
 KF P n (2.4)
m

Equation (2.4) in linearized form becomes:

x 1
log   log K F  log P (2.5)
m n

where, x = mass of adsorbate

m = mass of adsorbent

P = equilibrium pressure of adsorbate

KF and n are constants for a given adsorbate and adsorbent at a particular temperature. The value

of n is always greater than 1; indicating that the amount of gas adsorbed does not increase as

rapidly as the pressure.

2.14.4 The t-plot

Various thickness equations have been developed and reported in literature to relate thickness

of adsorbed film to pressure in gas-solid adsorption. The characteristics of these equations are:

 Compares an isotherm of a mesoporous materials with a standard type II isotherm of a

nonporous adsorbent with BET C constants similar to that of the microporous sample;

 Existence of a direct relation between thickness and specific volume of gas adsorbed;

 Thickness equation predicts V at various pressures for nonporous materials;

 Plotting experimental data against thickness equation data yields a straight line for

nonporous solids, and a deviated line for porous solids.

The t-values are in practice calculated with a thickness equation that describes the particular

standard reference curve


53
Practically, an adsorbate has never been reported to cover an adsorbent with a film of uniform

thickness. In many instances, it is assumed that the film thickness on pore walls is uniform

which then allows for the calculation of statistical thickness (t) from gas adsorption isotherms.

Several model equations have been reported in the literature developed to calculate statistical

thickness. The t-plot method permits the determination of micropore volume, surface area, and

help to obtain in principle information about the average pore size of the adsorbent.

De Boer et al. (1966) in their study of gas-solid adsorption proposed a master t-curve

multimolecular N2-adsorption to overcome limitations of Langmuir and the BET isotherms.

Their reports indicated an identical multilayer adsorption curves for variety of adsorbents

considered; provided no capillary condensation occur and no narrow pores are excluded during

the adsorption process. A master t-curve in terms of an average thickness of the adsorbed layer

for the adsorption of nitrogen at 78 K over the range 0.1–0.75 P/Po was developed. This

developed curve gave the same surface area with a fitted BET isotherm. If the statistical

V 
thickness (t) of one layer is 3.5 Å, then a value of 3.54 (Å) indicates an adsorption of n   
 Vm 

layers where V is the adsorbed volume, and Vm the monolayer capacity both in cm3/g STP

adsorbent.

2.14.4.1 Harkins-Jura model

Harkin-Jura model that was derived from an empirical equation of state for condensed films,

attempts to modify earlier models of Langmuir and BET. The model proposes that the

adsorptive film is of the condensed type (liquid) and hence its surface pressure, the difference

of the surface tension of the bare surface and film-covered surface is a linear function of the

area per molecule (Shanavas et al., 2011). The isotherm does not rely on parameters such as

molecular area and monolayer values, and thus gives reliable results compared to Langmuir

isotherm. The Harkins-Jura model is given as:


54
A
log x  B  (2.6)
Va 2
P
where x  , A and B are constants, Va is the adsorbed volume.
Po

A similar thickness equation was obtained by De Boer et al. (1966), which represents nitrogen

adsorption on solid adsorbents at 77 K:

13.99
log x  0.034  (2.7)
t2

And,

0.5
 
t Å   
13.99
 (2.8)
 0.034  logP / Po 

For carbon-like derived materials, a developed thickness equation is given as:

t (Å)  0.88P/Po   6.45( P / Po )  2.98 (2.9)

2.14.4.2 Halsey equation

The Halsey equation provides another alternative to calculate the statistical thickness. The

equation for nitrogen adsorption at 77 K was described as:

1/ 3
 5 
t Å   3.54  (2.10)
 lnP / Po 

2.14.5 Polanyi's potential theory

The Polanyi's potential theory occupies a unique position in the study of physisorption. For

different number of adsorbents, the theory has been successfully used to correlate data at

different temperatures and for different adsorbates, although it is reported not consistent with

monolayer adsorption on heterogeneous surfaces (Harris, 1969). The theory explained that the

adsorption process is facilitated by the existence of potential fields around the surface of the

55
adsorbent, and the adsorptive potential energy (En) needed to energize the adsorbate from its

equilibrium pressure to the saturation pressure is described by (Polany et al., 1970; Harris,

1969):

RT  P 
dP  RT ln sat 
Psat Psat
En   VdP   (2.11)
Pequil Pequil P P 
 equil 

2.14.6 Dubinin-Radushkevich equation

Dubinin-Radushkevich method, which is based on the Polanyi's potential theory, is widely used

to describe adsorption in microporous solids such as activated carbon and zeolites. The

equation has a semi-empirical origin and is based on the assumptions of a change in the

potential energy between the gas and adsorbed phases and a characteristic energy of a given

solid. This equation yields a macroscopic behaviour of adsorption loading for a given pressure

(Nguyen & Do, 2001). The proposed equation allows the calculation of micropore volume from

the adsorption isotherm.

In its basic form, the Dubinin-Radushkevich can be described as (Nguyen & Do, 2001; Misra,

1969):

V   P  
2

 exp   RT ln  / E 

 (2.12)
Vo    Po   
 

2 P 
2
 RT 
loge V  loge Vo    loge   (2.13)
 E   Po 

where, V = volume adsorbed (cm3/g); Vo = micropore volume (cm3/g). A plot of loge V vs.

loge2 p / po known as a characteristic curve plotted to test the suitability of the equation and/ or

to determine its range of applicability. If the model equations were applicable, the plot would

be a straight line with an intercept loge Vo and a slope  RT / E 2 , from which micropore

volume and characteristic energy (E) can be obtained (Nguyen & Do, 2001).

56
2.14.7 Kelvin equation

The Kelvin equation provides a correlation between pore diameter and pore condensation

pressure. The relative pressure where the pore condensation occurs depends on the pore radius.

Basic assumptions are:

 Pores of cylindrical shapes;

 No fluid-wall interactions;

 Zero (0) contact angle (θ).

The basic form of the Kelvin equation is described by (Siboni & Volpe, 2008; Mitropoulos,

2008; Aylmore, 1974):

P 2Vl Cos  1 
ln    (2.14)
Po RT  rm 

If contact angle θ = 0, then Equation (2.14) becomes:

P 2Vl  1 
ln   
RT  rm 
(2.15)
Po

where θ is the contact angle between solid and the condensed phase, and:

rP  rK  t (2.16)

When fluid-wall interactions are considered, the modified Kelvin equation for nitrogen at a

temperature of 77 K becomes:

4.15
rK (Å)  (2.17)
logPo / P 

where 𝛾 is the surface tension of liquid adsorbate (N/m), Vl is the molar volume of the bulk

liquid adsorbate (vol/mol) and the value is 34.68 cm3/mol for nitrogen at 77 K, rm is the mean

radius of curvature of the liquid/gas interface; rK is the Kelvin radius; rP is the pore radius; R

57
is the universal gas constant; T is the absolute temperature; and t is the statistical thickness of

the adsorbed film (deposited or retained).

For many practical purposes, this model is commonly a cylinder with infinite length and for

many mesoporous materials; the adsorption isotherm is of type IV. Hysteresis in capillaries

open at both ends, results from the fact that in desorption, evaporation takes place from a

hemispherical meniscus while in adsorption, condensation occurs from a cylindrical film. In

both cases, condensation or evaporation occurs at some distance t from the pore walls

(Mitropoulos, 2008). On assumption of zero contact angle, rm for a hemispherical meniscus will

be equal to rm  rK .

where rK is Kelvin radius, rK  rP  t while for a cylindrical interface, rm  2rK .

2.14.8 Thermodynamics of gas adsorption

Adsorption process is exothermic in nature, and the heats released might be the result of

energetic interactions between the adsorbate and adsorbent species. The thermodynamic

relation of gas adsorption is given by Equation (2.18).

Gads  H ads  TS ads (2.18)

where ∆Gads, ∆Hads, and ∆Sads are the Gibbs free energy change of adsorption, the enthalpy

changes of adsorption, and the entropy change of adsorption, respectively. ∆Gads is negative

because gas adsorption proceeds spontaneously. Furthermore, ∆Sads is negative because the

molecules moving at random in gas phase are fixed on the solid surface by adsorption. The

adsorbate amount and the heat released during the adsorption process could be used to

determine the thermodynamic quantities. The isosteric method is used to measure the molar

enthalpy at different temperatures and from which the heat of adsorption is calculated (van

58
Dongen & Broekhoff, 1969). The isosteric adsorption enthalpy and entropy can be determined

P 1
from the gradient and y-intercept of ln versus from the following equation:
Po T

P H ads 1 S ads
ln   (2.19)
Po R T R

The equation above is similar to the Clausius-Clapeyron relation given by:

  P
H ads   RT 2  ln 
T  Po 
(2.20)

2.15 Biodiesel and its origin

Biodiesel is a long chain fatty acid mono-alkyl esters (FAME) produced from acyl glycerol

(triglyceride) in vegetable oils and animal fats via transesterification with short chain alcohols

(in this study, methanol) (Issariyakul & Dalai, 2014). Biodiesel has good properties like

renewability, biodegradability, lower emission profile and excellent lubricity amongst others

(Sajjadi et al.,2016; Guo et al., 2010). Triglyceride (Figure 2.10) is the major component of

vegetable oils composed of three esters of fatty acid chain (acyl group) attached to the glycerol

backbone (glycerol group). Typical fatty acid chains attached to triglycerides found in

vegetable oils are usually from 10 to 24 carbon atoms. The major difference between various

vegetable oils is the type of fatty acids attached to triglyceride molecule. Rudolf Diesel (1858-

1913) invented the diesel engine in 1892 (Marchetti & Errazu, 2008) out of his desire to

improve on the steam engines of the late 1800s that were considered dangerous. In 1900, he

demonstrated the newly designed engine running on peanut oil at Paris World Exhibition

(Marchetti & Errazu, 2008). The diesel engine later became the engine of choice for power,

reliability and high fuel economy. Shortly after death of Rudolf Diesel in 1913, cheap diesel

fuel from petroleum became available. With the availability of cheap petroleum, the diesel

engine design was modified to match the properties of petro-diesel (Ma & Hanna, 1999) and

59
this discourages the use of vegetable oils in that era. Several researchers (Aransiola et al., 2014)

have reviewed the use of vegetable oil as a source of fuel. Their publications show satisfactory

performance of vegetable oils as plausible source material as fuels for use in diesel engines. In

spite of the benefits of vegetable oils as fuel in diesel engines, their high viscosity is a major

drawback when used as fuels in direct-injection diesel engines resulting in poor atomization

and engine deposits (Issariyakul & Dalai, 2014).

The World War II and the oil crises of the 1970s saw a brief interest in the use of vegetable

oils to fuel diesel engines. Unfortunately, the newer diesel engine designs could not run on

traditional vegetable oils, due to high viscosity compared to petro-diesel. In the early 1980s,

concerns over the degradation of the environment, energy security, and agricultural over

production once again brought the use of vegetable oils to the forefront. A renewed effort was

focused on the use of vegetable oils as sources of fuels and was debated at the first International

Conference on plants and vegetable oils as fuel in Fargo, North Dakota in August 1982 (Ma &

Hanna, 1999). Four major technologies adapted to reduce the viscosity of vegetable oils, such

as direct blending with petrol-diesel, micro-emulsification, pyrolysis, and transesterification

have been suggested by Subramaniam et al. (2013), Boro et al. (2012), and Schwab et al. (1987)

for use in biodiesel production. The advantages and disadvantages of these methods as listed

in Table 2.15 show that transesterification is the best choice. Currently, transesterification is

the most commonly employed method for biodiesel production. The biodiesel produced

through this process is burned directly in unmodified diesel engines, with very low deposit

formation. The idea of transesterification dated back to 1938 when it was found that the

glycerin component of the triglyceride has no calorific value and was likely to cause excessive

carbon deposit on the engine. The residue fatty acid of the oil after the removal of the glycerin

part through a catalyzed chemical reaction (transesterification) with a simple short chain

alcohol, is referred to as biodiesel, a term that made its first appearance in a paper published in

60
1988 (Issariyakul & Dalai, 2014). Over 90% biodiesel produced, is currently sourced from

edible vegetable oils that compete with food use (Sajjadi et al., 2016; Ashraful et al., 2014).

Food is a necessity of man after shelter; and the use of edible vegetable oils as biodiesel source

materials is counterproductive. Feedstock accounts for about 60–95% of the total cost of

biodiesel (Banković-Ilić et al., 2012) and the production of biodiesel from low-cost materials

will reduce its price and makes it more competitive with fossil-diesel. Non-edible rubber seed

oil can be exploited as low-cost industrial oil for biodiesel production and this will reduce

biodiesel cost and avoid food-fuel crisis in future (Morshed et al., 2011). Homogeneous base

catalysts such as NaOH, KOH and methoxides of sodium and potassium are most suitable for

industrial transesterification of edible vegetable oils that are generally low in free fatty acids

(FFAs). However, associated problems such as difficulty in catalyst-products separation, initial

catalyst cost, soap formation, low-grade by-product glycerol, reactor corrosion and large

volume washing water requirements induced further research for environmentally benign

heterogeneous catalyst for biodiesel synthesis from non-edible oils (Roschat et al., 2012; Zabeti

et al., 2009). Heterogeneous (solid) catalysts are separated from reaction mixtures by simple

filtration process. They are reused and recycled severally producing high-grade glycerol

(Uprety et al., 2016). In the recent past, successful ventures were reported on the utilization of

calcined wastes such as eggshell (Piker et al., 2016) and animal bones (Nisar et al., 2017) to

generate CaO as low-cost solid base catalysts for transesterification of oil to biodiesels with

yields above 96%. Betiku and Ajala (2014) also used a solid catalyst obtained from waste

plantain peels calcined at 500 oC for 3.5 h as direct basic heterogeneous catalyst to transesterify

pretreated yellow oleander (Thevetia peruviana) oil to biodiesel with a yield > 95% (w/w) at

RSM optimized conditions of methanol/oil ratio (0.3 v/v), reaction time (1.5 h), and calcined

plantain peels catalyst (3.0 w/v).

61
Table 2.15: Methods of viscosity reduction in vegetable oils for biodiesel production

(Adapted from Boro et al., 2012)

Methods Definition Advantages Disadvantages

Pyrolysis or thermal Method of conversion of one 1.Lower processing Energy intensive


cracking substance into another by cost, compatibility
application of heat with the aid with infrastructure,
of the catalyst in the absence of engines and fuel
air or oxygen standards, and feed
stock flexibility
2.The final products
are similar to
diesel fuel in
composition

Micro-emulsions A micro-emulsion is a colloidal 1. Fuel viscosity is Lower cetane


equilibrium dispersion of lowered number and energy
optically isotropic fluid 2. They can improve content
microstructures with spray
dimensions generally in the characteristics by
range of 1–150 and explosive
spontaneously formed from two vaporization of the
normally immiscible liquids and low boiling
one or more ionic or non-ionic constituents in the
amphiphiles. micelles

Direct use and Either use vegetable oil directly Liquid nature and 1. Higher viscosity
blending or is blended with diesel portability. Heat 2. Lower volatility
content ~80% of 3. The reactivity of
diesel fuel) readily unsaturated
available; hydrocarbon
renewability chains

Transesterification Transesterification (also called Renewability; higher Glycerol disposal


alcoholysis) is the reaction of a cetane number; and waste water
fat or oil with an alcohol to lower emissions; problem
form esters and glycerol higher combustion
efficiency

62
Figure 2.10: A typical structure of a triglyceride molecule

2.15.1 Biodiesel feedstock

Lipids and simple alcohols are the main feedstock required for the production of biodiesel.

There are over 350 oilseed crops identified worldwide as sources of lipids suitable for biofuels

(Ashraful et al., 2014). Lipids can also include animal fats, and more recently, micro-algae and

cyanobacteria (Kings et al., 2017). Over 95% of biodiesel are sourced from edible vegetable

oils that are regionally based: rapeseed and sunflower in European countries and Canada;

soybean oil in United States; palm oil in tropical countries such as Indonesia and Malaysia;

coconut oil in the Philippines (Bankovic-Ilic et al., 2012; Ashraful et al., 2014). Several

researchers have reported that the cost of biodiesel depends on the feedstock that accounts for

about 60-95% (Ashraful et al; 2014; Morshed et al., 2011). The debate over the food versus

fuel crisis on the use of edible oils for biofuel production simulated research into the use of

non-edible and used oils as a veritable source of raw material for biodiesel production (Onoji

et al., 2016a).

Oil seed bearing plants such as rubber tree, jatropha curcas, and karanja just to mention a few

have been described as promising sources of non-edible oil for biodiesel production to avert

food shortage. The rubber tree with a high seed oil yield (35–60%) and with the potential to

generate natural rubber for production of various rubber products placed it at an advantageous

position over other non-edible oilseeds (Onoji et al., 2016a; Atabani et al., 2013). The rubber

63
is the bioenergy crop of the next century considering its versatile use in biofuels and automobile

industry.

The most frequently used alcohols for biodiesel production are short-chain alcohols, such as

methanol (CH3OH), ethanol (C2H5OH) and butanol (C4H9OH). The choice of alcohol depends

on cost, sustainability, and performance characteristics (Encinar et al., 2007). In Brazil, for

example, ethanol is preferred because it is readily available, cheap and sustainable. In

transesterification with ethanol, unreacted ethanol forms an azeotrope that is expensive to

purify during recovery. Methanol is much cheaper than ethanol and the literature on biodiesel

is replete with methyl esters rather than ethyl esters (Yuan et al., 2008). Methanol is more easily

recycled because it does not form an azeotrope. Ethanolysis forms stable emulsions depicting

difficulty in product separation and this is contrary to unstable emulsions resulting from

methanolysis. The reaction rate is relatively slow with ethanol because the formation of

ethoxide anion is more difficult than that of the methoxide (OM Tapanes et al., 2008).

2.16 Catalysis in biodiesel production

Esterification and transesterification reactions are commonly employed in biodiesel synthesis.

Vegetable oils that have high content of FFAs will undergo pretreatment process (esterification

reactions) to reduce the levels of FFAs to an acceptable limit (<1%) prior to transesterification

process with base catalysts.

2.16.1 Acid-catalyzed esterification reaction

This method is suitable for processing biodiesel feedstock with high FFAs and other low-grade

oils that are not suitable for direct base transesterification. The catalysts used include sulphuric,

hydrochloric, sulfonic and phosphoric acids (Aransiola et al., 2014). However, the downstream

processing of huge waste water, high oil/methanol ratio, corrosiveness, catalyst recovery

64
problem, longer reaction time, and modulate temperatures and pressures are some of its

disadvantages. The esterification reaction is a simple reversible reaction as shown in Equation

(2.21) (Guo & Fang, 2011). FFA supplies hydroxide and the methanol supplies proton without

intermediate process and FFA is converted to ester (biodiesel) and water as a byproduct.

(2.21)

2.16.2 Transesterification of oils using homogeneous catalyst

Transesterification has been widely accepted as the most viable method to produce biodiesel.

Transesterification is the chemical reaction of vegetable oils, waste oils, animal fats, or

microalgae oils with excess primary alcohol in the presence of a suitable catalyst. The choice

of catalyst and biodiesel production route depends on the physicochemical properties of the

selected oil. The catalysts used are classified as: homogeneous (acid and base), heterogeneous

(acid and base), and enzymes. The triglyceride is converted stepwise to diglyceride and

monoglyceride intermediates, and finally to glycerol with biodiesel (FAME) formed at each

step (Issariyakul & Dalai, 2014; Thanh et al., 2012; Guo & Fang, 2011) as depicted in Figure

2.11. The breakdown of the triglyceride structure requires three steps:

1. The first step produces an intermediate tetrahedral;

2. The second step involves the breakdown of the unstable intermediate tetrahedral into

diglyceride ion and fatty acid ester (biodiesel), and;

3. The final step that involves the recovery of the catalyst by proton transfer.

The three mechanisms are repeated for cleavage of each fatty acid ester and then finally, three

fatty acid esters (biodiesels) and a glycerol is formed as shown in the overall Equation (2.22)

(Thanh et al., 2012).

65
Figure 2.11: Scheme for stepwise transesterification of triglyceride to methyl esters

(Adapted from Guo & Fang, 2011)

(2.22)

2.16.2.1 Alkali-catalyzed transesterification reaction

Hydroxides and alkoxides of alkaline metals as well as the carbonates of sodium and potassium

metals catalyze these reactions. The process is widely used in the laboratory, pilot and industrial

scale levels. The catalysts show a very high performance when the feedstock have low content

of FFAs such as edible vegetable oils of rapeseed, soybean, sunflower etc. However, the

shortcomings of this method such as energy intensive, glycerol recovery problems,

corrosiveness, and food-fuel crisis, have generated debates about their use in different form in

the past few years.

66
2.16.2.2 Two-step transesterification reaction

The method involves acid-catalyzed esterification followed by alkali-catalyzed

transesterification, and useful for processing high FFAs feedstock to biodiesel. It is a

combination of the process as described in Section (2.16.1), followed by a second step, alkali-

catalyzed transesterification, as described in Section (2.16.2.1) (Betiku & Ajala, 2014).

2.16.2.3 Mechanisms for homogeneous catalysis

Homogeneous catalysts used for oil transesterification are either acid or base, and usually in

liquid form. The protonation of the carbonyl group in the triglyceride molecule and the alcohol

attacking the protonated carbon to create a tetrahedral intermediate form the basis of acid

catalysis (Aransiola et al., 2014). However, in a homogeneous-base catalyzed reaction, the

creation of nucleophilic alkoxides from the alcohol to attack the electrophilic part of the

carbonyl group of the triglycerides forms the basis of the reaction. The mechanisms for acid-

catalyzed and base-catalyzed are well documented in the literature (Thanh et al., 2012). These

reactions demonstrate the conversion of triglyceride into diglyceride. The reaction mechanisms

of diglyceride and monoglyceride, which convert into monoglyceride and glycerol,

respectively, take place in the same way as for triglyceride. The overall reactions were earlier

shown in Equation (2.22).

2.16.3 Transesterification of oils using heterogeneous catalysts

Solid catalysts of acid and base origins are known as heterogeneous catalysts. They were

developed to overcome the limitations of homogeneous catalysts, although they have similar

mechanisms to the homogeneous type. Aransiola et al. (2014) report various types of

heterogeneous catalysts and their preparatory methods in their review paper.

67
2.16.3.1 Heterogeneous acid transesterification reaction

Heterogeneous acid catalysts have ability to catalyze both esterification and transesterification

reactions simultaneously, and are less corrosive, less toxic, and generate fewer environmental

problems (Zhang et al., 2011), and thus produce a low-cost biodiesel that is competitive with

commercial diesel. The acids contain variety of acid sites with different strengths of Brӧnsted

or Lewis acidity as compared to homogeneous acid catalysts despite their lower activity

(Aransiola et al., 2014; Thanh et al., 2012). At low temperature, the activity of acid catalysts

in transesterification is normally quite low, and to obtain a sufficient reaction rate, it is

necessary to increase the reaction temperature (> 170 oC) and this could increase the capital

costs thus the cost of biodiesel (Di Serio et al., 2008). Several solid acid-catalysts such as

Nafion-NR50, acid ion-exchange resins, sulfated zirconia (SZ), tungstated zirconia (WZ),

supported acid catalysts (with metal oxides as supports), zeolite-based catalyst (e.g., zeolite β,

HYzeolite) and heteropoly acids have been reported in the literature for the transesterification

of oil to biodiesel with good yields.

The mechanism of solid acid-catalyzed esterification reaction consists of the following steps:

Firstly, solid catalysts provide protons, and carbonyl carbon is protonated. This is followed by

nucleophilic attack of CH3OH on the carbonium ion to form a tetrahedral intermediate. The

last stage involves the migration and reformation of proton with the breakdown of intermediate

to produce biodiesel (Guo & Fang, 2011). As discussed in Section (2.16.2), three consecutive

reactions are required to complete the transesterification of a triglyceride molecule. The

mechanisms of solid acid-catalyzed transesterification of triglycerides show that triglycerides

are protonated at the carbonyl group on the surface of the catalyst. A nucleophilic attack by the

alcohol forms an unstable tetrahedral intermediate that leads to proton migration, followed by

breakdown of the tetrahedral intermediate with the assistance of solvent. After twice

repetitions, three new esters products are produced and the catalyst regenerated. During the

68
catalytic process, protonation of carbonyl group boosts the catalytic effect of solid-acid

catalysts by increasing the electrophilicity of the adjacent carbonyl carbon atom (Guo & Fang,

2011). Issariyakul & Dalai (2014) presented the mechanisms of solid acid-catalyzed reactions

for both esterification and transesterification in a more compacted form. In the esterification

process, free fatty acid initially adsorbs on the active acid site of the catalyst that leads to

carbonation during their interaction. This is followed by an attack of methanol to produce an

unstable tetrahedral intermediate that finally leads to the removal of water molecule to produce

methyl ester (biodiesel). The reaction mechanism for the transesterification part occurs in a

similar manner to that of esterification. The formation of methyl ester stems from an

elimination of diglyceride, monoglyceride, and glycerol from the tetrahedral intermediate when

triglyceride, diglyceride, and monoglyceride are adsorbed in the acidic site, respectively.

2.16.3.2 Heterogeneous base transesterification reaction

The use of heterogeneous base catalyst has received a wider research attention over the last

decade. Such catalysts are known to transesterify low-quality oils with moderately high FFAs

and water contents. Though low in catalytic activity, their performance could be improved by

simple calcinations process. Heterogeneous base catalysts are available in different forms such

as hydrotalcites, metal oxides, mixed-metal oxides, metallic salts, supported-based catalyst and

base zeolites to mention a few (Guo & Fang, 2011). They can be generated from low-cost waste

materials such as eggshell (Piker et al., 2016), plantain peels (Betiku & Ajala, 2014) and animal

bones (Nisar et al., 2017) with their catalytic activity enhanced by calcinations. Hydrotalcites

(HTs) are a class of anionic and basic clays known as layered double hydroxides (LDHs) with

the formula Mg6Al2(OH)16CO3.4H2O. HTs consist of positively charged brucite-like layers and

interstitial layers formed by CO32 anions and water molecules compensate the positive charge

resulting from the substitution. LDHs have strong alkali sites and high stability with good

69
adjustability of composition and structure (Guo & Fang, 2011). Metal oxides are composed of

cations possessing Lewis acid and anions with Brӧnsted base. They are classified as single

metal oxides (e.g., CaO, MgO and SrO) and mixed metal oxides (e.g., A-B-O type metal

oxides, where A is alkaline-earth metal (Ca, Ba, Mg), alkaline metal (Li) or rare earth metal

(La) and B is a transition metal (Ti, Mn, Fe, Zr, Ce) (Kawashima et al., 2008; Liu et al., 2008).

Reports on the use of inorganic solid bases (metallic salts) are rare in literature, although such

materials are easy-to-use and could be sourced from low-cost materials. Available reports

indicate that sodium silicate, vanadyl phosphate, calcium zincate and calcium methoxide have

been investigated as heterogeneous catalysts for biodiesel synthesis with yield of almost 100%

under certain conditions (Guo & Fang, 2011).

Alkali metals (Li, Na, K) and alkaline-earth metals (Mg, Ca, Ba) are the most common sources

of super basicity, and selected as the active species of supported solid base catalysts for

biodiesel synthesis. They are frequently used in the metallic form or as various ionic forms of

hydroxide, halide, carbonate and nitrate. Alumina, silica, zinc oxide, zirconium oxide and

zeolite have been used as supports for these catalysts (Issariyakul & Dalai, 2014; Guo & Fang,

2011) in their preparation by impregnation followed by calcinations. The reaction mechanism

pathway for solid base-catalyzed transesterification seems to follow a similar mechanism to

that of a homogeneous base catalyst. First, ion exchange proceeded after methanol adsorbed

on the surface of solid base, producing active specie (CH3O-) which is strongly basic and highly

active. Secondly, nucleophilic attack of CH3O- on the carbonyl carbon of triglyceride formed

a tetrahedral intermediate. Thirdly, rearrangement of the intermediate results in the formation

of ester (biodiesel). Finally, protons are converted to diglyceride ion to generate diglyceride.

This sequence is then repeated twice to yield glycerol and biodiesel.

70
2.16.4 Enzyme-catalyzed transesterification reaction

Lipases as enzyme-catalysts for biodiesel synthesis have been classified as environmentally

friendly approach to solving energy problems. This new area of research seems to overcome

the challenges of using conventional catalysts-feedstock pretreatment, high energy

requirements, catalyst removal problems and waste water treatment (Aransiola et al., 2014).

Lipases as biocatalysts can be classified as: extracellular lipases and intracellular lipases, and

are produced by microorganisms, animals and plants (Gog et al., 2012; Antczak et al., 2009).

The main drawbacks of lipase-catalyzed transesterification reaction are the high cost of enzyme

production, higher reaction times as compared to base-catalyzed reaction, potential hazard of

explosion, regeneration and reuse of biocatalyst limited with a long operating time. However,

in order to reduce cost of operation and improve catalyst stability and reusability, lipases can

be immobilized on several materials (Modi et al., 2007; Noureddini et al., 2005), and have been

characterized by high selectivity and efficiency, absence of side reactions, and high yield of

biodiesel (Koh et al., 2011; Dizge & Keskinler, 2008).

2.16.5 Heterogeneous catalyst from natural sources

Sourcing for sustainable heterogeneous catalysts derived from waste biomass materials and

other natural sources as alternative to conventional heterogeneous catalysts of chemical origin

is a promising area of research (Aransiola et al., 2014). The conventional heterogeneous

catalysts (acid or base) require high synthesis cost, and sometimes-expert skills are required

for processing them. Therefore, sourcing for solid base catalysts from biomass could reduce

the cost of biodiesel and serve as a platform for converting waste to wealth. Betiku and Ajala

(2014) obtained 95% yield of yellow oleander oil biodiesel with calcined biomass (waste

plantain peels) as solid base catalyst. In addition, the catalytic activity of CaO obtained from

natural and biological waste materials such as eggshell, limestone calcite, dolomite and animal

71
bones through calcinations has been investigated as solid base catalysts for oil

transesterification (Nisar et al., 2017; Piker et al., 2016).

2.17 Transesterification of rubber seed oil

Transesterification reaction is a process generally employed to produce fatty acid mono-alkyl

ester (FAME) popularly called biodiesel when vegetable oil or animal fats reacts with excess

short-chain alcohol (in this case, methanol) in the presence of a suitable catalyst (acid, base or

enzymes) with glycerol as by-product (Onoji et al., 2016a). The free fatty acids (FFAs) and

moisture contents of the oil/fats determine the choice of the catalyst. The transesterification

processes available in literature for rubber seed oil biodiesel production are conventional

method, in-situ, enzymatic, supercritical, co-solvent, ultrasonic-assisted, and microwave-

assisted. Several factors such reaction time, temperature, FFAs and moisture contents, pressure,

capital cost, catalyst type and reagents determine the applicable method of production. Edible

vegetable oils with low FFA (<1 wt. %), commonly employ homogeneous base catalysts such

as NaOH and KOH. Homogeneous base catalyst will form soap during transesterification of

rubber seed oil with simple alcohols because of its high FFA. In a 2-step process, esterification

reaction of rubber seed oil with excess alcohol in the presence of homogeneous acid catalyst

(H2SO4) will reduce the FFA content. The esterified oil is transesterified with homogeneous

base catalyst to produce a higher yield of rubber seed oil's biodiesel and glycerol as waste.

However, the downstream processing of the biodiesel to recover the catalyst and other waste

products possess challenges to industrialists with the attendant high capital costs.

Heterogeneous catalyst that catalyzes both esterification and transesterification reactions is

seen as a better option for high FFA oils. They are stable, active at low temperature, have high

selectivity, relatively cheap and could be locally sourced (Di Serio et al., 2008).

72
2.18 Transesterification methods for rubber seed oil

The transesterification methods available for biodiesel production from RSO are conventional,

in-situ, enzymatic, supercritical, co-solvent, ultrasonic-assisted and microwave-assisted.

2.18.1 The conventional method

The common conventional homogenous base catalysts for RSO transesterification reaction are

NaOH, KOH, carbonates and the alkoxides of Na and K (Ahmad et al., 2014a; Ibrahim & Pillai,

2011) with biodiesel yields ranging from 86.7-97.8 wt.%. Homogeneous acid catalysts such as

H2SO4, HCl, and H3PO4 are used for RSO esterification (pre-treatment) process prior to

transesterification reaction (Ahmad et al., 2014b, Ahmad et al., 2014c; Bokhari et al., 2014).

Heterogeneous catalysts such as cation exchange resin, solid metal oxides and their silicates

are tolerant to high FFAs RSO and can be used several times without much loss to activity

(Omorogbe et al., 2013). Gimbun et al. (2012) used lime-based catalyst derived from cement

clinker to obtain 96.9 wt. % yield of biodiesel from high FFA RSO. Krishnakumar et al. (2013)

synthesized biodiesel from rubber seed oil with solid CaO catalyst calcined at 850 oC for 1.5 h

and reported a yield of 98 wt. %.

2.18.2 The in-situ method

This involves contacting the raw-milled seed directly with the alcohol in the presence of a

catalyst (acid/alkali) instead of the pre-extracted oil. The advantages of the process include

reduced reaction time, higher yield, smaller wastewater formation and lower consumption of

reagents (Ma & Hanna, 1999). Abdulkadir et al. (2014) and Widayat et al. (2013) employed

in-situ method and obtained optimum biodiesel yields of 75 wt. % and 91.05 wt. percentage

respectively during RSO transesterification.

73
2.18.3 The enzymatic method

This method uses enzymes such as lipases as biocatalysts. Though characterized by low

reaction rates, when the enzymes are immobilized, they can be re-used several times (Ma &

Hanna, 1999). The process is FFA friendly and the milder conditions allowed biodiesel yield

of at least 90% (Awolu & Layokun, 2013) and can be used to transesterify rubber seed oil to

biodiesel.

2.18.4 The supercritical method

This process is tolerant to high FFA and moisture content and recommended by researchers for

variety of feeds such as rubber seed oil and waste vegetable oils (Sanjel et al., 2014). Under

the supercritical conditions, the alcohol becomes “super acid” thereby eliminating the use of

catalyst creating a single phase with low reaction time (2-4 min) and low mass transfer

limitations are reduced. The high temperature (300–400 oC), high pressure (80–100 atm) and

high molar alcohol/oil ratio (42:1) are some of its inherent disadvantages (Thanh et al., 2012).

2.18.5 The co-solvent assisted method

This is a new greener biotechnological method because of minimum waste generation and low

energy demand. The co-solvent creates a single phase that gives rise to increased reaction rates

within a short time. Among the researched co-solvents such as tetrahydrofuran (THF), 1-4-

dioxane, acetone and di-ethyl ether, THF was the first reported co-solvent used for biodiesel

production (Thanh et al., 2012).

2.18.6 The ultrasonic-assisted method

The immiscibility of alcohol and vegetable oils hinders mass transfer rates between the phases.

This limits the reaction rates and the yield of biodiesel (Guo & Fang, 2011). Widayat and Kiono

74
(2012) employed ultrasonic-assisted technology in the extraction and transesterification of

rubber seed oil simultaneously to biodiesel at 40 oC and 40 Hz. A developed model equation

for the transesterification step based on response surface methodology shows that ultrasound

could reduce process time and increase biodiesel yield significantly.

2.18.7 The microwave-assisted method

This is an alternative heating method to the conventional method of heating because the

microwave energy selectively energizes the catalyst's interaction with the reactants (Hincapie

et al., 2014; Di Serio et al., 2008). Microwave is non-ionizing radiations with low energy. The

changing electrical field component of microwave radiation interacts with the dipoles of the

reactants molecules increasing their rotations. This reflects an increase in reaction rates and

higher biodiesel yield in shorter time, as compared to conventional systems heating (Hincapie

et al., 2014; Guo & Fang, 2011).

2.19 Factors affecting the transesterification of rubber seed oil

Transesterification of rubber seed oil to synthesize biodiesel is affected by number of factors.

The parameters considered in this thesis are as follows:

2.19.1 FFA content of rubber seed oil

The FFA content of rubber seed oil is reportedly high with the range 23.471–45 wt. % (Ahmad

et al., 2014a; Omorogbe et al., 2013; Ebewele et al., 2010b). The reduction of FFA values of

crude rubber seed oil is significant to avoid soap formation and downstream processing

problems when a homogeneous base catalyst is used in transesterification reaction (Ibrahim &

Pillai, 2011). Homogeneous acid catalyst such as H2SO4 is employed in esterification of the oil

with alcohol to reduce the FFA to less than 1 wt.% prior to transesterification. Catalyst

75
concentration followed by alcohol/oil ratio is the most influential parameters compared to

reaction temperature and time (Ahmad et al., 2014b).

2.19.2 Alcohol/oil molar ratio

Alcohol/oil molar ratio is one important variable that affects the yield of biodiesel. The

stoichiometric ratio for transesterification requires three moles of alcohol and one mole

triglyceride to yield three moles of fatty acid esters and one mole of glycerol (Ma & Hanna,

1999). The molar ratio depends on the catalyst type and may be as high as 30:1 for acid

catalyzed reaction and 6:1 for base reaction for butanol-soybean oil (Freedman et al., 1986).

Various researchers (Bokhari et al., 2014; Ahmad et al., 2014b; Abdulkadir et al., 2014)

reported a 6:1 molar alcohol/oil ratio for maximum biodiesel yield of 75–96.8 wt. % when

homogeneous catalyst such as KOH is employed for rubber seed oil transesterification. Ahmad

et al. (2014b) observed that alcohol/oil molar ratio is the most influencing factor in rubber seed

oil biodiesel production up to certain limit, beyond which the amount of methanol has no effect

on the biodiesel yield while a higher methanol concentration may result to downstream

processing problems.

2.19.3 Catalyst concentration

Different catalyst loadings for the successful transesterification of rubber seed oil to biodiesel

have been reported in the literature. Gimbun et al. (2012) reported an optimum 6 wt. % with a

yield of 92.3% using a limestone-based catalyst for rubber seed oil transesterification. Optimal

conditions such as catalyst loading, calcination temperature and time are usually determined

through optimization process (Ahmad et al., 2014a; Ahmad et al., 2014c). Catalyst loading was

reported the second most influencing factor that determines higher of biodiesel (Ahmad et al.,

2014b; Widayat & Kiono, 2012). The physical appearances of rubber seed oil biodiesel are

76
determined by the concentration of the catalyst with higher concentration giving rise to a darker

colour (Widayat et al., 2013). Dhawane et al. (2015) studied the influence of using flamboyant

pods derived steam activated carbon as catalyst (0.5–5 wt.%) on the transesterification of

rubber seed oil at 60 oC, 15:1 methanol/oil ratio, 750 rpm, and 60 min reaction time. they

reported a maximum FAME yield of 89.3% at catalyst loading of 3.5 wt.%. Further increase in

catalyst concentration decreases the yield due to soap formation during the reaction, viscous

and emulsified reaction mixture at higher catalyst loadings.

2.19.4 Reaction temperature and time

Generally, increase in temperature increases the rate of a chemical reaction and the yield of

rubber seed oil biodiesel (Abdulkadir et al., 2014; Gimbun et al., 2012). Dhawane et al. (2015)

reported 60 oC as most desirable temperature for transesterification reaction when methanol is

used. Gimbun et al. (2012) observed that increasing the temperature above the boiling point of

methanol (65 oC) to about 70 oC reduces the yield from 96.9 to 95.8 wt. % when rubber seed

oil was transesterified with derived CaO as a base catalyst. The decrease yield is explained by

the evaporation of methanol when reaction temperature is above 65 oC. Most work cited in

literature suggests that 1 h reaction time is sufficient for the completion of a transesterification

reaction with base catalyst. Solid catalyst leaching may increase may increase when a higher

reaction time is considered (Dhawane et al., 2015).

2.19.5 Agitation speed

Agitation speed plays significant role in the esterification and transesterification of vegetable

oils to esters because of mass transfer limitations imposed by the immiscibility of the reaction

mixtures. Dhawane et al. (2016) in their study reported the influence of agitation speed on the

optimization of Iron (II) doped carbonaceous catalyst for biodiesel production from rubber seed

77
oil at 12:1 methanol/oil ratio, 55 oC, and 1 h reaction time. They observed that at 500 rpm, the

yield was 94.5%, and subsequently decreases to 85.01% at 750 rpm that was attributed to the

leaching of the adsorbed Iron (II) from the surface into the reaction mixture. Further reduction

in yield to 75.34% was reported when agitation speed increases to 100 rpm.

2.20 Thermodynamics of internal combustion ignition engines

The reaction paths for complex fuel mixtures such as diesel, biodiesel or gasoline are not well

defined compared to the actual path of combustion process for simple fuels such as hydrogen

and methane. However, the first and second laws of thermodynamics can be applied to analyze

the combustion performance of internal combustion ignition engines. First law of

thermodynamics can be used to relate the end states of mixtures undergoing a combustion

process, and its application does not require the details of the process as it fails to give the best

insight into the engine's operation (Abassi et al., 2010). Energy balance on the cylinder is made

with the aim of obtaining power and efficiency using first law. Otto and Diesel cycles are the

common methods applied to analyze the performance of engines using the first law.

2.20.1 Overview of reciprocating engines

A reciprocating engine also called internal combustion ignition engine is a piston cylinder

device (Figure 2.12) that has a wide range of applications in automobiles, truck, light aircraft,

ships, and electric power generators. The piston reciprocates in the cylinder between two fixed

positions, known as the top dead center (TDC) and the bottom dead center (BDC). The TDC is

the position of the piston when it forms the smallest volume in the cylinder, while the BDC is

the position of the piston when it forms the largest volume in the cylinder (Ҫengel & Boles,

1994). The distance between the TDC and BDC is the largest distance that the piston can move

in one direction, and it is called the stroke of the engine. The diameter of the piston is called

78
the bore. The air or air-fuel mixture is drawn into the cylinder through the intake valve, and the

combustion products expelled from the cylinder through the exhaust valve. The minimum

volume formed in the cylinder when the piston is at TDC is the clearance volume, while the

volume displaced by the piston as it moves between TDC and BDC is the displacement volume.

The ratio of the maximum volume formed in the cylinder to the minimum (clearance) volume

is the compression ratio (r ) of the engine (Ҫengel & Boles, 1994):

Vmax V BDC
r  (2.23)
Vmin VTDC

Reciprocating engines are classified as either compression-ignition (CI) engines (e.g., diesel

engine) or spark-ignition (SI) engines (e.g., petrol engine) depending on how the combustion

process is initiated. They operate as either four-stroke or two-stroke internal combustion

engines depending on designs.

intake exhaust
valve valve

TDC
Bore

Stroke

BDC

Figure 2.12: Schematic diagram for a reciprocating engine cylinder (Adapted from

Ҫengel & Boles, 1994)

79
2.20.2 Air-standard assumptions
The working fluid remains a gas throughout the entire cycle in the gas power cycle.

Compression-ignition (diesel), spark-ignition (petrol), and the conventional gas turbine engines

are devices that operate on gas cycle. During the combustion process, the composition of the

working fluid changes from air and fuel to combustion products during the course of the cycle.

The working fluid closely resembles air at all times because air is predominantly nitrogen that

hardly undergoes chemical reactions in the combustion chamber. The working fluid is expelled

as exhaust gases instead of returning to its initial state at a point in the cycle. This tells us that

the working fluid does not undergo a complete thermodynamic cycle, though internal

combustion engines operate on a mechanical cycle (i.e., the piston returns to its starting position

at the end of each revolution). To simplify the analysis of actual gas power cycles, the following

air-standard assumptions are usually applied:

 The working fluid is air. It circulates continuously in a closed loop, and behaves as an

ideal gas,

 All the processes which make up the cycle are internally reversible,

 The combustion process is replaced by a heat addition process from an external source,

 A heat rejection process that restores the working fluid to its initial state replaces the

exhaust process.

The air-standard assumption is called the cold-air standard assumption if the air has constant

specific heats whose values are determined at room temperature (25 oC). This simplified model

enables us to study qualitatively the influence of major parameters on the performance of the

actual engines (Ҫengel & Boles, 1994).

80
2.20.3 Diesel cycle-The ideal cycle for compression ignition engines

The Diesel cycle (Figure 2.13) is the ideal cycle for CI reciprocating engines. The CI engine,

first proposed by Rudolf Diesel in the 1890s is very similar to the SI engine. In SI engines, the

air-fuel mixture is compressed to a temperature below the auto-ignition temperature of the fuel,

and the combustion process is initiated when a spark plug is fired (Ҫengel & Boles, 1994). In

CI engines, only air is compressed to a temperature above the auto-ignition temperature of the

fuel, and combustion starts on contact as the fuel is injected into the hot air. Therefore, a fuel

injector in diesel engines replaces the spark plug and carburetor. In gasoline engines, a mixture

of air and fuel is compressed during the compression stroke, and the compression ratio is

limited by the onset of auto-ignition or engine knock. In diesel engines, only air is compressed

during the compression stroke, thereby eliminating the possibility of auto-ignition. Therefore,

diesel engines can be designed to operate at much higher compression ratios, typically between

12 and 24 (Ҫengel & Boles, 1994).

The fuel injection process in diesel engines starts when the piston approaches TDC and

continues during the first part of the power stroke. Therefore, the combustion process in these

engines takes place over longer intervals and thus, diesel cycle is approximated as a constant-

pressure (process 2 → 3) heat addition process. This is the only process where Diesel cycle

differs from Otto cycle. The remaining three processes are the same for both ideal cycles

(processes 1 → 2, 3 → 4, and 4 → 1). Diesel cycle, like the Otto cycle, is executed in a closed

system (piston-cylinder).

81
(a)
P Qin
2 3

isentropic

isentropic Qout
1
v

T
(b)
Qin
P =constant 3
2
4
Qout
1 v =constant

Figure 2.13: P–v (a) and T–s (b) diagrams for the ideal Diesel cycle (Adapted from Ҫengel

& Boles, 1994)

Under the cold-air standard assumptions, the amount of heat added to the working fluid at

constant pressure and rejected at constant volume can be expressed as:

Qin  Q23  W23  (U ) 23  P2 v3  v2   U 3  U 2  (2.24)

Qin  h3  h2  C p T3  T2  (2.25)

Qout  Q41  W41  (U ) 41  U 4  U1  Cv T4  T1  (2.26)

W41  0 (2.27)

82
The thermal efficiency of the ideal diesel cycle under the cold-air standard assumptions

becomes:

T 
T1  4  1
T T
 4 1  1  1 
Wnet Q T
nth , Diesel   1  out
 T3  T2 
(2.28)
Qin Qin T 
T2  3  1
 T2 

Cp
where   , the specific heat ratio and equal to 1.4 for ideal gas (Azoumah et al., 2009).
Cv

If (r ) is defined as the engine's compression ratio and (rc ) represents the cutoff ratio, expressed

as the ratio of the cylinder after and before the combustion process, then Equation (2.28) can

be expressed as (Azoumah et al., 2009):

1  rc  1 
nth ,Diesel  1    (2.29)
r  1   rc  1

V1 V
where r  , and rc  3 ; V1 is the maximum volume, and V2 represents the clearance volume
V2 V2

as earlier discussed.

The combustion chamber temperature (T3 ) under ideal conditions can be evaluated from the

measured exhaust gas temperature (T4 ) by the relation (Azoumah et al., 2009):

 1
r
T3  T4   (2.30)
 rc 

The compression stroke (process 1 → 2) is assumed an isentropic process, and the following

equations used to determine the pressure and temperature at stage 2.

 1
V 
T2  T1  1  (2.31)
 V2 


V 
P2  P1  1  (2.32)
 V2 

83
Process 2 →3 ( P  constant heat addition to the air-fuel mixture, i.e. P2  P3 ):

P2V2 P3V3
 (2.33)
T2 T3

V 
T3  T2  3  (2.34)
 V2 

Process 3 → 4 (isentropic expansion of the combustion gases):

 1
V 
T4  T3  3  (2.35)
 V4 


V 
P2  P3  3  (2.36)
 V4 

Process 4 →1 (constant-volume heat rejection process with no work interactions is shown in

Equations (2.26) and (2.27), respectively.

The mean effective pressure (MEP) is determined as:

Wnet Wnet
MEP   (2.37)
Vmax  Vmin V1  V2

2.20.4 Mathematical models for internal compression ignition engines

The analysis of the performance of internal compression ignition (diesel) engines based on the

second law of thermodynamics is made possible by first-law mathematical modeling of the

various processes inside the cylinder and its subsystems. Examples of such mathematical

models include zero-dimensional models also referred to as single-zone models, two-zone,

four-zone or multi-zone models (Rakopoulos & Giakoumis, 2006). The single-zone model

assumes the working fluid in the engine to be a thermodynamic system (mixture of gases) that

undergoes energy and mass exchange with the surroundings, and the energy released is

computed using first law of thermodynamics. In the two-zone models, the working fluid in the

engine is split into two zones (burned and unburned gases) that are analyzed separately (Hago

84
& Morin, 2010; Rakopoulos & Giakoumis, 1997). Computer simulations of internal

combustion engine cycles have been of tremendous use in design studies, as diagnostic tools

in analyzing data obtained from experiments, and in understanding the complex processes

occurring in the combustion chamber. Quasi-dimensional model, a double-zone model has

been the most frequently used model because it predicts the performance of the engine better

than the single zone models (Rakopoulos & Giakoumis, 1997).

As discussed above, the Diesel cycle is useful for analyzing diesel engine’s performance. The

compression stroke of Diesel engine can be simulated from the following equations based on

ideal gas behaviour.

PV  nRT (2.38)

dp 1   R  dv R 
  1   P  
d V   Cv  d Cv 
(2.39)

The volume of the cylinder at any crank angle (θ) can be defined as (El-Seesy et al., 2017;

Heywood, 1988):

 2

V  Vc 1  rc  1   1  Cos   2  Sin 2  
 1

0.5 


(2.40)

l
 (2.41)

where Vc  clearance volume as earlier defined, rc  cutoff ratio, l  connecting rod length

(m), and   crank radius (m).

Differentiating Equation (2.40) with respect to the crank angle (θ) gives:

1 
 Vc  rc  1Sin  0.5 2Sin2 
dV
(2.42)
d 2 

The net heat release rate (HRR) in the cylinder during the combustion process using a single-

zone combustion model can be determined from the expression described by (Bayindir et al.,

2017; Ibrahim, 2016):

85
dQ  dV 1 dP
 P  V (2.43)
d   1 d   1 d

The cumulative heat release (CHR) is the integral of HRR over a restricted crank angle ( )

interval that depends on the HRR curve and given as (Guardiola et al., 2017):

   1
 dQ      1  PdV     1VdP (2.44)

Where Q  apparent heat release rate (J), P  cylinder pressure (bar), and V instantaneous

volume of cylinder (m3).

Computer software like MATLAB can be used to simulate Equation (2.38) to Equation (2.44)

and the results obtained can be compared with those from experiments.

To model the performance of a compression ignition (CI) engine according to the first law of

thermodynamics, Rakopoulous & Giakoumis (2006) suggested the following simplifying

assumptions:

(a) Spatial homogeneity of pressure for two-zone models;

(b) Spatial homogeneity of temperature for the whole cylinder or for each zone considered;

(c) Working fluid is considered an ideal gas;

(d) Gas properties (enthalpy, internal energy, etc.) are modeled using polynomial relations with

temperature (and pressure);

(e) Heat released from combustion is distributed evenly throughout the cylinder;

(f) Blow-by losses are neglected;

(g) Enthalpy associated with pressure of injected fuel is negligible;

(h) Spatially averaged, instantaneous (time resolved) heat transfer rates are used to estimate

heat transfer to the cylinder walls;

(i) Dissociation is usually, but not always neglected;

(j) No heat transfer occurs between burned and unburned zones;

(k) Work required to transfer fluid from the unburned zone to the burned zone is negligible.

86
2.20.5 Models for gas mixture properties

One of the processes that control engine power, efficiency, and emission inside the diesel

engine cylinder, is the combustion of fuel-air mixture (Heywood, 1988). The components that

make up the working fluid in the engine cylinder such as oxygen, carbon dioxide, carbon

monoxide, water vapour, fuel vapour, nitrogen, etc. are treated as ideal gases for simplification

in engine performance evaluations.

Thus,

m
PV  RT  nRT (2.45)
M

where P  pressure, V  volume, m  mass of gas, T  temperature, R  universal gas constant,

M  molecular weight, and n  number of moles.

The mixture of diesel and biodiesel properties can be determined from individual properties

using the following relations:

mix  i xi (2.46)


i 1

where mix is the property of the mixture, i is the property of pure diesel or biodiesel, and xi is

the fraction of diesel or biodiesel.

Air-fuel ratio is defined by:

 A i xi
   (2.47)
 F  mix i 1 i xi  A F i

The lower heating value (LHV) of the fuel mixture as given by Bayraktar (2005) is:

 x LHV 
i i i
LHVmix  i 1
(2.48)
 x i 1
i i

The general chemical reactions occurring in the engine cylinder for diesel/biodiesel blend-air

mixture with humid air is approximated by the following reactions (Meisami et al., 2017).

87
aC14 H 25  bC18 H 34O2  cO2  3.76N 2  xH 2O  dCO2  eCO  fO2  gNO  hNO2  iN2  jH 2O

(2.49)
where O2  3.76 N 2  xH 2O is the inlet air entering the combustion chamber.

x  7.655 (2.50)

Pv
  0.622
Pa (2.50-a)

where  represents specific humidity of air, Pv and Pa are the partial pressure of vapour and

dry air, respectively, and combustion reaction based on one mole of the fuel a  b  1 .

Özkan (2015) in a comparative study on energy and exergy analyses of a diesel engine reported

a combustion equation for diesel fuel (No. 2) based on Equation (2.49).

C14 H 25  35.264O2  3.76 N 2   14.929O2  0.017CO  14.04CO2  0.00236C14 H 25  0.4911NO 


0.0161NO2  132.343N 2  12.363H 2 O

(2.51)

2.20.6 First and second laws of thermodynamics for internal combustion engines

Energy analysis of a process is based on the first law of thermodynamics, which embodies the

principle of energy conservation. It is a traditional method used to assess the performance and

efficiency of energy systems and processes. The first law deals with the quantity of energy and

does not fully analyze the performance of internal combustion ignition engine.

According to the first law, the energy balance for steady-state control volume (combustion

chamber volume) of the engine is given as:

Q  W  H (2.52)

where Q represents the heat transfer through the combustion chamber to cooling water and

lubrication systems per mole of fuel, W represents the mechanical useful work done by the

engine in one cycle per unit mole of fuel, and H represents the enthalpy change between the

products  H p  and reactants  H r  .


 

   

88
Thus,

 
H  H p  H r (2.53)

When we substitute Equation (2.53) into Equation (2.52), we obtain an expression:

 
Q  H r  W  H p (2.54)

The enthalpy terms can be evaluated from the following equations:

  0 
H r   ni  h f   h  (2.55)
r  i

  0 
H p   ne  h f   h  (2.56)
p  e

where, ni and ne are the moles number of reactants and products, respectively, per mole of fuel

0
obtained directly from the balanced combustion equation, h f is the standard enthalpy of


formation and  h represents the difference in enthalpy between any given state and the

enthalpy of ideal gas at 25 oC and 1 atm.

Introducing Equation (2.55) and Equation (2.56) into Equation (2.54) yields:

 0   0 
Q   ni  h f   h   W   ne  h f   h  (2.57)
r  i p  e

 
where Q , W , H r , and H p are in kJ/kmol of fuel.

The lower heating value LHV  at constant pressure equals the enthalpy at complete

combustion when water is present as a vapour and based on first law, it is described by the

relation (Caton, 2000):

0 0
LHV  H P0 ,T0   ni h f ,i   ne h f ,e (2.58)
r p

89
where H P ,T represents the enthalpy of the combustion at reference dead state pressure P0 
0 0

and temperature T0  .

The second law of thermodynamics can be used to obtain a model expression for the maximum

useful work an engine can deliver, based on the assumption that an internal combustion engine

is an open system that exchanges heat and work with its surrounding. Exergy (also called

Availability) is defined as the maximum amount of useful work that can be obtained by bringing

a system to a state of thermodynamic equilibrium (thermal, mechanical, and chemical) with the

common components of the environment by means of a reversible process (Peduzzi et al.,

2016). Exergy analysis based on the second law of thermodynamics is a thermodynamic

analysis technique for systems and processes that deal with the quality of energy. It has been

increasingly applied on energy systems over the last few decades because of its advantages

over energy analysis based on first law (Abassi et al., 2010; Rakopoulos & Giakoumis, 1997).

In evaluating exergy, the characteristics of the reference environment (temperature, pressure,

and chemical composition) must be specified. Thus, exergy depends on the properties of the

system or process, and the environment. Exergy is not conserved like energy, but destroyed by

processes such as combustion, friction, mixing and throttling that decrease the ability of the

system to produce useful work (Abassi et al., 2010; Rakpoulos & Giakoumis, 1997). The

destruction of exergy, called irreversibility is the source of non-maximum conversion of fuel

energy into useful mechanical work in a diesel (and Otto) engine. The reduction of

irreversibility can lead to better engine performance through a more efficient exploitation of

fuel (Rakopoulos & Giakoumis, 2006).

From the second law of thermodynamics, the heat transfer (Q) can be expressed as follows:

Q
 S (2.59)
TA

where TA is the temperature of the surrounding and S is the entropy.

90
Combining Equation (2.52) and Equation (2.59) results in:

W  H  TA S  (2.60)

Maximum work is achieved when the pressure and temperature of the products equals that of

the atmosphere.

Under these conditions:

Wmax  G PA ,TA


(2.61)

where G P ,T represents the change in the Gibbs free energy for the conversion of reactants
A A

r  to products  p  at atmospheric temperature and pressure, and will be at its maximum at the
complete combustion of the fuel (Caton, 2000).

The availability (exergy) conversion efficiency   , which measures the effectiveness of any

internal combustion engine is determined from the ratio of the actual work delivered compared

with this maximum work, and expressed as follows:

W W
 
Wmax  G PA ,TA
(2.62)

2.20.7 Second law of thermodynamics analysis (availability)

The second law of thermodynamics analysis applies exergy (availability) equations in the

engine cylinder of internal combustion ignition engine. The different forms of availability that

exist for the analysis of the engine cylinder are chemical availability, flow availability, thermal

availability, mechanical availability and fuel availability (Caton, 2000).

The chemical exergy (fuel availability) associated with the burning of fuel in the cylinder, and

for hydrocarbon liquid fuel of the general type, C x H y on kg basis is given as (Rakopoulos &

Giakoumis, 2006):

91
 y 0.042 
a fch  LHV 1.04224  0.011925  
 x x  (2.63)

For fuel of the form C z H y O p S q (Meisami et al., 2017; Özkan, 2015; Rakopoulos & Giakoumis,

2006; Stepanov, 1995):

 y p q y 
a fch  LHV 1.041  0.1728  0.0432  0.2196 1  2.0628  (2.64)
 z z z z 

Thermo-mechanical availability is another form cited by Rakopoulos & Giakoumis (1997):

a ftm  U  P0V  T0 S  G0
(2.65)

where G0 is the working medium’s Gibbs free enthalpy at ambient pressure P0 , and

temperature T0 , calculated based on dead state compositions, and neglecting kinetic and

potential energy terms.

When the system moves from one state to another, the change in exergy as defined by Lior &

Rudy (1988) for ideal Otto cycle is also applicable for Diesel cycle.

 T   P 
a ftm  CP Tn1  Tn   Tn CP loge  n1   R loge  0 
  Tn   Pi  (2.66)

Pi  x i P (2.67)

where n (0, 1, 2, 3, 4) represents the state of the components, Pi represents the partial pressure

of component i .

If T1 represents the intake temperature, and T0 equals the temperature of the environment, then

T1  T0 for n = 0 (i.e., the environment), and the exergy for the intake process becomes:

P 
a ftm , 01   RT0 loge  0  (2.68)
 Pi 

The exergy of the exhaust gases includes thermo-mechanical exergy and chemical exergy

components described by (Meisam et al., 2017; Özkan, 2015):

92
 P
   
a ftm   yi  hi,T  hi,T0  T0 si,T  si,T0  RT0 loge  (2.69)
i  P0 

 y 
a fch  RT0  yi loge  i 
i  yi , 0  (2.70)

a fexh  a ftm  a fch


(2.71)

where a ftm represents the thermo-mechanical exergy, a fch represents the chemical exergy and

a fexh represents the total exhaust exergy in molar basis.

The availability balance can be made on the inlet manifold and the exhaust manifold for each

fuel composition and the results can be used to estimate the exergy efficiency of the engine for

different fuel composition as follows (Rakopoulos & Giakoumis, 2006):

Availability out in product Loss  destruction


  1
Availability in Input (2.72)

If we consider the cylinder as a closed system and for a four-stroke engine based on second-

law, Meisami et al. (2017), Rakopoulos & Giakoumis (2006), and Heywood (1988) describe

the exergy efficiency per cycle as:


𝑊𝑏
𝜀1 = (2.73)
𝑚̇𝑓1 𝑎𝑓𝑐ℎ

where Wb represents the brake work production, and 𝑚̇𝑓1 , represents the total mass flowrate

of fuel entering the cylinder per cycle.

The efficiency 1  can then be compared to the brake thermal efficiency 𝜂𝑡ℎ as defined by first

law of thermodynamics (Meisami et al., 2017; Rakopoulos & Giakoumis, 2006).


𝑊𝑏
𝜂𝑡ℎ = (2.74)
𝑚̇𝑓1 𝐿𝐻𝑉

Thus,
𝐿𝐻𝑉
𝜀1 = 𝜂𝑡ℎ 𝑎 (2.75)
𝑓𝑐ℎ

93
Another approach to second-law exergy efficiency for the cylinder is as defined by Alkidas

(1988):

Wb Wb
2  
Wmax Wb  I (2.76)

where I represents irreversibilities due to combustion and heat transfer.

For a closed system considered for the cylinder, mass transfer is limited in or out of the cylinder

during a single cycle, and the exergy of the fuel is described by the relation (Lior & Rudy,

1988):

a fch ,closed  a fch  P  Pi v (2.77)

where, v  specific volume (vol/mol), Pi is the partial pressure of the components (diesel and

biodiesel once vapourized) which is determined by using Raoult's law as follows:

Pi  xi P (2.78)

The specific volume v  is calculated from equations of state assuming ideal gas behaviour for

the air-fuel mixture.

RT
v (2.79)
P

where R is the ideal gas constant, T is the temperature, and P is the pressure in the cylinder.

If the assumption of zero work done by the exhaust gases as they return to the environment as

discussed earlier (Equation (2.27)) is applied, then the following changes in exergy for the

process could be written as (Lior & Rudy, 1988):

Exergy added during compression (process 1 → 2)  a2  a1 (2.80)

Exergy lost during combustion (process 2 →3)  a2  a3 (2.81)

Exergy extracted during expansion (process 3 → 4)  a3  a4 (2.82)

Exergy lost from exhaust (process 4 → 1)  a4  a1 (2.83)

94
Therefore, the effectiveness of the closed system (piston-cylinder) on the premise of only one

control mass is given below:

sys,eff 
a3  a4   a2  a1  (2.84)
ain

where:

a3  a4   a2  a1  represents the net exergy extracted for intended use, and
ain represents decrease in input fuel exergy.

2.21 Performance evaluation of diesel engines

The ruggedness, fuel economy, and high-power delivery of diesel engines has attracted its use

as engines of choice in industrial applications. The performance and efficiency of an engine is

determined by evaluating engine parameters such as brake mean effective pressure (BMEP),

brake power (BP), brake specific fuel consumption (BSFC), brake thermal efficiency (BTE %),

engine torque, and observing the exhaust gas temperature. The gaseous emissions from diesel

engines increased in the past few decades due to increase in the number of operating power

plants, vehicles, ships, light aircrafts, and agricultural tractors using diesel as fuel. Due to

increasing stringent environmental regulations to protect the environment, the use of diesel-

biodiesel blends has been encouraged to reduce the levels of exhaust gas emissions such as

carbon monoxide (CO), carbon dioxide (CO2), smoke opacity, particulate matters (PM), and

oxides of nitrogen (NOx), etc into the environment (Aldhaidhawi, et al., 2017). The high

content of oxygen (10–12 %) in biodiesel enables near 100% combustion of diesel-biodiesel

blends in diesel engines. The analyses of these parameters will determine the level of

effectiveness of the engine and compliance to environmental laws.

95
2.21.1 Brake mean effective pressure (BMEP)

BMEP is the average mean pressure in the cylinder that would produce the measured brake

power output. This pressure is calculated as the uniform pressure in cylinder as the piston rises

from top to bottom of each power stroke. The BMEP is that part of the indicated mean effective

pressure IMEP the gives the brake power.

BP
BMEP  IMEP  Losses  (2.85)
Ce  N

where Ce represents the engine capacity, N represents the number of working cycles per

rev / s
second, and for a 4-stroke engine, N  ; for a 2-stroke engine, N  rev / s .
2

2.21.2 Brake power (BP)

The brake power (also called engine shaft power) of an engine is the difference between the

indicated power (IP) and the power lost to mechanical friction (FP). The indicated power is the

actual rate of work done by the working fluid (e.g. diesel oil) on the piston measured in

kilowatts. The mechanical efficiency is the measure of the ability of the engine to overcome

the frictional power loss that includes frictions in the bearings, pistons, and other mechanical

parts of the engine (Dwivedi et al., 2013).

60
BP  IP  FP  2N s (2.86)
Te

IP  IMEP  A  l  N (2.87)

BP
 mech  (2.88)
IP

where,  mech represents the mechanical efficiency, IP is in Watts, N s represents the shaft

speed in rev/s, Te represents the effective Torque (Nm), IMEP is in Pa, A is the area of piston

(m2), l is the length of stroke (m), and N as defined in Section (2.21.1).

96
2.21.3 Brake specific fuel consumption (Bsfc)

Bsfc is defined as the rate of fuel by the engine per brake horsepower developed, and it is a

measure of the efficiency of the engine in converting the fuel supplied to produce useful work.

It is desirable to obtain a lower value of Bsfc because it indicates that the engine is more

efficient.

𝑚̇𝑓
𝐵𝑠𝑓𝑐 = (2.89)
𝐵𝑃

where 𝑚̇𝑓 is the mass flowrate of fuel.

2.21.4 Brake thermal efficiency (BTE %)

This is the ratio of the brake power to the energy generated from the engine by burning the

fuel. The energy in the fuel is usually assumed as the lower heating value LHV  .

𝐵𝑃
𝜂𝑡ℎ = (2.90)
𝑚̇𝑓 (𝐿𝐻𝑉)

2.22 Performance and emissions evaluation of diesel engine using diesel-biodiesel blends

of rubber seed oil

Few studies have examined the utilization of rubber seed oil biodiesel as an alternative engine

fuel. Onoji et al. (2016a) in their review, confirm the appropriateness of such promising fuel

for use in diesel engines to boost the economy of most sub-Saharan African countries that

depend on imported fuel. Ramadhas et al. (2005a) examined the performance of unmodified

diesel engine that run on rubber seed oil, diesel, and rubber seed oil biodiesel-diesel blends at

different loads. In all cases considered, the thermal efficiency increases with load to a certain

limit, and begins to decrease. The lower blends of biodiesel increase the thermal efficiency and

reduce the fuel consumption. The low thermal efficiency for pure biodiesel (B100) is due to

97
the reduction in calorific value of biodiesel that results to increased fuel consumption compared

to B10 (10% biodiesel-90% diesel). Pure rubber seed oil has the lowest thermal efficiency in

all cases considered, and for all loads.

The brake specific fuel consumption (Bsfc) variation with engine loads for all cases considered,

decreases with increase in the load. Lower values of Bsfc were observed for lower blends of

biodiesel/diesel, but the trend reverses for higher blends of B50-B100. This is because biodiesel

is about 14% lower in calorific value than diesel fuel. The Bsfc of rubber seed oil was higher

than those of the blends and diesel because of high viscosity and low atomization.

The carbon monoxide emission increases with load capacity in all the cases considered. This

is typical with all internal combustion engines since the air-fuel ratio decreases with increase

in load, and becomes greater than the stoichiometric value. Biodiesel blends have lower CO

emissions compared to those from diesel, biodiesel and rubber seed oil.

The profile of carbon dioxide emissions in the all cases considered by the researchers increases

with load. The CO2 emissions for the blends are lower than those reported for diesel and

biodiesel. Biodiesel combustion emits more CO2 than diesel fuel because of oxygen content

(10–12 %) of biodiesel that aids higher combustion rate.

In consideration of the variation of smoke opacity of the different fuels considered with loads,

lower values were observed for biodiesel blends compared to diesel and rubber seed oil. The

blend of B20 appears to give the lowest opacity while rubber seed oil and diesel gave similar

results.

The exhaust gas temperature increases with loads for all the cases considered by the researchers

and for all fuels tested. The exhaust gas energy loss is lower for blends up to B20 because of

lower exhaust gas temperature observed. This fact is confirmed by the higher brake thermal

efficiency and lower brake specific fuel consumption obtained for B20 and below. Pure rubber

98
seed oil gave the highest exhaust gas temperature in all loads, indicative of higher energy loss

and higher concentration of NOx in the exhaust gas stream.

The foregoing results of engine performance and emissions analyses for blends of rubber seed

oil biodiesel tested by Ramadhas et al. (2005a) are similar to those reported for biodiesel from

pongamia oil (Prabhahar et al., 2011), jatropha curcas (Elango & Senthilkumar, 2011), other

non-edible oil biodiesels (Ashraful et al., 2014), and rubber seed oil-diesel blends (Ramadhas

et al., 2005b). The NOx emissions for diesel and biodiesel blends follow an increasing trend

with loads for all types of biodiesels. NOx is generally formed at higher combustion

temperatures, and lower values are for diesel fuel compared with the blends. In all cases, pure

biodiesel (B100) generates higher concentration of NOx compared with diesel and the blends

(Elango & Senthilkumar, 2011). Geo et al. (2010) injected diethyl ether (DEE) through the

intake port during suction stroke at different flowrates of 100, 150, and 200 g/h into rubber

seed oil that was injected directly inside the cylinder at the end of compression stroke in order

to study the engine performance. Their results indicate that the brake thermal efficiency

improves from 26.5% with neat rubber seed oil to a maximum of 28.5% with DEE injection

rate of 200 g/h. Smoke opacity also reduce from 6.1 to 4 BSU along with marginal reduction

in hydrocarbons and CO emissions with DEE injection. Other researchers (Senthilkumar &

Purushothaman, 2012) examined the use of high FFA rubber seed oil biodiesel-diesel blend in

a diesel engine with respect to performance and emission analysis. Their report for B00, B5,

and B100 indicated similar trends for Bsfc, BTE, and CO emissions up to 100% load with those

reported by Ramadhas et al. (2005a).

CHAPTER 3: METHODOLOGY AND EXPERIMENTAL PROCEDURES

99
3.0 Introduction

This chapter documents all methods and techniques used to achieve results in this thesis. This

includes, but not limited to the collection of rubber seeds, seed oil extraction, characterisation,

and the determination of the physicochemical properties of the extracted oil. It also discusses

the experimental procedures and methods used in the transesterification process, and the

modelling approach used in this thesis. The oil extraction process was modeled to determine

the optimum extraction process variables. The kinetics of the oil degradation under thermal

conditions was studied. The synthesis and characterisation of catalyst from waste rubber seed

shell for transesterification of the rubber seed oil to biodiesel is reported in this chapter.

Finally, the produced biodiesel was characterised, and tested in an unmodified diesel engine to

determine its performance and emission evaluation in compliance with environmental laws.

3.1 Materials and Chemicals

Matured and fresh rubber seeds (Hevea brasiliensis) used in this work was handpicked from

different clones of NIG800 series at the plantations of Rubber Research Institute of Nigeria

(RRIN), Iyanomo, Benin City.The seeds are usually available in the early months of dry season,

usually July to August each year but could be harvested until October.

Various sizes of Duran beakers, flasks, measuring cylinders and separating funnels were used

in this work. Other glass wares used include density bottles, pipettes, burettes, etc. Muslin cloth

with Whitman filter paper was used to enclose a known weight of milled rubber seed powder

into the extraction thimble during the extraction process.

All chemical reagents used in this study were of analytical grades manufactured by BDH

Chemicals Ltd., Poole England and GFS Chemicals, Inc., 867 McKinley Ave., Columbus, OH

43223, and they include n-hexane, NaOH, KOH, CaCl2, methanol, ethanol, phenolphthalein,

100
starch, KI, CCl4, Na2S2O3, ethanoic potassium hydroxide, HCl, p-anisidine, iso-octane, diethyl

ether, and ethyl alcohol.

3.2 Experimental procedures for seed oil extraction

This work was carried out in the Biochemical Engineering Laboratory at the Department of

Chemical Engineering, Landmark University, Omu-Aran, Kwara State, Nigeria, using the

experimental procedures described in this section.

3.2.1 Rubber seed preparation

Figure 3.1 shows the flowchart for the preparation of rubber seed powder from which oil was

extracted. The rubber seeds were examined for freshness, washed 3–4 times with hot distilled

water to get rid of dust and solid impurities, and dried at room temperature for 48 h (Onoji et

al., 2017a). Separation of the chaffs from the seeds was carried out by winnowing.

Approximately 500 g of seeds were de-shelled manually using laboratory mortar and pestle to

free the kernels from the shells. The seed shells and kernels were weighed separately and their

weights recorded (see Section B.1 of Appendix B for calculations). About 100 g of seed kernel

was dried for 5 h at 105 oC using U CLEAR heating drying oven (Model DHG-9053A) to

constant weight to determine the initial moisture content of the seeds. Moisture content after

conditioning was determined by calculating the weight difference of the sample before and

after oven drying (see Section B.1 of Appendix B for calculations). The remaining seed kernels

were processed to 15% absolute moisture content required for higher oil yields before

extraction (Bhuiya et al., 2015). A portion of the processed kernel was milled and sieved into

five-particle sizes (0.5, 1, 1.5, 2 and 2.5 mm) and were used to determine which particle size

would give the maximum oil yield from the available sizes (Onoji et al., 2017a).

Raw rubber seeds

101
Sun drying

(4 to 5 days)

Manual cracking

(5 days)
Winnowing

Grinding

Rubber seed powder (for oil extraction and optimisation process)

Figure 3.1: Flowchart of rubber seed powder preparation

3.2.2 Extraction of the seed oil

The main industrial process parameters influencing oil extraction are as follows: type of

solvent, kernel particle size, seed/solvent ratio, extraction time and temperature (Elkhaleefa &

Shigidi, 2015). The solvent chosen is usually based on cost, behaviour towards the matrix, and

toxicity. In this study, five-particle sizes of rubber seed kernel (0.5, 1, 1.5, 2 and 2.5 mm) were

considered to determine the particle size that will give the maximum oil yield (Onoji et al.,

2016b). For each particle size considered, the extractor was charged with 50 g of milled rubber

seed kernel packed in a muslin cloth (Figure C.1 of Appendix C), and placed in a thimble of

102
the Soxhlet extractor. A 500-mL round-bottom glass flask was filled with 225 mL n-hexane

solvent and tightly fixed to the end of the extractor (Figure C.2 of Appendix C). Heating was

provided for the setup by a heating mantle regulated at 60 oC for 45 min. Extraction time was

considered to end when the first drop of the extracting solvent recycled back into the thimble.

Each experiment was conducted in three replicates, and the mixture was concentrated at 65 oC

using rotary evaporator (Figure C.3 of Appendix C) under vacuum to recover the extracted oil

(Figure C.4 of Appendix C) (Onoji et al., 2017a, 2016b). The oil yield was gravimetrically

determined using Equation (3.1), and average values recorded in Table D.1 of Appendix D and

presented in Figure 4.1. The particle size that gave the maximum oil yield was employed for

further oil extraction.

𝑚𝑎𝑠𝑠 𝑜𝑓 𝑒𝑥𝑡𝑟𝑎𝑐𝑡𝑒𝑑 𝑜𝑖𝑙 (𝑔)


𝑅𝑆𝑂 𝑦𝑖𝑒𝑙𝑑 (𝑤𝑡. %) = × 100 (3.1)
𝑚𝑎𝑠𝑠 𝑜𝑓 𝑟𝑢𝑏𝑏𝑒𝑟 𝑠𝑒𝑒𝑑 𝑘𝑒𝑟𝑛𝑒𝑙 𝑢𝑠𝑒𝑑 (𝑔)

3.3 Characterization of the seed oil

The physico-chemical properties of the extracted rubber seed oil were determined to ascertain

its industrial applications. Standard procedures as described by American Society for Testing

and Methods (ASTM) and Association of Official Analytical Chemists (AOAC, 1990) were

used to characterize the oil. The extracted rubber seed oil was analyzed for density, specific

gravity, oAPI, saponification value, acid value, % free fatty acids (FFA), iodine value, peroxide

value, kinematic viscosity, refractive index and pH using standard procedures as described by

ASTM and AOAC (1990).

The acid value was determined by titration method, while the iodine value was obtained by

Wijs' method. The specific gravity measurement was carried out at room temperature using

standard specific gravity bottles. The state and colour of the oil were noted using visual

inspection at room temperature (Onoji et al., 2016b). Other parameters evaluated are aniline

point, diesel index, cetane number (Mofijur et al., 2014), higher heating value (HHV)

103
(Demirbaş, 1998), mean molecular mass of fatty acids (Ajiwe et al., 1995), and average

molecular mass of oil (Fillières et al., 1995). The procedures and calculations of these

parameters are presented in Sections E.1.1 through E.1.12 of Appendix E. ASTM standard

procedures were followed to determine cloud, flash, freezing, boiling, and pour points,

respectively. The cold filter plugging point (CFPP) was estimated using the correlation cited

by Verma et al. (2016).

3.4 Spectroscopy analyses of the seed oil

The extracted oil was characterized by gas chromatography-mass spectroscopy (GC-MS),

Fourier transform infrared spectroscopy (FT-IR) and nuclear magnetic resonance (NMR)

techniques.

3.4.1 Fatty acid profile of the seed oil

The fatty acid profile was determined qualitatively using GC-MS (QP2010 Plus Shimadzu,

Japan) system equipped with flame ionization detector (FID) with capillary column DB-1

(length 30 m x diameter 0.25 mm x film thickness 0.25 μm). Helium was used as carrier gas

with a linear velocity of 49.2 cm/s and a purge flow of 3 mL/min. The column oven temperature

was programmed at 70 oC and was quickly ramped at 10 oC/min until it reached 280 oC and

held for 5 min. A sample of the oil was mixed with 1 mL HPLC grade n-hexane. A portion of

this mixture (0.5 μL) was injected into the GC at 250 oC at split ratio of 20:1. A column pressure

of 116.9 kPa and column flowrate of 1.80 mL/min was maintained. The mass spectrometer

(MS) was operated in electron ionization mode with the following parameters: ion source

temperature 200 oC, interface temperature 250 oC and solvent cut time 2.5 min. It was scanned

from m/z 30 to 350, and identification of the peaks was performed by comparing retention

times with those of National Institute of Standards and Technology (NIST) library analyzed

104
under the same conditions (Onoji et al., 2016b). The area percentage method was used to

estimate the amount of fatty acid compositions of the oil (Sanjel et al., 2014).

3.4.2 Fourier transform infrared spectroscopy

FT-IR spectroscopy can be employed to identify the various functional groups present in oils

and gives an insight into formation of primary and secondary oxidation products when the oil

is under thermal stress (Ogbu & Ajiwe, 2016). In this study, FT-IR was used to characterize

the oil at room temperature (28 oC) in order to assess the oxidation state of the oil prior to

thermal treatment using FT-IR spectrometer (Model: TENSOR 27, Bruker Optics Inc, USA)

equipped with a detector having a spectral range 4000–500 cm-1. The FT-IR was analyzed using

OPUS spectroscopy software supplied with the instrument.

3.4.3 Nuclear magnetic resonance spectroscopy

The NMR spectroscopy is a useful nondestructive tool that complements GC-MS in the

characterization of lipids and identification of compounds present. In this study, the 1H NMR

spectrum was recorded on Bruker ultra-shield TM 500-MHz NMR spectrophotometer using

deuterated chloroform (CDCl3) as solvent (δ = 7.26 ppm) containing a small amount of

tetramethylsilane (TMS) as an internal standard (δ = 0 ppm) and put in a 5-mm diameter tube.

The 13C NMR experiments were performed using the same equipment, and the carbon atom in

CDCl3 was taken as 77.42 ppm. All other peaks were assigned with respect to it. About 25–30

mg of RSO was dissolved in 1 mL of CDCl3, and the experiments were performed at 25 oC

(Onoji et al., 2016b).

3.5 Kinetics of thermal oxidative degradation of the seed oil

105
Vegetable oils (mainly triglycerides) undergo chemical reactions such as oxidation,

isomerization, polymerization and hydrolysis when subjected to prolonged heating process

(Onoji et al., 2016b). These reactions affect the physico-chemical properties of these oils and

their quality. The thermal study of oils before their industrial applications permits data to be

obtained on the influence of time and temperature on their thermal oxidation (Gouveia de Souza

et al., 2004).

In this study, classical chemical analysis methods were used to determine the change in

peroxide value, iodine value and refractive index of RSO when subjected to heat treatment

(Oderinde et al., 2009). Rubber seed oil has been reported to be thermally stable up to 250 oC

under heat treatment (Aravind et al., 2015). For this reason, the peroxide value, iodine value

and refractive index were determined at 100, 150, 200 and 250 oC at various time intervals

using methods prescribed by ASTM and AOAC (1990). The procedures outlined in Sections

E.1.7, E.1.8, and E.1.9 of Appendix E was followed at the conditions of temperature and time

considered for each case. The iodine value is a measure of the degree of unsaturation to quantify

the amount of double bonds present in the oil. It reflects its susceptibility to oxidation

(Hrušovský et al., 2013), and was used in the present study to generate data for kinetic plots.

The broad reaction rate expression proposed by several researchers for the kinetics of thermo-

oxidative degradation of RSO is expressed in Equation (3.2) (Santos et al., 2002; Van Boekel,

1996; Ramaswamy et al., 1989).

−𝑑𝐶 ⁄𝑑𝑡 = 𝑘𝐶 𝑛 (3.2)

where 'C' is iodine value at any time, 'n' is order of reaction, and 'k' is the reaction rate constant

whose temperature dependence is commonly described by Arrhenius Equation (3.3).

𝑘 = 𝐴𝑜 exp(− 𝐸𝑎 ⁄𝑅𝑇) (3.3)

106
T is the absolute temperature, R is the universal gas constant, Ea is the activation energy, and

Ao is the pre-exponential factor. The thermo-oxidative degradation of vegetable oils generally

follows a first-order reaction as given in Equation (3.4) (Ramaswamy et al., 1989).

𝑙𝑛 𝐶𝑡 ⁄𝐶𝑜 = 𝑘𝑡 (3.4)

where 𝐶𝑡 iodine value at time t is, 𝐶𝑜 is initial iodine value.

3.6 Optimization study of the oil extraction process using RSM and ANN–GA techniques

RSM based on Box–Behnken experimental design, and ANN–GA was used to model and

optimize the extraction process conditions for rubber seed oil production. A three-level, three-

factor design was employed in this study (Table 3.1). The extraction process parameters

investigated were rubber seed powder weight (X1), solvent volume (X2) and extraction time

(X3). The BBD generated 17 experimental runs that were randomized to minimize the

unpredictable variations in the observed responses due to uncontrolled extraneous factors. The

experimental runs included 12 factorial points, and 5 centre points that provide information on

the interior of the experimental regions to evaluate the curvature effect. The experiments were

carried out in series as described in the preliminary assay for the determination of optimum

particle size. An average particle kernel size of 0.5 mm that gave the maximum oil yield was

used to obtain experimental oil yields that were employed in RSM and ANN modeling (Onoji

et al., 2017a)

107
Table 3.1: Experimental range and levels of independent process variables for seed oil
extraction via Box-Behnken design (Onoji et al., 2017a)

Independent process variables Symbols Coded factor levels

(uncoded) –1 0 +1

Rubber seed powder weight (RSPW) (g) X1 40 50 60

Solvent volume (mL) X2 200 225 250

Extraction time (min) X3 40 45 50

3.6.1 Modeling of seed oil extraction process variables via RSM

The experimental oil yields obtained from BBD, and the process parameters were statistically

analyzed using RSM and least-squares method to fit a second-order polynomial model

generated by the Design-Expert software trial version 8.0.3.1 (Stat-Ease Inc., Minneapolis,

MN, USA) for the complete BBD (Onoji et al., 2017a). The relationship between the response

and the input variables is given in Equation (3.5) (Baş & Boyaci, 2007b):

Y  f ( X 1 , X 2 ,.....X n )   (3.5)

where Y is the response, f is the unknown function of the response, X1, X2, …. Xn denote the

actual independent process variables, n is the number of the independent variables and 𝜀 the

statistical error that represents other sources of variability not accounted for by f. It is generally

assumed that ε has a normal distribution with mean zero and variance (Baş & Boyaci, 2007b).

The relationship between independent coded and actual values is defined by Equation (3.6):

𝑋𝑖 −𝑋𝑖
xi = (3.6)
∆𝑋𝑖

where xi is the coded value of the ith factor, Xi is actual (experimental) value of the ith factor in

the uncoded units, 𝑋𝑖 is the average of the low and high values for the ith factor, and ∆𝑋𝑖

represents the step change for the ith factor.

108
The general second-order polynomial regression model shown in Equation (3.7) was used to

develop the empirical model that establishes the correlation between the coded process

parameters (x1, x2 and x3) and the response variable 𝑌 (% yield).

𝑌 = 𝛽𝑜 + ∑𝑘𝑖=1 𝛽𝑖 𝑥𝑖 + ∑𝑘𝑖=𝑖 𝛽𝑖𝑖 𝑥𝑖2 + ∑𝑘𝑖<𝑗 𝛽𝑖𝑗 𝑥𝑖 𝑥𝑗 + 𝜀 (3.7)

where 𝛽o, 𝛽i, 𝛽ii, and 𝛽ij represent the regression coefficients of intercept, linear, quadratic and

interactive terms, respectively, while xi and xj represent the coded process parameters.

The regression coefficients of Equation (3.7) were estimated using the least square estimation

method based on multiple regression technique. In this method, it is assumed that random errors

are identically distributed with a zero mean and a common unknown variance. The residual is

an estimate of the corresponding ε i which is the difference between the actual value (𝑌𝑖 ) and

the predicted value as shown in Equation (3.8).

𝜀𝑖 = 𝑌𝑖,𝑒𝑥𝑝− 𝑌𝑖,𝑝𝑟𝑒𝑑 (3.8)

The objective of the optimization step is to minimize the sum of the squares of the residuals,

often called the sum of squares of the errors (SSE). Thus,

SSE    i2   Yi ,exp  Yi , pred 


n n
2
(3.9)
i 1 i 1

The residuals may be expressed in different as in Equation (3.10), and SSE may be written as

shown in Equation (3.11) (Baş & Boyaci, 2007b).

ε =Y-x𝛽 (3.10)

SSE   T   Y  x  Y  x 
T
(3.11)

Differentiating the SSE with respect to β, yields a vector of partial derivatives (Equation

(3.12)), as follows:


SSE   2 x T Y  x  (3.12)


109
When Equation (3.12) is equated to zero, we have xβ = Y that is solvable directly to

obtain the coefficients β from Equation (3.13).

xT x  xT Y (3.13)

The rearrangement of Equation (3.13) in matrix form is given in Equation (3.14) (Baş &

Boyaci, 2007b):

  Cx T Y (3.14)

where C  x T x  is a squared matrix. After the regression coefficients are obtained from
1

Equation (3.14) using the experimental response Y, the model-predicted response could

be easily calculated using the developed model from Equation (3.7) with all the

regression coefficients incorporated.

The developed second-order regression model that adequately describes the behavior of the

process with respect to the operating variables was used to create contour (2D) and surface

responses (3D) plots. This was used to explore the designed space, and to predict the optimal

conditions for RSO extraction yield. The variability of the response variable was determined

by the adjusted multiple coefficient of determination (R2adj.) which is a measure of the amount

of variation around the mean explained by the model and adjusted for the number of terms in

the model. In addition, the adequate precision (signal-to-noise ratio) was calculated using

Equation (3.15) to verify that the quadratic regression model is adequate (Onoji et al., 2017a;

Cerino Córdova et al., 2011).

Max (𝑌)−𝑀𝑖𝑛 (𝑌)


𝐴𝑑𝑒𝑞𝑢𝑎𝑡𝑒𝑝𝑟𝑒𝑐𝑖𝑠𝑖𝑜𝑛 = 2
(3.15)
√𝜔𝜎
𝑛

where Max (Y) is the maximum predicted oil yield; Min (Y) is the minimum predicted oil yield;

ω is the number of model parameters (including βo); σ2 is residual mean square; and n is the

number of experimental runs.

110
3.6.2 Modeling of seed oil extraction process variables via ANN

A commercial ANN program, Neural Power trial version 2.5 (CPC-X software) was employed

in this study. RSO yield was predicted by using back-propagation multi-layer feed-forward

neural network algorithm. A topology of 3–3–1 network (Figure 3.2) corresponding to the

number of neurons within the input, hidden and output layers, respectively was used for the

training. A sigmoid transfer function was adopted for the study. The input vector consisted of

the rubber seed powder weight (X1), solvent volume (X2), and extraction time (X3) while the

output vector (Y) was the rubber seed oil (RSO) yield (Onoji et al., 2017a). The functional

relationship to be estimated by the ANN model could be expressed as follows:

Y  f X1 , X 2 , X 3  (3.16)

where X1, X2, X3, and Y are as defined earlier. The output of a neuron is computed from Equation

(3.17) (Shokri et al., 2011):

 n 
Oi  f   wij I j  bi  (3.17)
 j 1 

where Oi is output of the ith neuron, f is transfer function, bi is bias of the ith neuron, wij is

synaptic weight corresponding to jth synapse of ith neuron, Ij is jth input signal to the ith neuron

and n is number of input signals to the ith neuron.

bias bias
s
Rubber seed powder weight, X1 (g)

Solvent volume, X2 (mL) Y (%)

Extraction time, X3 (min)

Figure 3.2: Schematic of 3-3-1 developed ANN topology (Onoji et al., 2017a)

111
The experimental oil yields obtained from the 17 experimental runs of BBD was divided into

two sets. Experimental oil yields of 14 set of data was used for ANN training and cross

validation, and the other set of 3 data was used for testing of the trained network. The network

was trained using the back-propagation method implemented by Levenberg-Marquardt

algorithm depicted in Figure 3.3 (Onoji et al., 2017a). The training was carried out by adjusting

the connection weights and biases between neurons with the aim of reducing the mean square

error (𝑀𝑆𝐸) between the predicted and experimental outputs below an acceptable threshold

(Gueguim Kana et al., 2012). A network will memorize results instead of generalizing when it

is over trained. In such a case, the resulted model can perfectly predict the data similar to

training data, but will perform poorly if new set of data are submitted to it. The training was

completed when the network is able to predict the given output. The trained model was

validated by using it to predict RSO yield of the 3 data that were not used during the training

phase. The validated model was subsequently used as evaluation function (Equation 3.18) for

Genetic Algorithm (GA) optimization process to predict the optimal values of RSO extraction

variables (X1, X2, X3 and Y) applied in developing the model (Onoji et al., 2017a; Kostić et al.,

2013; Gueguim Kana et al., 2012).

1
f (X )  (3.18)
1 eX

112
Inputs: X1, X2, X3
Experiments Process Simulation

Actual Output (YA)

ANN model Predicted MSE=YA - YP


Output
(Yp)
Validate
ANN model

Adjust weights and


biases based on MSE

Min (MSE) MSE<Ɛ Save


Save best
best weights
Back-propagation min (MSE) and biases
weiuddudududud
NO YES
ududuights and

Figure 3.3: Levenberg-Marquardt training flowchart for ANN model (Adapted from

Onoji et al., 2017a).

3.7 Catalyst preparation and characterization

The rubber seed shells (RSS) freed from the kernels in Section 3.2.1 were stored in

polyethylene bags for analyses. The RSS was washed 3–4 times with hot distilled water to get

rid of dust and solid impurities. It was then dried in a laboratory oven at 110 oC for about 5 h,

and stored in a desiccator. The cooled dried RSS was ground to powder with a grinder, and

sieved through a 60-mesh size. The powdered raw RSS was analyzed by thermo-gravimetric

analyzer (TGA) in the range of 30–950 oC at 10 oC/min, under nitrogen environment to

determine a suitable calcination temperature of the catalyst (Onoji et al., 2017b). The RSS

powder was calcined from 40–800 oC at heating rate of 10 oC/min in an electric muffle furnace

(Carbolite, Parson Lane, Hope Valley S33 6RB, England, Model: RWF 12/5) for about 3 h to

113
remove organic materials in it and stored in screwed bottles for analyses. Raw and calcined

RSS (Figure E.1 of Appendix E) were dissolved in deionized water to determine the basic

property (pH). Energy dispersive X-ray spectroscopy (EDX 3600B-XRF Skyray Instrument)

analysis was employed to determine the elemental compositions of the raw and the calcined

RSS. Samples were pulverized using Knife Mill GRINDOMIX (model: GM20), and pelletized

with Mini-pellet press kit Asia (model: GS01152). The equipment was calibrated using pure

silver standard sample. The diffraction patterns of raw and calcined RSS samples were

analyzed with a diffractometer (XRD EMPYREAN mini-material analyzer, manufactured by

PANalytical B.V., Holland) to observe their amorphous and crystalline structures. Sample was

pulverized to homogeneous size and loaded into XRD sample holder. The diffractometer

equipped with PIXcel-3D detector employed Cu-Kα1 radiation source (λ =1.54059 Å) at 45

kV and 40 mA. Scanning was recorded in a continuous mode over 2o-Theta range from 4.0131o

to 79.9849o with a step size 0.026o at 13.77 second per step. The size and morphology of

catalyst was examined with scanning electron microscope (SEM-Phenom ProxMVE

016477830, manufactured by Phenom World, Eindhoven, Netherlands). The RSS sample was

dispersed on a stub using a sticky carbon tape before coating with palladium gold. Vacuum

was created before carrying out the analysis by inserting the stub into the equipment (Onoji et

al., 2017b).

Surface area and porosity, pore volume, and pore-size distribution are important parameters in

heterogeneous catalysis and catalyst design (Storck et al., 1998). Prior to analysis, about 335.3

mg of raw sample was weighed and outgassed at 250 oC for 11 h under a stream of nitrogen

gas to remove moisture and impurities. Similar analysis was carried out using 382.7 mg

calcined RSS catalyst at same conditions. These analyses were to evaluate the textural

properties of the raw and calcined RSS in order to determine the influence of calcination

temperature (Onoji et al., 2017b).

114
The accessibility of active sites is determined by the total surface area, and thus related to

catalytic activity. Gold App surface area and porosity analyzer (Model: V-sorb 2800p) was

used to estimate the surface textural properties of the raw and calcined RSS from isotherms

generated by Brunauer-Emmett-Teller (BET) model (Equation 3.19) (Vasudevan et al., 2011)

using nitrogen adsorption/desorption at 77.3 K.

P / Po   C  1  P 
1
(3.19)
V 1  P / Po  VmC Po VmC

where P is the partial pressure, Po is the saturation pressure, Vm is the maximum amount of

nitrogen adsorbed/unit mass of catalyst at high pressure conditions required to form a complete

monolayer over the entire surface of the catalyst, and C is the BET constant related to isosteric

heat of adsorption. Barrett-Joyner-Halenda (BJH) model calculated the pore-size distributions

of the raw and calcined RSS. The total pore volume is defined as the volume of liquid N2

corresponding to the amount adsorbed at a relative pressure of P/Po = 0.997, after which N2

desorption commences.

3.8 Biodiesel production from the extracted seed oil and analysis

The rubber seed oil used for the biodiesel production was earlier extracted as in Section 3.2.2.

The high FFA of the oil allows for two-step production routes: Esterification and

transesterification.

3.8.1 Acid-catalyzed esterification of the seed oil

The rubber seed oil used in the study had an initial acid value of 18.02±0.141 mg KOH/g oil,

and FFA level of 9.01±0.07% (Onoji et al., 2016b). Reduction of FFA to <1% prior to

transesterification reaction is necessary to avoid catalyst consumption in soap formation and a

low yield of biodiesel because of the basic nature of the synthesized catalyst (Betiku & Ajala,

2014). 100 mL rubber seed oil was measured into a 250-mL one-neck glass reactor and
115
preheated to 105 oC on a magnetic stirrer hot plate to remove moisture and volatile impurities.

Then, 1.5 % vol/vol H2SO4 catalyst based on methanol volume was mixed with 45 mL of

analytical grade methanol (>99%) and stirred for 5 min. The mixture was added to the oil

maintained at 60 oC, and the reaction continued for 1 h at a stirring rate of 600 rpm. The

esterified oil was transferred into a separating funnel and allowed to stand for 2 h (Figure F.1

of Appendix F). The pretreated oil was separated from the water formed (Equation (3.20)), and

excess methanol was evaporated in a rotary evaporator prior to acid value determination by

standard titration procedure described by ASTM. Similar experiments were carried out using

3, 4.5, and 6 % vol/vol H2SO4 and their acid values determined were recorded in Table F.1 of

Appendix F. The 6% vol/vol H2SO4 obtained for <1% FFA (90.5% conversion) was employed

to esterify the oil used for transesterification reaction (Onoji et al., 2017b).

H2SO4 (catalyst)
RCOOH + CH3OH ↔ RCOOCH3 + H2O
FFA methanol methyl ester water (3.20)

3.8.2 Biodiesel production from the esterified seed oil

A three-factor, three-level (Table 3.2) central composite design using Design-Expert® dx8

software version 8.0.7.1 (Design Ease Inc., USA) coupled with response surface methodology

was used to statistically analyze the experimental data that consisted of 8 factorial points, 6

axial points, and 6 center points. A transesterification reaction of the esterified seed oil was

performed according to the experimental design set by RSM and designed by employing CCD.

Twenty experimental runs were carried out for the optimization process under different

combination of process variables. In each run, and for the reaction time (X1), 180 mL of oil

specified by methanol/oil ratio (X2), and RSS catalyst (X3) as depicted in Table 3.2 were reacted

in a 500 mL three-necked round bottom reactor equipped with a reflux condenser, and magnetic

116
stirrer maintained at 400 rpm throughout the experiment (Onoji et al., 2017b). Heat was

supplied by magnetic hot plate at 60 oC for a reaction time set by the CCD. At the end of the

reaction time, the reaction was quenched in an ice-bath and the content of the reactor was

transferred into a separating funnel for 2 h to separate glycerol and other impurities from the

biodiesel (rubber seed oil methyl ester) phase (Figure F.2 of Appendix F). The biodiesel was

poured into a rotary evaporator to remove excess methanol, and washed thrice with deionized

water at 50 oC to remove entrained glycerol and residual catalyst. Washed biodiesel was dried

over anhydrous sodium sulphate at 50 oC (Dhawane et al., 2016). Dried biodiesel (Figure F.3

of Appendix F) was decanted and filtered to remove sodium sulphate, and the yield was

estimated using Equation (3.21).

𝑊𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑝𝑟𝑜𝑑𝑢𝑐𝑡 𝑜𝑏𝑡𝑎𝑖𝑛𝑒𝑑 × 𝐹𝐴𝑀𝐸%


𝐵𝑖𝑜𝑑𝑖𝑒𝑠𝑒𝑙 𝑦𝑖𝑒𝑙𝑑 (%) = × 100 (3.21)
𝑊𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑟𝑒𝑓𝑖𝑛𝑒𝑑 𝑟𝑢𝑏𝑏𝑒𝑟 𝑠𝑒𝑒𝑑 𝑜𝑖𝑙

where FAME% is the concentration of fatty acid methyl esters (FAME) analyzed by gas

chromatography-mass spectrometry (GC-MS).

Table 3.2: Experimental range and levels of independent process variables for biodiesel

production via central composite design

Independent process variables Symbols (uncoded) Coded factor levels

–1 0 +1

Reaction time (min) X1 60 65 70

Methanol/oil ratio (v/v) X2 0.20 0.25 0.30

RSS catalyst (g) X3 2.5 3.0 3.5

117
3.9 Characterization of the produced biodiesel

The physico-chemical properties such as density, color, moisture content, oAPI gravity,

kinematic viscosity, acid value, saponification value, iodine value, and calorific value were

determined. These properties including oxidation stability at 110 oC, cetane number, diesel

index, pour point, cloud point, flash point, fire point and cold filter plugging point of the methyl

ester (biodiesel) produced at optimum conditions were determined following the procedures

prescribed by ASTM standards and other methods (Onoji et al., 2017a). The oxidation stability

was determined by using PetroOXY automatic oxidation stability tester (Protest® Instruments

GmbH & Co. KG, Dahlewitz, Germany). Figure F.4 of Appendix F and Equation (3.22)

developed by Botella et al. (2014) with adjusted R2 of 0.915 at 95% confidence interval was

used to estimate the Rancimat value of the oxidation stability of the produced biodiesel without

additive.

𝑅𝑎𝑛𝑐𝑖𝑚𝑎𝑡 (𝑚𝑖𝑛) = [(31.89 − 20.63𝑓)(𝑃𝑒𝑡𝑟𝑜𝑂𝑋𝑌(𝑚𝑖𝑛)) + (−214.65 + 319.68𝑓)]

(3.22)

where f = 0 if no additive is used, and 1 if an additive has been mixed with biodiesel while

PetroOXY (min) is the value obtained from the Tester.

The fatty acid methyl ester (FAME) content of the biodiesel was determined by GC Agilent©

7890B and MS Pegasus 4D GCXGC (LECO) located in Wits University Johannesburg, and

the value obtained was compared with the correlation of Equation (3.23) developed by

Felizardo et al. (2006).

FAME (%) = – 45.055ln𝜇 + 162.85 (3.23)

whereµ is the kinematic viscosity (mm2/s) at 40 oC.

118
3.10 Reusability test for the catalyst used in biodiesel production

Different loadings (2, 2.2, 2.5, 3, and 3.5 g) of the prepared catalyst at optimum reaction

conditions of RSM, and at 60 oC were used to produce biodiesel in order to determine the level

of loss in activity in the reusability test. After the completion of each reaction, the biodiesel

yield was determined. The catalyst was separated from the reaction mixture by filtration,

centrifuged, washed with acetone to remove impurities blocking the active sites, and oven dried

at 110 oC for 3 h before each successive reuse. Leaching of active sites into the reaction mixture

was determined by atomic absorption spectroscopy (AAS) (Onoji et al., 2017b).

3.11 Statistical analysis of the biodiesel production by RSM

The general second-order polynomial regression model described by Equation (3.24) was used

to develop the empirical model that describes the correlation between the transesterification

uncoded process parameters (X1, X2, and X3), and the predicted response variable, RSOME

yield (%) (Onoji et al., 2017b).

𝑅𝑆𝑂𝑀𝐸 (%) = 𝛽𝑜 + ∑𝛼𝑖=1 𝛽𝑖 𝑥𝑖 + ∑𝛼𝑖=𝑖 𝛽𝑖𝑖 𝑥𝑖2 + ∑𝛼𝑖<𝑗 𝛽𝑖𝑗 𝑥𝑖 𝑥𝑗 + 𝑒 (3.24)

where βo, βi, βii, and βij represent the regression coefficients of intercept, linear, quadratic, and

interactive terms, respectively, while xi and xj, represent the coded independent process

parameters as defined in Equation (3.25), and 'e' the experimental error of analysis.

𝑋𝑖 −𝑋𝑖
xi = (3.25)
∆𝑋𝑖

Xi is actual (experimental) value of the ith factor in the uncoded units, 𝑋𝑖 is the average of the

low and high values for the ith factor, and ∆𝑋𝑖 represents the step change for the ith factor.

119
The adequate precision ratio (signal-to-noise ratio) which verifies the adequacy of the

developed model was estimated by Equation (3.26) (Cerino Córdova et al., 2011).

Max (𝑌)−𝑀𝑖𝑛 (𝑌)


𝐴𝑑𝑒𝑞𝑢𝑎𝑡𝑒 𝑝𝑟𝑒𝑐𝑖𝑠𝑖𝑜𝑛 = 2
(3.26)
√𝜔𝜎
𝑛

where Max (Y) is the maximum predicted RSOME yield (%), Min (Y) is the minimum predicted

RSOME yield (%), ω is the number of model parameters (including 𝛽𝑜 ), σ2 is residual mean

square, and n is the number of experimental runs. The derived model was used to create surface

(3-D) plots to explore the designed space, and predict the optimal conditions for RSOME.

3.12 Analysis of variance (ANOVA) study for biodiesel production

One-way analysis of variance (ANOVA) was used for evaluating statistical significance of the

experimental polynomial model, coefficients, process factors, and the quality of the model fit.

Optimal process conditions for maximizing the yield of RSOME were determined by solving

the model equation using the Design-Expert® dx8 software version 8.0.7.1. The coefficient of

variance (% C.V.), a ratio of standard deviation to mean value was determined to check the

capability of the model to accurately predict yields, and a low value is desirable.

3.13 Modeling of biodiesel production process variables by ANN

ANN Neural Power trial version 2.5 (cpc-x software) was employed in the study. RSOME

yield (%) was predicted using the supervised back-propagation multilayer feed-forward

algorithm. The network consists of an input layer (three neurons), hidden layer (three neurons),

and output layer (one neuron). The same process parameters and experimental data used for

the RSM modeling were deployed in the ANN architectural design. Training of the network,

cross validation and testing were carried out, and the ANN-Genetic Algorithm model was

successfully used as an evaluation function for the optimization of its input space to predict the

120
optimal values (Onoji et al., 2017a) of RSOME yield (%) and process parameters (X1, X2, and

X3) applied in developing the model.

3.14 Data validation for RSM and ANN models

The predicted optimized responses obtained from RSM and ANN models were compared with

the average experimental values from triplicate runs obtained using the optimum conditions in

order to evaluate the efficiencies of these optimization techniques. The percent relative error

(PRE) (Zahedi & Azarpour, 2011) and absolute average deviation (AAD) (Betiku & Ajala,

2014) were determined and these were used to identify the best model by comparing evaluated

values. The PRE and AAD values were calculated using Equations (3.27) and (3.28),

respectively.

𝑜𝑝𝑡𝑖𝑚𝑎𝑙 𝑎𝑐𝑡𝑢𝑎𝑙 𝑦𝑖𝑒𝑙𝑑−𝑜𝑝𝑡𝑖𝑚𝑎𝑙 𝑝𝑟𝑒𝑑𝑖𝑐𝑡𝑒𝑑 𝑦𝑖𝑒𝑙𝑑


𝑃𝑅𝐸 (%) = × 100 (3.27)
𝑜𝑝𝑡𝑖𝑚𝑎𝑙 𝑎𝑐𝑡𝑢𝑎𝑙 𝑦𝑖𝑒𝑙𝑑

𝑅𝑆𝑂𝑀𝐸𝑖,𝑎𝑐𝑡𝑢𝑎𝑙 −𝑅𝑆𝑂𝑀𝐸𝑖,𝑝𝑟𝑒𝑑 1
𝐴𝐴𝐷 (%) = {[∑𝑛𝑖= ( )] 𝑛} × 100 (3.28)
𝑅𝑆𝑂𝑀𝐸𝑖,𝑎𝑐𝑡𝑢𝑎𝑙

Where n is the number of experimental runs, while RSOMEi, actual is the actual yield (%), and

RSOMEi, pred is the predicted yield (%).

3.15 Combustion of biodiesel-diesel blends in internal compression ignition engine

Five litres of commercial diesel was purchased from Nigerian National Petroleum Corporation

(NNPC) fuel station located at Refinery road, Effurun, Delta State, Nigeria. The diesel and the

rubber seed oil biodiesel were characterized to ascertain its uniformity with ASTM (American

Society for Testing and Materials) standards for biodiesel and diesel fuels. The biodiesel-diesel

121
blends were prepared and tested for performance, and combustion characteristics using internal

compression ignition engine.

3.15.1 Biodiesel-diesel blending analysis via ultrasonication technique

Figure 3.4 depicts the experimental setup for the blending of biodiesel mixtures using the

ultrasonication blending technique. Four different blends of B10, B20, B30, and B50 were

prepared on volumetric basis for test in internal compression ignition engine. Pure diesel (B00)

and biodiesel (B100) fuels were also tested to determine their levels of environmental

friendliness with those of their blends. For each blend (Bxx), where B stands for biodiesel

blend, and xx for the % volume of biodiesel in the blend (e.g. B10 means 10% vol biodiesel,

and 90% vol diesel), the measured volume of biodiesel was poured into a batch blender. The

ultrasonicator horn was lowered to a depth of 2 cm within the biodiesel and then switched on.

Then, an appropriate measured volume of diesel was added to the biodiesel at a constant rate

and allowed to blend for 5 min under standard atmospheric conditions. The ultrasonicator is

then switched off and the mixture emptied into a measuring cylinder for combustion analysis

in an internal compression ignition engine.

122
Ultrasonic
transducer

Ultrasonic processor

Thermometer

Ultrasonic horn

biodiesel – diesel blends

Figure 3.4: Experimental setup for biodiesel-diesel blending analysis

3.15.2 Methodology and experimental setup for combustion study of biodiesel blends

The experimental research was performed on a naturally aspirated single-cylinder, four-stroke,

direct-injection, air-cooled, TecQuipment TD202 (Model: 1B20) Hatz diesel test engine that

was coupled to TecQuipment TD200 small engine test bed shown in Figure 3.5, and connected

to a separate Versatile Data Acquisition System (VDAS) to accurately monitor and record all

measured data on a computer system. The test engine is located in the Biofuels Laboratory at

Petroleum Training Institute, Effurun, Nigeria. The detailed technical specifications of the test

engine are listed in Table 3.3.

123
Figure 3.5: Experimental TD202 diesel test engine for performance and combustion

analyses located at Petroleum Training Institute, Effurun, Nigeria

Table 3.3: Technical specifications for TD 202 diesel test engine

Engine make/model Hatz/1B20

Engine type Single cylinder, direct injection and air-cooled

Displacement (Capacity) 232 cm3

Number of strokes 4

Aspiration/starting system Natural/Recoil starter

Cylinder bore/connecting rod length 69 mm/104 mm

Stroke/crank radius 62 mm/31 mm

Compression ratio 22:1

Absolute maximum power 3.5 kW at 3600 rpm

Continuous rated power 3.1 kW at 3000 rpm

Orifice diameter/Coefficient of discharge 0.0185 m/0.6

124
The test engine set included a robust cost effective and efficient hydraulic dynamometer to

control the engine load as shown schematically in Figure 3.6. There are no load resistors needed

because the engine power is dissipated into the water that passes through the dynamometer.

The engine was allowed to warm up before the experiments were conducted to ensure

parameters were being analyzed at steady state (Özener et al., 2014). The performance and

operating parameters, such as the engine torque (Nm), engine power (Watts), exhaust gas

temperature (oC), ambient air temperature, TA (oC), air-box differential pressure, ∆P (Pa), and

ambient pressure, PA (mbar) were measured at the engine speeds of 1500, 2000, 2500, 3000,

and 3500 rpm at fixed 100% engine loading and were digitally displayed on instrument

modules. They were recorded after the conditions of the engine stabilized. The parameters

considered for the engine performance and emission analysis that indicate the applicability of

the produced biodiesel in the engine are air/fuel ratio, brake mean effective pressure (BMEP),

brake thermal efficiency (BTE %), break power (BP), engine torque, brake specific fuel

consumption (BSFC), and exhaust gas temperature for performance evaluation consideration.

The CO, CO2, NOx, total hydrocarbons (THCs), and smoke opacity were considered for the

emission analysis with respect to the test engine. The engine power (Watts) was displayed

digitally on the instrument module (Figure 3.5) and could be calculated using the module

displayed speed (rpm) and the torque (Nm) based on Equation (3.29) used by Ude et al. (2017).

TN
BP  (3.29)
9549.3

where:

BP represents the engine brake power (kW); T represents the developed torque (Nm), and N is

the engine speed (rpm).

125
The air velocity (m/s), air flowrate (kg/s), air/fuel ratio, and the brake mean effective pressure

(bar) are described by Equations (3.30), (3.31), (3.32), and (3.33), respectively (TecQuipment

Ltd, 2011).

2∆𝑃
𝑈 = √𝜌 (3.30)
𝑎𝑖𝑟

𝐶𝑜 𝜋𝑑2 2𝑃 ∆𝑃
𝑚̇𝑎 =
4
√ 𝑅 𝐴𝑇 (3.31)
𝑎 𝐴

𝐴𝑖𝑟 𝑚̇𝑎
𝑟𝑎𝑡𝑖𝑜 = (3.32)
𝐹𝑢𝑒𝑙 𝑚̇𝑓

60 × 𝐵𝑃 × (𝑆𝑡𝑟𝑜𝑘𝑒/2)
𝐵𝑀𝐸𝑃 = (3.33)
0.1 ×𝑆𝑝𝑒𝑒𝑑 ×𝐸𝑛𝑔𝑖𝑛𝑒 𝐶𝑎𝑝𝑎𝑐𝑖𝑡𝑦

where:

U = velocity of air (m/s);

∆P = air-box differential pressure (Pa);

ρair = density of air (1.2754 kg/m3);

d = diameter of the orifice plate (= 0.0185 m for the test engine used);

Co = coefficient of discharge (= 0.6 for the test engine used);

𝑚̇𝑎 = mass flowrate of air (kg/s);

𝑚̇𝑓 = mass flowrate of fuel (kg/s)

PA = ambient pressure (Pa); TA = ambient air temperature (K);

Ra = Gas constant for air (287 J/kg K)

BMEP = brake mean effective pressure (bar);

BP = brake power (Watts);

126
Engine capacity = 232 cm3 for the test engine used, and speed expressed in rpm.

Stroke = 4, for the test engine.

The engine speed was measured using an optical sensor, while the torque was measured using

a load cell. The volumetric flowrate of fuel at each engine speed, for each blend was calculated

using the time (measured by a stopwatch) the engine takes to consume 8 mL of fuel using the

8-mL pipette fitted gauge (AVF1 option). The inlet air flowrate was measured using an orifice

plate of diameter 18.5 mm. A k-type thermocouple and a differential pressure transducer were

fitted near the orifice plate in order to measure the air temperature and pressure, respectively,

needed for the inlet air flowrate calculation. An engine cycle analyzer (ECA 100) received the

signals of both the pressure transducer and shaft encoder.

A Quintox flue gas analyzer (Model: KM 9106) manufactured by KANE International Ltd.,

Swallow fields, Welwyn Garden City, Herts, AL 1JG, UK, and supplied with combustion

analysis display unit was employed to measure the exhaust gas emissions (CO, CO2, NOx, and

THC) at the above stipulated engine speeds, while the smoke opacity was measured with AVL

DiSmoke 4000 opacity smoke meter.

127
Fuel tank
Computer
Data Acquisition System

8 mL
Torque input

Speed input
Air box
pressure 16mL
input Exhaust Temperature
Thermocouple input

To exhaust Optical
Engine gas sensor
cycle analyzer
Signal
Differential
analyzer pressure
Cylinder pressure Air temperature Signal
transducer
Signal Orifice
Shaft encoder
Plate
& rpm sensor
Load cell
Air box
Fuel
Hydraulic flow
dynamometer Test Engine

Fuel flow

Figure 3.6: A schematic layout of experimental TD202 diesel test engine

128
CHAPTER 4: RESULTS AND DISCUSSION

4.0 Introduction

This chapter documents and discusses all the results obtained in this study. The results were

compared with the literature, and discrepancies if any were explained.

4.1 Extraction of seed oil and determination of the best seed particle size

The results depicted in Figure 4.1 showed that a seed particle size of 0.5 mm gave the maximum

oil yield of 40.3 wt. % among the available sizes during the oil extraction (Onoji et al., 2017a).

The data for the plots are shown in Table D.1 of Appendix D. The findings of this study concur

with those of Menkiti et al. (2015) who extracted Terminalia catappa L seed oil using n-hexane

solvent, and obtained a yield of 60.45 wt. % from the smallest kernel particle size 0.5 mm

compared to the other sizes (1, 1.5, 2, and 2.5 mm) considered. A similar trend in the reported

work of Reshad et al. (2015) where the smaller particle size of 1 mm gave a rubber seed oil

yield of 49.36 wt. % for 8 h extraction time was also observed. The smaller particle size creates

a larger surface area that enhances oil yield by facilitating solvent diffusivity in the seed

powder. In the present study, particle size of 0.5 mm was considered for oil extraction for the

modeling and optimization processes, and chemical analyses. However, for industrial

application, 1 mm particle size with a yield of 40.1 wt. % is preferred to 0.5 mm size as less

energy is required to process the 1 mm kernel size (Onoji et al., 2017a).

129
45

40
Oil yield (%)

35

30

25
0.5 1 1.5 2 2.5
kernel size (mm)

Figure 4.1: Effects of kernel particle size on the rubber seed oil yield

4.2 Physical characterization of the seed and seed oil

Table 4.1 presents the analysis of rubber seed and the extracted seed oil (Onoji et al., 2016b).

The computations are presented in Section B.1 of Appendix B. The kernels moisture content

(9.7 wt.%) of rubber seed evaluated was higher than the reported value of 6.5 wt.% for yellow

horn seed kernels (Li et al., 2013). The determined rubber seed oil moisture content is low

(1.73 wt.%) compared to castor oil (8 wt.%) and shea nut oil (10 wt.%) (Asuquo et al., 2012),

Afzelia africana oil (2.10 wt.%) and Hura crepitans oil (1.90 wt.%) (Ogbu & Ajiwe, 2016).

High moisture content initiates oxidation process that reduces the oil's shelf life. The low value

for this oil is an indication that it could be stored for a long period without appreciable

deterioration in value. The ash content of the rubber seed oil (0.001 wt.%) is lower than the

prescribed ASTM ash limit (0.01 wt.% max) and those of Afzelia africana oil (0.006 wt.%) and

Hura crepitans oil (0.004 wt.%) (Ogbu & Ajiwe, 2016). The low ash value is an indication that

the oil lacks trace metals that catalyze oxidation reactions that cause rancidity, high acidity and

130
other unpleasant characteristics during storage. The carbon residue (0.4 wt.%) is close to the

ASTM standard limit of 0.35 wt.% max for diesel fuels (Demirbaş, 1998), but higher than other

vegetable oils such as sunflower (0.03 wt.%), rapeseed (0.05 wt.%) and olive (0.09 wt.%)

(Anastopoulos et al., 2009). The high volatile content of the oil suggests that it could be a

plausible source material for biodiesel synthesis for use in internal compression ignition

engines.

Table 4.1: Physical characterization of rubber seed and the seed oil
Parameter Composition (wt.%)

Rubber seed

Kernel 40.47 ± 0.07

Kernel moisture 9.7 ± 0.07

Shell 49.83± 0.14

100

Rubber seed oil

Conradson carbon residue 0.4 ± 0.028

Moisture content 1.73 ± 0.028

Ash content 0.001 ± 0.0001

Volatile matter 97.869± 0.056

100

values are mean ± standard deviation of duplicate data.

4.3 Characterization of the seed oil

The extracted rubber seed oil was subjected to physico-chemical analysis following the

procedures prescribed in AOAC (1990) and ASTM methods. A comparison of the properties

of the oil with values in the literature (Onoji et al., 2016a; Aravind et al., 2015; Reshad et al.,
131
2015) for rubber seed oil is presented in Table 4.2. The results of the analyses indicate the

potentials of rubber seed oil as an alternative feedstock to edible oils for commercial production

of biodiesel, and other industrial applications of oils where non-edible oil is preferable (Onoji

et al., 2017a).

4.3.1 Colour, pH, density and Specific gravity

The colour of the rubber seed oil was dark brown after clarification, and it remained liquid at

room temperature (Onoji et al., 2016b). The pH value of 6.0 obtained for the oil compares

favorably with those reported for castor oil (6.8) and luffa cylindrica seed oil (3.93) (Asuquo

et al., 2012). This value is an indication of the presence of reasonable amount of free fatty acid

and the advantageous utilization of the oil in soap making (Hosamani & Katagi, 2008). The

density at 25 oC was determined using a pycnometer and was found to be 0.886 g/cm3, which

implies that the oil is less than water with the absence of heavy elements. The specific gravity

at 15 oC was determined as 0.909 using standard specific gravity bottles and the oAPI gravity

of 24.1 was calculated for the oil (Onoji et al., 2016b). Specific gravity is an indication of the

energy content of a fuel. A fuel with a high specific gravity (low oAPI) has a higher heating

value than a fuel with a low specific gravity (high oAPI). The physical characteristics seem to

be similar to other common vegetable oils, and the oil was validated as environmentally

friendly.

4.3.2 Iodine, peroxide, saponification and acid values

Iodine value (IV) measures the degree of unsaturation in a fat or vegetable oil and determines

its oxidative stability. The iodine value of rubber seed oil (137.02 g I2/100 g oil) was estimated

by Wijs' method and falls within the range specified for semi-drying oils (100–150 g I2/100 g)

(Asuquo et al., 2012). The higher iodine value may be due to high-unsaturated fatty acid

132
content, as observed in GC-MS analysis of oil. The peroxide value (PV) determines the extent

to which the oil has undergone rancidity when stored, heated, or in contact with air and could

be used to assess the quality and stability of oils. Oils become rancid when the PV range from

20–40 mequiv O2/kg oil (Bora et al., 2014). Standard Organization of Nigeria (SON, 2000) and

the Nigerian Industrial Standards (NIS, 1992) specify a peroxide value of 10 mequiv O2/kg oil

for edible oils. The PV obtained for rubber seed oil in this study was 10.46 mequiv O 2/kg oil,

and it is within the range reported in literature for RSO (1.6–16 mequiv O2/kg oil) (Onoji et al.,

2016a), and close to Nigerian standards for edible oils. This indicates that the rubber seed oil

can be stored for a long period without deterioration. The saponification value (SV) indicates

the amount of alkali required to convert the oil into soap and is an index of average molecular

mass of fatty acid in the oil sample (Toscano et al., 2012; Anastopoulos et al., 2009). As oil is

mainly triglycerides, it allows for comparison of the average fatty acid chain length. The SV

of 195.3 mg KOH/g RSO obtained in this study (Onoji et al., 2016b) lies between those of

sunflower (186) and coconut oils (265) (Aravind et al., 2015), and within the range of 195–205

mg KOH/g oil for edible palm oils as specified by SON (2000) and NIS (1992). This shows

that rubber seed oil has lesser tendency to saponify at elevated temperatures than coconut oil.

The acid value (AV) of rubber seed oil was 18.20 mg KOH/g oil, which indicates high levels

of free fatty acids (9.10 wt.%), a limiting factor in its use as a source of food. The acceptable

limit for edible oils is ≤ 10 mg KOH/g oil. Therefore, chemical re-esterification or physical

refining would be required to de-acidify the oil (Ebewele et al., 2010b). However; this value

is low enough and is within the limits for industrially useful oils (Oyedeji & Oderinde, 2006).

4.3.3 Kinematic viscosity

Kinematic viscosity (KV) is a measure of the resistance of oil to shear. The dynamic (absolute)

viscosity for the rubber seed oil was digitally determined at 40 oC using Brookfield DV - I +

133
Viscometer (Model LVDV - |+) manufactured by Brookfield Engineering Laboratories Inc.,

Middleboro, MA 02346 USA. It was operated at 100 rpm with spindle No. 61 and a Torque of

59.2%. The average KV of the RSO was calculated as 35.6 cP and converted to 40.18 mm2/s

based on Equation (4.1). This is a low value compared to jatropha curcas L. oil (77.4 mm2/s)

but higher than that of conventional diesel (<5 mm2/s) (Reshad et al., 2015).

𝐴𝑏𝑠𝑜𝑙𝑢𝑡𝑒 𝑣𝑖𝑠𝑐𝑜𝑠𝑖𝑡𝑦 (𝑐𝑃)


𝐾𝑖𝑛𝑒𝑚𝑎𝑡𝑖𝑐 𝑣𝑖𝑠𝑐𝑜𝑠𝑖𝑡𝑦 (𝑚𝑚2 ⁄𝑠) = (4.1)
𝐷𝑒𝑛𝑠𝑖𝑡𝑦 (𝑔⁄𝑐𝑚3 )

4.3.4 Cold flow properties

The flow properties of RSO were observed under low temperatures, and the parameters

considered are pour point (PP), cloud point (CP) and cold filter plugging point (CFPP),

respectively.

4.3.4.1 Pour point (PP)

The PP is the lowest temperature at which the oil suffers from gel formation and attains semi-

solid state and become deprived of its flow ability (Verma et al., 2016). The value of –6 oC

observed for this rubber seed oil is lower than that of Jatropha oil (4 oC) (Reshad et al., 2015;

Mazumdar et al., 2012) and castor oil (–2 oC), comparable to waste cooking oil (–7 oC), but

higher compared to canola (–18 oC) (Reshad et al., 2015). The value obtained is close to

reported value of –9 oC for RSO (Aravind et al., 2015). This suggests that it can be employed

under cold climatic conditions as a source material for lubricating oil formulation.

4.3.4.2 Cloud point (CP)

CP is the temperature at which a cloud or haze first appears in a liquid when cooled under

carefully controlled conditions. The observed CP for rubber seed oil was 5.5 oC. This value is

lower than 10–11 oC reported for Jatropha oil (Mazumdar et al., 2012). This suggests that RSO

134
is mainly unsaturated and a good source material for biodiesel synthesis for use in cold regions

compared to Jatropha oil (Onoji et al., 2016b).

4.3.4.3 Cold filter plugging point (CFPP)

The CFPP measures the lowest temperature at which a fuel gives trouble-free flow in a fuel

system. The CFPP value of –0.025 oC was estimated from Equation (4.2) cited by Verma et al.

(2016).

CFPP  0.8537  CP  4.72 (4.2)

The low CFPP confirms that the rubber seed oil is mainly unsaturated fatty acids and its use is

favourable in cold climatic regions.

4.3.5 Aniline point

The aniline point (AP) of oil is defined as the minimum temperature at which equal volumes

of aniline and oil are miscible. It is a measure of the aromatic content of oil and an evaluating

factor for diesel index determination and cetane number. At room temperature (28 oC),

complete miscibility of rubber seed oil with an equal volume of aniline was observed. The oil

was cooled to 10 oC in an ice-bath and mixed with equal volume of aniline to form a 2-phase

mixture. The mixture was heated gradually until complete miscibility was observed at 21 oC

(69.8 oF) as the aniline point (Onoji et al., 2016b).

4.3.6 Boiling, freezing, flash and fire points

The boiling point was determined as 119 oC using Stanhope-Setastill apparatus (Model: 11860-

3U). This suggests that moisture present could be removed at 105 oC without affecting its

properties prior to industrial applications such as biodiesel and lubricant synthesis. The freezing

point was observed at –18 oC using Stanhope-Seta Cloud and Pour Point Refrigerator Surrey,

135
England (Test Bath No. 4) calibrated to –51 oC. The fire and flash points were determined by

using Stanhope-Seta Pensky-Martens (A and B closed Cup-Model 34100-2/34000-0U) multi-

flash equipment. The flash and fire points were observed as 240.3 and 256 oC, respectively.

This shows that the oil can be stored safely at room temperature.

4.3.7 Refractive index (RI)

The refractive index (RI) is the ratio of the velocity of light in vacuum to the velocity of light

in a medium (in this case, RSO) and an indication of the level of saturation of the oil. RI was

determined at 20 oC using Abbe Refractometer (Model: 60/ED) and calibrated chart. The value

of 1.4707 calculated for RSO falls within the range reported in the literature for vegetable oils

(Rashad et al., 2015).

4.3.8 Other parameters

Other parameters determined for RSO are cetane number, diesel index, higher heating value,

and average molecular weight.

4.3.8.1 Cetane number (CN)

Cetane number is a measure of the tendency a fuel to knock in a diesel engine. It is an indication

of the ignition quality of a fuel. CN for RSO was calculated as 43.42 using the correlation

developed by Mofijur et al. (2014) as expressed in Equation (4.3).

𝐶𝑁 = 46.3 + (5458⁄𝑆𝑉) − (0.225 × 𝐼𝑉) (4.3)

This value compares favourably with those reported in the literature (45–49.73) for RSO

(Omorogbe et al., 2013; Kumar et al., 2010). However, with the recent price instability of crude

oil in the world markets, there is renewed interest in the use of vegetable oils such as RSO to

produce biodiesel for use in diesel engines.

136
4.3.8.2 Diesel index (DI)

The diesel index of RSO is computed using Equation (4.4) as follows:

𝐴𝑛𝑖𝑙𝑖𝑛𝑒 𝑝𝑜𝑖𝑛𝑡 (℉)× 𝐴𝑃𝐼 𝑔𝑟𝑎𝑣𝑖𝑡𝑦 (60 ℉)


𝐷𝑖𝑒𝑠𝑒𝑙 𝑖𝑛𝑑𝑒𝑥(𝐷𝐼) = (4.4)
100

The value 15.71 obtained is below minimum standard of 50 specified for diesel fuel. Therefore,

this oil cannot be used directly on diesel engines without chemical conversion because of its

high viscosity that will result to poor atomization in engine during operation.

4.3.8.3 Higher heating value (HHV)

The HHV of vegetable oils can be calculated by using SV and IV obtained from simple

chemical analyses using common laboratory equipment. In this study, a simple correlation (R2

= 0.9999) developed by Demirbaş (1998) as expressed in Equation (4.5) was used to estimate

HHV for RSO.

𝐻𝐻𝑉 (𝑀𝐽⁄𝑘𝑔) = 49.43 − [0.041(𝑆𝑉) + 0.015(𝐼𝑉)] (4.5)

The value obtained (39.37 MJ/kg) is within the range reported in literature for RSO (36.1–44

MJ/kg) (Onoji et al., 2016a; Reshad et al., 2015) and comparable to those reported by Rahman

et al. (2014) for jatropha curcas L. oil (38.66).

4.3.8.4 Average molecular weight of rubber seed oil

Vegetable oils are mainly triglycerides of three fatty acid chains with a glycerol backbone. The

average molecular weight of vegetable oils can be estimated using simple analytical processes

in laboratories. Higher molecular mass indicates higher heating values for the oil. The

determination of the molecular weight of vegetable oil is important for biodiesel production

reactions because the quantity of reagents used is based on molecular weight of the vegetable

137
oil. The average molecular weight of the RSO was calculated using Equation (4.6) reported by

Fillières et al. (1995).

𝐴𝑣𝑔. 𝑚𝑜𝑙𝑒𝑐𝑢𝑙𝑎𝑟 𝑤𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑜𝑖𝑙 (𝑔⁄𝑚𝑜𝑙. ) = 3𝑀𝑤𝐹𝐴 + 𝑀𝑤𝑔𝑙𝑦 − 3𝑀𝑤𝑤𝑎𝑡𝑒𝑟 (4.6)

where 𝑀𝑤𝐹𝐴 is the mean molecular weight of fatty acids present, 𝑀𝑤𝑔𝑙𝑦 is the molecular weight

of glycerol (92.09), and 𝑀𝑤𝑤𝑎𝑡𝑒𝑟 is the molecular weight of water (18.01).

The mean molecular weight of fatty acids of RSO was estimated from Equation (4.7) reported

by Ajiwe et al. (1995) and calculated as 286.74.

𝑀𝑒𝑎𝑛 𝑚𝑜𝑙𝑒𝑐𝑢𝑙𝑎𝑟 𝑤𝑒𝑖𝑔ℎ𝑡 𝑜𝑓 𝑓𝑎𝑡𝑡𝑦 𝑎𝑐𝑖𝑑𝑠 = (56⁄𝑆𝑉) × 1000 (4.7)

The average molecular weight of the RSO estimated from Equation (10) is 898 g/mol. The

value obtained compares favourably with those reported by Anastopoulos et al. (2009) for

sunflower oil (876), rapeseed oil (992), olive oil (857), and used frying oil (882).

138
Table 4.2: Physico-chemical properties and other characteristics of rubber seed oil.

Parameters Refs.a, b, c This study


Colour Golden yellow, light/dark brown Dark brown
Density, g/cm3 @ 25 oC 0.857–0.943 0.886± 0.002
Specific gravity @ 15 oC 0.91 0.909± 0.002
o
API gravity @ 15 oC NA 24.1± 0.282
Ph 6 6± 0.141
Oil content (wt.%) 40–50 43± 0.141
Iodine value, g I2/100 g oil 113–146 137.02± 0.028
Peroxide value, meq. O2/kg oil 1.6–16 10.46± 0.098
Saponification value, mg KOH/g oil 183.91–235.28 195.30± 0.282
Acid value, mg KOH/g oil 1.68–42.41 18.20± 0.141
Free fatty acid (%FFA as oleic acid) 0.84–42.412 9.10± 0.07
Kinematic viscosity, mm2/s @ 40 oC 6–66 40.18± 0.028
Refractive index @ 20 oC 1.46–1.47 1.4707±
0.0003
Pour point, oC –9 to –1.5 –6
Cloud point, oC 3–4 5.5
Cold filter plugging point, oC NA –0.025
Flash point, oC 72–295 240.3
Fire point, oC 298 256
Aniline point, oC (oF) NA 21 (69.8)
Boiling point, oC NA 119
Freezing point, oC NA –18
Cetane number 45–49.73 43.42
Higher heating value (HHV), MJ/kg 36.1–44 39.37
Mean mol. weight of fatty acids, (g/mol) NA 286.74
Average molecular weight of RSO, (g/mol) NA 898
Diesel index NA 15.71
a
Onoji et al., 2016a; bAravind et al., 2015; c Reshad et al., 2015.
NA: means not available; values are mean ± standard deviation of duplicate data.

139
4.3.9 Fatty acid profile of the seed oil
Gas chromatography-mass spectrometry (GC-MS) analysis was employed to characterize the

oil to determine its fatty acid profile. The identification of compounds and their structures was

based on data from NIST library. The results presented in Table 4.3 and Figure 4.2 indicates

that the oil is mainly unsaturated. The fatty acids present are palmitic (13.85%), stearic

(16.82%), oleic (64.11%), and cis-erucic (5.22%). The total unsaturated fatty acid composition

of the rubber seed oil was 69.33% with a saturated content of 30.67% (Onoji et al., 2016b).

Table 4.3: Fatty acids profile of rubber seed oil

Fatty acid/systemic name Chemical formulae Compositions (wt. %)


Ref.q This study
Myristic (C14:0) / C14H28O2 2.2 –
Tetradecanoic
Palmitic (C16:0)/ C16H32O2 0.23–10.6 13.85
Hexadecanoic
Palmitoleic (C16:1)/ C16H30O2 0.23–0.25 –
Hexadec-9-enoic
Stearic (C18:0)/ C18H36O2 5.69–12 16.82
Octadecanoic
Oleic (C18:1)/ cis-9- C18H34O2 12.7–42.08 64.11
Octadecenoic
Linoleic(C18:2)/ cis-9-cis- C18H32O2 39.6–52.84 –
12-Octadecadienoic
α-Linolenic (C18:3)/ cis-9- C18H30O2 2.38–26 –
cis-12-cis-15-
Octadecatrienoic
Arachidonic (C20:0)/ C20H40O2 0.66–0.97 –
Eicosanoic
Erucic acid (C22:1)/cis-13- C22H42O2 – 5.22
Docosenoic
Total saturated 8.78–23.57 30.67
Total monounsaturated 12.93–42.33 69.33
Total polyunsaturated 41.98–78.84 Trace
(linoleic, linolenic, etc)
q
Onoji et al., 2016a

Although the composition of this oil did not follow the trend of some of the reported fatty acid

profile for rubber seed oil in the literature (Onoji et al., 2016a), genetic variations could be

140
responsible for the disparity observed in the fatty acid profile of seed oil from NIG 800 clonal

series of Nigerian rubber seeds. The high content of oleic acid and the presence of cis-erucic

acid indicate the potential use of seed oils from the NIG800 series as a source of oleo chemicals.

This could replace the oleo chemicals from petroleum currently used for the oleo chemical and

steel industries. The mean molecular weight of fatty acids present in the rubber seed oil was

computed as 280.81 using Equation (4.8). This value compares favourably with 286.74

obtained using chemical analysis method earlier shown in Equation (4.7) with 2.06 %

deviation.

f i
Mean molecular weight of fatty acids = f (4.8)
M i

wi

where f I = % composition of fatty acids from GC-MS analysis and M wi = molecular weight

of a fatty acid.

Figure 4.2: Gas chromatogram of the rubber seed oil with identified peaks: peak 4

(palmitic acid), peak 5 (oleic acid), peak 6 (stearic acid), peak 10 (cis-erucic acid) (Onoji

et al., 2016b)

141
4.3.10 FT-IR and NMR spectroscopy of the seed oil

The structural characterizations of rubber seed oil were further performed by FT-IR, 1H NMR,

and 13C NMR analyses to confirm results from GC-MS analysis. These analyses are essential

for industrial applications of the oil.

4.3.10.1 FT-IR analysis of the seed oil

Further insight into the chemical compositions of the rubber seed oil was provided by the FT-

IR spectrum at room temperature (28 oC) based on Table G.1 of Appendix G. It has been

reported that very low proportions of compounds present in a mixture exhibit very weak band

which are not detectable in the IR spectrum (Guillèn & Cabo, 2002). The absorption bands of

vegetable oils at intervals of 3200 and 3600 cm-1 are usually due to O–H stretching vibration

of alcohols, carboxylic acids, and hydroperoxides (Ogbu & Ajiwe, 2016). These bands range

is conspicuously absent at room temperature for the oil under study. This indicates the absence

of a free hydroxyl functional group (O–H) for alcohols, phenols, primary and secondary

oxidation products for the extracted oil, as shown in the FT-IR spectrum (Figure 4.3). Data

showed that triglyceride (TG) was the main component in this rubber seed oil. The strong

absorption band of the ester carbonyl functional group of TG, i.e. C=O was observed around

1742 cm-1. The stretch vibration of C–O at 1160 cm-1 is attributed to the presence of ester

groups. The FT-IR spectrum presents a fingerprint region (1461–585 cm-1) that can be used as

an analytical tool to detect rubber seed oil adulteration. Table 4.4 shows the functional groups

and modes of vibration present in this rubber seed oil at room temperature (Onoji et al., 2016b).

The results compared favourably and within the ranges reported for rubber seed oils by other

researchers (Reshad et al., 2015; Bakare et al., 2006).

142
400 900 1400 1900 2400 2900 3400 3900

wavelength (cm-1)

Figure 4.3: FT-IR spectrum of the rubber seed oil

143
Table 4.4: Wavenumbers and their functional groups for FT-IR analysis of the rubber seed oil

Wavenumber (cm-1) Functional group Description of vibration

3009 Alkenes This is due to =C–H stretching of non-conjugated unsaturation (methylene group)

2921–2853 Alkanes These strong peaks are attributed to the stretching of C–H (methyl group)

1742 Carbonyl This strong and sharp peak is attributed to –C=O stretching of esters

1710 Carboxylic acids This assigned peak is due to stretching vibration of C=O

Fingerprint region (1461–585 cm-1)

1461 Alkenes The medium signal can be attributed to bending frequency of C–H

1377 Alkenes The assigned peak is due to C–H bending vibrations of alkenes (CH2 group)

1238 Carboxylic acids, Esters The assigned peak is attributed to stretching vibration of C–O

1160 Esters C–O stretching vibration attributed to ester groups

1118–1033 Esters =C–O–C stretching vibration of ester groups

721 Aromatics Assigned to out-of-plane bending vibration of saturated carbon atom (C–H )

585 Alkanes The weak signal is due to C–H vibration

144
4.3.10.2 1H NMR analysis of the seed oil

The assignments of the signals of the rubber seed oil spectrum to the different kinds of

hydrogen atoms of the acyl chains based on the literature provided by Guillén & Ruiz (2003)

as in Table G.2 of Appendix G are presented in Table 4.5. The 1H NMR spectrum of the rubber

seed oil shown in Figure 4.4, with ten different signals of variably significant intensity

depending on the proportions of the different acyl groups, is comparable with those of most

vegetable oils reported in the literature (Guillén & Ruiz, 2003) and that of rubber seed oil

presented by other researchers (Reshad et al., 2015; Bakare et al., 2006; Okieimen et al.,

2005b).

The signals are due to the protons of the triacylglycerols and are proportional to the number of

hydrogen atoms of each kind that are present in the oil sample. Signal 1 (Figure 4.4 and Table

4.5) is produced by overlapping of doublet signals of methyl group protons of the saturated and

omega-9 (i.e. oleic acid) acyl groups, and it appears at 0.80 and 0.81 ppm, respectively. Signal

2 is a singlet due to methyl protons of erucyl group (i.e. cis-erucic acid) and appears at 0.89

ppm. Signal 3 is due to the protons of the saturated methylene group of all acyl chains and

appears at 1.19 and 1.23 ppm. Signals 4 and 6 are due to methylene protons in the β and α

position and appear at 1.53 and 2.22 ppm, respectively. Signal 5, between 1.94 and 1.98, are

due to α-methylene protons in relation to a single double bond (allylic protons). Signal 7, an

overlapping signal between 2.68 and 2.72 ppm, is due to signals from α-methylene protons in

relation to double bonds (bi-allylic protons). Signal 8, which appears between 4.05 and 4.23

ppm, is due to the protons on carbon atoms 1 and 3 of the glyceryl group. Signal 9, which

appears at 5.25, is due to the proton on carbon atom 2 of the glyceryl group that overlaps with

signal 10 (olefinic protons of different acyl groups) at 5.26 ppm.

145
Table 4.5: 1H NMR spectrum analysis of the rubber seed oil

Signal Chemical shift Functional group Assignments

(ppm)

1 0.80 and 0.81 –CH3 (terminal methyl protons (saturated and oleic acids) all acyl chains

except linolenyl

2 0.89 –CH3 (terminal methyl protons (cis-erucic acid) acyl chains

3 1.19 and 1.23 –(CH2)n - (saturated methylene proton) all acyl chains

4 1.53 –OCO–CH2–CH2– (β-methylene protons (carbonyl) acyl chains

5 1.94-1.98 –CH2–CH=CH (allylic methylene protons) all acyl chains

6 2.22 –OCO–CH2– (α-methylene protons) acyl chains

7 2.68 and 2.72 =HC–CH2–CH= (bis-allylic methylene protons) acyl chains

8 4.05-4.23 –CH2OCOR (methylene protons on carbons 1, 3) glyceryl group

9 5.25 >CHOCOR (proton on carbon atom 2). It overlaps with glyceryl group

signal 10

10 5.26 –CH=CH– (olefinic protons) acyl chains


1
H NMR chemical shift (2.8 ppm) (Sadowska et al., 2008) usually attributed to the presence of linolenic and

linoleic chains (CH=CHCH2CH=CH), is not present in rubber seed oil used in this study. This is in satisfactory

agreement with the GC-MS and 13C NMR analyses of the oil.

146
Figure 4.4: 1H NMR spectrum of the rubber seed oil

4.3.10.3 13C NMR analysis of the seed oil


13
The C NMR spectrum for the oil under study is depicted in Figure 4.5 and presented in

Figures G.1 through G.5 of Appendix G. The spectrum could be grouped into four distinct

spectra regions (A, B, C, and D) as presented in Table 4.6 (Onoji et al., 2016b). The results are

similar to previous studies by other researchers on rubber seed oil (Okieimen et al., 2005b).

When the 13C NMR is compared with spectra in Figure G.6 of Appendix G (Sadowska et al.,

2008), it shows the possible absence of linolenic and linoleic acids as confirmed by the data

from GC-MS analysis (Table 4.3).

147
Figure 4.5: 13C NMR spectrum of the rubber seed oil

Table 4.6: 13C NMR analysis of carbon type present in the rubber seed oil

Group Carbon environment Chemical shift, δ (ppm) Functional group

A Aliphatic 13.91–33.97 CH3

B Glyceryl 61.93–77.16 CH2–OCOR

C unsaturated alkenes 126.99–131.99 CH=CH

D carbonyl 172.41–172.82 C=O

carboxylic acid 178.64 COOH

4.4 Kinetic study of thermo-oxidative degradation of the seed oil

When vegetable oils are heated under extreme conditions experienced in frying or in industrial

applications, they undergo thermal oxidative reactions that are reflected in remarkable changes
148
in peroxide, iodine, and refractive index values, respectively (Oderinde et al., 2009). The

average values of the peroxide, iodine, and refractive index of the Hevea brasiliensis seed oil

used in this study with respect to temperature and time on heating obtained from Table E.1 of

Appendix E are summarized in Table 4.7 (Onoji et al., 2016b).

Table 4.7: Effect of heating on peroxide, iodine, and refractive index values of rubber

seed oil

Temp (oC) Time (min) Peroxide value Iodine value Refractive index
(mequiv O2/kg oil) (g I2/100 g oil) @ 20 oC
100 30 4.02 ± 0.25 117.72 ± 1.10 1.4651 ± 0.013
60 4.50 ± 0.12 117.43 ± 0.16 1.4322 ± 0.048
120 4.81 ± 0.01 117.24 ± 0.05 1.4147 ± 0.018
180 5.00 ± 0.22 117.02 ± 0.02 1.3514 ± 0.043
240 5.40 ± 0.28 116.98 ± 0.66 1.3011 ± 0.001
300 5.61 ± 0.15 116.84 ± 0.05 1.2430 ± 0.011
150 30 4.38 ± 0.04 117.56 ± 0.36 1.4630 ± 0.004
60 4.66 ± 0.08 117.32 ± 0.16 1.4290 ± 0.001
120 4.98 ± 0.12 117.12 ± 0.24 1.4130 ± 0.019
180 5.24 ± 0.05 116.84 ± 0.32 1.3850 ± 0.005
240 5.56 ± 0.35 116.60 ± 0.87 1.3651 ± 0.006
300 5.82 ± 0.16 115.40 ± 0.84 1.3436 ± 0.002
200 30 4.62 ± 0.12 117.38 ± 0.09 1.3647 ± 0.002
60 4.88 ± 0.08 117.02 ± 0.02 1.3230 ± 0.001
120 5.00 ± 0.12 116.61 ± 0.38 1.2991 ± 0.002
180 5.34 ± 0.05 116.32 ± 0.31 1.2911 ± 0.001
240 5.76 ± 0.08 115.68 ± 0.98 1.2834 ± 0.002
300 5.94 ± 0.04 115.24 ± 0.05 1.2433 ± 0.001
250 30 4.74 ± 0.02 117.21 ± 0.50 1.3020 ± 0.002
60 4.90 ± 0.07 116.60 ± 0.28 1.2750 ± 0.001
120 4.98 ± 0.11 116.20 ± 0.28 1.2331 ± 0.001
180 5.20 ± 0.07 115.40 ± 0.28 1.2311 ±0.001
240 5.50 ± 0.07 114.82 ± 1.32 1.2212 ± 0.005
300 5.70 ± 0.14 114.48 ± 0.22 1.2011 ± 0.010
values are mean ± standard deviation of duplicate data

This shows that rubber seed oil undergoes thermal degradation that results to oxidative

rancidity and formation of unstable hydroperoxides. Moisture present in rubber seed oil hastens

its oxidation process (hydrolysis) that generates free fatty acids (FFAs) and their oxidized

compounds as depicted in Figure 4.6.


149
triacylglycerol-oil free fatty acids glycerol acrolein
O
O CH2OC-R R COOH CH2OH CH2
heat heat
R`-COCH O (H2O) R`COOH + CHOH (O2) CH
CH2OC-R`` R`` COOH CH2OH HC=O
(toxic)
heat

free radicals Na, K, Ca, etc.

heat oxidation heat oxidation

dimers and polymers, soap formation oxidation products


increased viscosity (foaming) lactones,
methyl ketones,
aldehydes, etc.

browning (flavor and aroma)

browning

Figure 4.6: Overall process of oil degradation (Adapted from Onoji et al, 2016b)

During thermal hydrolysis, the oil decomposed to produce di- and mono-acylglycerol, glycerol,

and FFAs that accelerate further the hydrolysis of the oil. The major thermal degraded

byproducts of rubber seed oil are low molecular volatile compounds, such as aldehydes,

ketones, carboxylic acids, and short chain alkanes and alkenes (Tables 4.4, 4.5, and 4.6) and

non-volatile dimers and polymers. The rubber seed oil used in this study is rich in

monounsaturated oleic acids (64.11 wt.%) and could form dehydroxydimer, ketohydrodimer,

monohydrodimer, and dehydrodimer of oleic acid if defectively stored. Dimers and polymers

are large molecules with molecular weight range of 692-1600 Da. These compounds are formed

during thermal oxidation by a combination of –C–C–, –C–O–C–, and –C–O–O–C–bonds and

have hydroperoxy, epoxy, hydroxyl, carbonyl groups, and –C–O–C–C–, and –C–O–O–C–

linkages (Choe & Min, 2007). The aforementioned groups and linkages were present in the oil
150
used for this study as deduced from FT-IR and NMR results (Tables 4.4, 4.5, and 4.6). This

establishes the possibility of generating dimers and polymers from rubber seed oil at elevated

temperatures considered in the kinetic study. The mechanism for the formation of acyclic

polymers generated from oleic acid component of the rubber seed oil during thermal

degradation is shown in Figure 4.7. The degradation process also results in loss of unsaturation

(iodine value) in the fatty acids as shown in Table 4.7.

Figure 4.7: Acyclic polymer formations from oleic acid during thermal oxidation of oils

(Adapted from Onoji et al., 2016b)

The iodine data in Table 4.7 were used to generate the kinetic data for the thermo-oxidative

behaviour of the oil under heat treatment. The reaction follows a first order mechanism as

depicted in Figure 4.8. The double bonds in vegetable oils increase the refractive index and the

151
progressive decrease in refractive index on heating as seen in Table 4.7 confirms loss of

unsaturation.

Time (min)
0
0 100 200 300 400 ln (Ct/Co) at 100
deg.C
-0.005 ln (Ct/Co)150 deg. C
R² = 0.914

-0.01 ln (Ct/Co) 200 deg. C


ln (Ct / Co)

ln (Ct/Co) 250 deg. C


R² = 0.866
-0.015

R² = 0.990
-0.02

-0.025 R² = 0.981

-0.03

Figure 4.8: ln (Ct/Co) vs. time for iodine values at 100, 150, 200 and 250 oC (Adapted from

Onoji et al., 2016b)

The kinetic parameters for degradation of rubber seed oil at 100, 150, 200, and 250 oC are given

in Table 4.8. The Arrhenius plot (lnk vs. 1/T) for the degradation kinetics in the temperature

range 100-250 oC is shown in Figure 4.9 with coefficient of determination R2 equals 0.893.

The activation energy (Ea) and the pre-exponential factor determined are 13.07 kJ/mol and

161x10-3 s-1, respectively (Onoji et al., 2016b). The high activation energy suggests that the oil

is less susceptible to thermal deterioration than other vegetable oils, such as Hura crepitans

seed oil (Ea = 1.989 kJ/mol) (Oderinde et al., 2009). The proposed Arrhenius equation for the

kinetics of degradation of rubber seed oil in the temperature range 100-250 oC is given in

Equation (4.9).

152
k (s-1) = 161 x 10-3 exp (-13.07/RT) (4.9)

Table 4.8: Kinetic parameters for degradation of rubber seed oil

Parameters/Temperature 373 K 423 K 473 K 523 K

k (s-1) 2.0 x 10-3 5.0 x 10-3 6.0 x 10-3 7.0 x 10-3

Ea = 13.07 kJ/mol

Ao = 161 x 10-3 s-1

1/T (K-1)
-4.5
0.0017 0.0019 0.0021 0.0023 0.0025 0.0027 0.0029
-4.7

-4.9

-5.1
ln k

-5.3
y = –1572x –1.826
-5.5
R² = 0.893
-5.7

-5.9

-6.1

-6.3

-6.5

Figure 4.9: Arrhenius plot for kinetics of thermo-oxidative degradation of rubber seed oil

153
4.5 Parametric effects and optimization oil extraction using RSM

4.5.1 Full quadratic regression model

The RSM full quadratic regression model for the extraction of rubber seed oil using BBD based

on Equation (3.7) is given in coded extraction process parameters as (Onoji et al., 2017a):

Y (wt. %) = 36.66 + 1.71x1 + 0.77x2 + 3.69x3 – 4.48x1x2 – 3.28x1x3 – 2.88x2x3 + 0.11x12 –


1.24x22 + 0.96x32 (4.10)

The mathematical model generated by Design-Expert software trial version 8.0.3.1 (Stat-Ease

Inc., Minneapolis, MN, USA) was used for the analysis. The correlation between the predicted

and experimental yields data is depicted in Figure 4.10, and data obtained from Table 4.9.

154
44

42
Predicted oil yield (wt.%)

40
y = 0.9996x + 0.0152
38
R2 = 0.9996
36

34

32

30

28
28 30 32 34 36 38 40 42
Experimental oil yield (wt.%)

Figure 4.10: Correlation of predicted and experimental oil yields of rubber seed oil

extraction

155
Table 4.9: Box–Behnken design for the rubber seed oil extraction process optimization by RSM and ANN (Onoji et al., 2017a)

Std. x1 x2 x3 Oil yield (wt. %) Predicted yield (wt. %) Residual value


run RSM ANN RSM ANN
1 –1 –1 0 28.64 28.56 28.641 0.080 0.001195
2 1 –1 0 41.00 40.94 41.531 0.062 0.535070
3 0 0 0 36.44 36.66 36.656 –0.22 0.216070
4 0 0 0 36.64 36.66 36.656 –0.016 0.016071
5 –1 0 –1 29.00 29.04 29.000 –0.042 0.000305
6 1 0 –1 39.00 39.02 39.000 –0.024 0.000257
7 –1 0 1 43.00 42.98 43.000 0.024 0.000470
8 0 0 0 36.76 36.66 36.656 0.01 0.103930
9 0 –1 –1 29.00 29.04 29.000 –0.038 0.000497
10 0 1 –1 36.44 36.34 36.440 0.10 0.000012
11 0 –1 1 42.06 42.16 41.529 –0.10 0.530890
12 0 1 1 38.00 37.96 38.000 0.038 0.000005
13 –1 1 0 39.00 39.06 39.000 –0.062 0.000093
14 1 1 0 33.45 33.53 33.450 –0.080 0.000063
15 1 0 1 39.88 39.84 39.883 0.042 0.000197
16 0 0 0 36.68 36.66 36.656 0.024 0.023929
17 0 0 0 36.76 36.66 36.656 0.10 0.10393
X1: rubber seed powder weight (g), X2: solvent volume (mL), X3: extraction time (min)

156
The analysis of the regression coefficients showed that the linear terms of x1, x2 and x3 were

positive, whereas the quadratic term of x2 was negative. This means that the oil yield was

enhanced by X1, X2 and X3. However, at a certain level of solvent and with more seed powder

weight, the oil yield decreases as the solvent becomes saturated. The curvature of the surface

response was significantly determined by x22. The coefficient of determination (R2) of the

model in predicting the yield was 0.9996. The residual values (Table 4.9) for the RSM in each

case were found to be less than unity and confirm that the R2 value for the model is close to 1.

4.5.2 Optimization of oil extraction process variables based on RSM

The BBD within the experimental range of data for the independent process parameters,

predicted optimal rubber seed oil yield of 42.98 wt. % at the following conditions: rubber seed

powder weight (RSPW) 60 g, solvent volume 250 mL, and extraction time 45 min. Using these

predicted values, the experiment was validated in triplicate and the average optimum rubber

seed oil yield was 42.64 wt. % (Section D.2 of Appendix D). This is close to the model

predicted value (within ± 0.797 % relative error) at same optimum conditions based on

Equation (4.11) (Zahedi & Azarpour, 2011). The absolute average deviation (AAD) obtained

from Table 4.9 and based on Equation (4.12) (Betiku & Ajala, 2014) was ±0.385 %. Evaluation

of R2 and AAD values together was used to check the accuracy of the model (Baş & Boyaci,

2007b). The low value of AAD and R2 = 0.9996 (close to unity) confirms the validity of the

model in predicting the oil. Hence, the regression model gave a good estimate of the response

of the process for low values obtained.

𝑒𝑥𝑝𝑒𝑟𝑖𝑚𝑒𝑛𝑡𝑎𝑙 𝑣𝑎𝑙𝑢𝑒−𝑝𝑟𝑒𝑑𝑖𝑐𝑡𝑒𝑑 𝑣𝑎𝑙𝑢𝑒


𝑃𝑒𝑟𝑐𝑒𝑛𝑡 𝑟𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝑒𝑟𝑟𝑜𝑟 = × 100 (4.11)
𝑒𝑥𝑝𝑒𝑟𝑖𝑚𝑒𝑛𝑡𝑎𝑙 𝑣𝑎𝑙𝑢𝑒

𝑌𝑖,𝑒𝑥𝑝 −𝑌𝑖,𝑝𝑟𝑒𝑑 1
𝐴𝐴𝐷 = {[∑𝑛𝑖=1 ( )] 𝑛} × 100 (4.12)
𝑌𝑖,𝑒𝑥𝑝

157
Where n is the number of experimental runs, while Yi, exp is the experimental oil yield

and Yi, pred is the predicted oil yield.

4.5.3 Analysis of variance (ANOVA) study for seed oil extraction


The statistical significance of the quadratic regression model equation, the individual factors,

their interactions and the goodness of the fit by analysis of variance (ANOVA) are shown in

Table 4.10. The quality of the model fit was evaluated by the Fisher's test (F-value), the

probability value (p-value), the 'lack of fit', the coefficient of determination (R2) and adjusted

coefficient of determination (R2adj). The p-value should be less than 0.05 for the variables to

have a significant effect on the response value - in this case, RSO yield (Onoji et al., 2017a).

158
Table 4.10: Analysis of variance (ANOVA) for response surface quadratic model

Source of variance Sum of squares Degrees of freedom Mean square F-value p-value

Model 303.18 9 33.69 1980.33 <0.0001

RSPW, x1 23.44 1 23.44 1377.89 <0.0001

Solvent, x2 4.79 1 4.79 281.56 <0.0001

Extraction time, x3 108.81 1 108.81 6396.43 <0.0001

x1x2 80.19 1 80.19 4714.29 <0.0001

x1x3 43.01 1 43.01 2528.56 <0.0001

x2x3 33.06 1 33.06 1943.66 <0.0001

x12 0.047 1 0.047 2.79 0.1388

x22 6.47 1 6.47 380.39 <0.0001

x32 3.87 1 3.87 227.49 <0.0001

Residual 0.12 7 0.017

Lack of fit 0.05 3 0.017 0.96 0.4917

Pure error 0.069 4 0.017

Corrected total 303.29 16

RSM: R2 = 0.9996, R2adj. = 0.9991, Root mean square error (RMSE) = 0.088,

Mean = 36.57, Standard deviation = 0.13, Adequate precision = 144.2,

Absolute average deviation (AAD) = ± 0.385 %, % Coefficient of variation = 0.355,

F- Fisher; p- probability.

The quadratic regression model has an F-value of 1980.33 and a p-value <0.0001, which

indicates that the derived model is statistically significant at the 95% confidence level. The

higher the F-value for the specific independent process parameter, the higher will be the effect

of that variable. The main effects, i.e. x1, x2 and x3 are significant terms based on their calculated

p-values (<0.0001). The quadratic effect terms of x22 and x32 are significant, but x12 term is not

significant (p>0.05), with a small F-value of 2.79. The cross-product terms of x1x2, x1x3 and

x2x3, are all significant. In Table 4.11, 95% confidence levels (low and high) represent the range

159
that the true coefficients are found in 95% of the time. If one limit is positive and the other is

negative, then the coefficient zero could be true, which indicates that the factor has no effect.

In this study, x12 has no significant effect on RSO yield having both negative and positive

values as shown in Table 4.11. However, even though x12 has a p-value >0.05 (Table 4.10),

indicating that it has no significant effect on the RSO yield, the magnitude of the change caused

is insignificant; hence, it was retained in the regression model Equation (4.10). In general, if

there is large number of insignificant model terms, model reduction by reducing the ranges of

parameters may improve the model (Dharma et al., 2016). If a model displays a 'lack of fit', it

means that several important terms were excluded from the model, or the presence of unusually

large residuals arising from fitting the model (Dharma et al., 2016). In that case, the model

does not sufficiently describe the relationship between the independent process variables and

the response variable. In this study, the 'lack of fit' of low F-value of 0.96 and high p-value of

0.4917 implied that the model was not significant relative to pure error (0.069). The p-value of

the 'lack of fit' parameter is greater than 0.05, indicating that there is good fit between the

quadratic regression model and the experimental oil yield obtained (Dharma et al., 2016). The

R2 reflects the variability of the dependent variable, which is explained by its relationship with

the independent process variables. In this study, the R2 value is 0.9996, which indicates that

99.96% of the variability in the rubber seed oil yield is attributed to the independent factors

considered (rubber seed powder weight, solvent volume and extraction time), and only 0.04%

was not explained by the model (Onoji et al., 2017a). The R2adj value is 0.9991 and the root

mean squared error (RMSE) is 0.088. These indicate that the estimated model fits the

experimental data satisfactorily, as suggested by Lee et al. (2010). The low values of standard

error observed in the intercept and all the model terms (Table 4.11) confirm that the regression

model equation fits the data well, with acceptable prediction. The variance inflation factor

(VIF) values shown in Table 4.11 indicate that the center points are orthogonal to all other

160
factors in the model. The signal-to-noise ratio is measured by the statistical parameter called

'adequate precision'. In general, an adequate precision >4 is desirable for a good model

(Dharma et al., 2016). In this study, the calculated adequate precision is 144.2 (Table 4.10),

indicating that the signal is adequate and the quadratic model can be used to navigate the

designed space. In addition, the 'coefficient of variation' is low (0.355%), which indicates that

the experimental data are accurate and reliable.

Table 4.11: RSM regression analysis for rubber seed oil extraction

Factor Regression Degrees of Standard error 95% Cnf. level 95% Cnf. level VIF
coefficient freedom (low) (high)
Intercept 36.66 1 0.058 36.52 36.79 1.00

x1 1.71 1 0.046 1.60 1.82 1.00

x2 0.77 1 0.046 0.66 0.88 1.00

x3 3.69 1 0.046 3.58 3.80 1.00

x1x2 -4.48 1 0.065 -4.63 -4.32 1.00

x1x3 -3.28 1 0.065 -3.43 -3.12 1.00

x2x3 -2.88 1 0.065 -3.03 -2.72 1.00

x12 0.11 1 0.046 -0.044 0.26 1.01

x22 -1.24 1 0.046 -1.39 -1.09 1.01

x32 0.96 1 0.046 0.81 1.11 1.01

Cnf: confidence; VIF: variance inflation factor

4.5.4 Parametric effects of interactive factors on the seed oil yield


Figure 4.11 (a) shows the response surface plot that indicates the influence of solvent volume

and rubber seed powder weight on the extraction yield when extraction time was held constant

at 45 min. It was observed that a higher seed powder weight significantly increases the oil yield

at low solvent volume. However, a slight reduction in oil yield was observed when a higher

solvent volume and low seed powder weight was used.

161
are

e predicted value
w predicted value

44.1124
50
owder weight (g)
(ml) 45
R S O

) = 50.00 40

35
%

30

25

200.00
40.00
212.50 45.00
225.00 50.00

237.50 55.00
Solvent volume (ml) Rubber seed powder weight (g)
250.00 60.00

Figure 4.11 (a): Interactive effects between solvent volume and rubber seed powder

weight on oil yield for RSM response surface (3D) plot

Figure 4.11 (b) depicts the response surface plot on the mutual interaction between extraction

time and rubber seed powder weight at constant solvent volume (225 mL). It was observed that

oil yield increased significantly with extraction time at low rubber seed powder weight. This

may be the result of more interaction between rubber seed weight and solvent (Onoji et al.,

2017a). This agrees with the findings of Reshad et al. (2015). Nevertheless, a reduction in oil

yield was observed at higher rubber seed weight and low extraction time.

162
are

44.1124
owder weight (g) 50
(min)

45
= 250.00
R S O

40

35
%

30

25

40.00 40.00
42.00
45.00
44.00
50.00
46.00
Extraction time (min) 48.00 55.00
Rubber seed powder weight (g)
50.00 60.00

Figure 4.11 (b): Interactive effects between extraction time and rubber seed powder

weight on oil yield for RSM response surface (3D) plot

Figure 4.11 (c) describes surface plot for the relationship between extraction time and solvent

volume when the rubber seed powder weight was held constant at 50 g. A significant increase

in oil yield was observed at low solvent volume and high extraction time. However, a marginal

reduction (9.6%) in oil yield was observed when solvent volume and extraction time are both

at high levels.

163
re

e predicted value 44.1124


w predicted value

50
(ml)
(min) 43
R S O

r weight (g) = 40.00 36

29
%

22

15

250.00
50.00
48.00 237.50
46.00 225.00
44.00
42.00 212.50 Solvent volume (ml)
Extraction time (min) 40.00 200.00

Figure 4.11 (c): Interactive effects between extraction time and solvent volume on oil yield

for RSM response surface (3D) plot

4.6 Parametric effects and optimization of oil extraction using ANN model

The optimum number of neurons in the hidden layer that gave good prediction of the outputs

of both training and cross validation sets in this study was iteratively obtained as three. The

genetic algorithm (GA) of the model predicted rubber seed oil yield of 43 wt. % at the following

conditions: rubber seed powder weight 40 g, solvent volume 202 mL, and extraction time 49.99

min. Using these values, the experiment was validated in triplicate, and average optimum

rubber seed oil yield was 42.96 wt. % (Section D.3 of Appendix D). This is close to the model

predicted value (within ±0.093 % relative error) at same extraction conditions. The model was

found to have an R2 value of 0.9982, R2adj value of 0.9964, RMSE of 0.2054, and AAD of

±0.064%. Figure 4.12 compares the predicted and experimental values of the rubber seed oil

164
yield, and the R2 obtained (0.997) shows a good correlation between the data obtained. This

confirms that the model proved suitable for adequate representation of the actual relationship

among the selected factors.

0.044
0.042
Predictrd yield (wt. %)

0.04 y = 0.9979x + 0.0762


0.038
R2 = 0.9979
0.036
0.034
0.032
0.03
0.028
28 30 32 34 36 38 40 42
Experimental yield (wt. %)

Figure 4.12: Correlation of predicted and experimental rubber seed oil yields by ANN

model

Figures 4.13 (a–c) depicts the interactive effects on oil yield between two extraction

independent variables while maintaining the third variable at constant level. The curvature

nature of the 3D surfaces suggests significant interactions of solvent volume with rubber seed

powder weight, extraction time with rubber seed powder weight, and solvent volume with

extraction time. It was also observed that the contour plots obtained by the ANN model were

very similar to those obtained by the RSM model.

165
Figure 4.13 (a): Interactive effects of rubber seed powder weight and solvent volume on

oil yield for ANN response surface (3D) and contour (2D) plots

166
Figure 4.13 (b): Interactive effects of rubber seed powder weight and extraction time on

oil yield for ANN response surface (3D) and contour (2D) plots.

167
Figure 4.13 (c): Interactive effects of extraction time and solvent volume on oil yield for

ANN response surface (3D) and contour (2D) plots

168
The level of importance of the process variables considered for the rubber seed oil extraction

is shown in Figure 4.14, and it is apparent that solvent volume of 50.06% is the most significant

variable on rubber seed oil yield, followed by rubber seed powder weight of 42.47% and lastly,

extraction time of 7.471%.

Figure 4.14: Percentage contribution of process variables to seed oil extraction via ANN

model (solvent volume 50.06%; RSPW 42.47%; extraction time 7.471%)

4.7 Comparison of RSM and ANN models for the seed oil extraction
The accuracies of both RSM and ANN models were evaluated based on the values obtained

for R2, AAD and % relative error. The results show that both optimization techniques gave good

predictions, due to the values of R2 that are relatively close to unity (i.e. 0.9996 and 0.9982 for

RSM and ANN, respectively). However, ANN proved to be a superior model, with lower

values of % relative error (±0.093 %) and AAD (±0.064 %) compared to the RSM model with

relative error ±0.797 % and AAD ±0.385 % (Onoji et al., 2017a).

169
4.8 Catalyst preparation and characterization

This section discusses the preparation of the catalyst derived from rubber seed shells, and the

characterization of the catalyst for biodiesel production.

4.8.1 Thermo-gravimetric analysis of the rubber seed shell catalyst


To study the influence of calcination temperature on weight loss (%), the powdered RSS was

subjected to thermal analysis (30–950 oC) in a nitrogen atmosphere at 10 oC/min heating rate.

The thermogram of RSS in Figure 4.15 shows that less water was evaporated from the surface

of the catalyst in the temperature range 30–348 oC. This indicates that less water moisture from

air was adsorbed on the surface of the catalyst than for the common heterogeneous catalysts

and it confirms its superior hydrophobicity (Onoji et al., 2017b)

120

100
Mass loss (%)

80

60

40

20

0
0 100 200 300 400 500 600 700 800 900
Temperature (oC)

Figure 4.15 TGA profile for the rubber seed shell (RSS) catalyst

The result also shows that the decomposition temperature of RSS catalyst was about 348 oC.

This confirms the stability of the catalyst for the transesterification reaction. The TGA result

from this study is similar to the TGA curve for RSS by other researchers (Sun & Jiang, 2010).

170
Major mass loss (about 90%) was observed between 348 and 465 oC, which resulted from

decomposition of macromolecules and removal of organic compounds such as cellulose and

hemicelluloses (348–400 oC), and lignin (220–465 oC). At temperature range 465-800 oC, an

additional small and steady mass loss was also observed. This may be attributed to the

formation of gaseous components because of decomposition process. Figure 4.15 also shows

that the RSS calcined at 800 oC or above did not indicate any further mass loss, which means

that the composition of the calcined RSS was mainly CaO and SiO2 as observed from XRD

results. It also confirms the high thermal stability of the calcined RSS at 800 oC and above.

Based on TGA results, 800 oC was chosen as the calcination temperature of RSS to produce

the catalyst (Onoji et al., 2017b).

4.8.2 XRF elemental analysis and basic property of rubber seed shell catalyst

XRF spectra of raw and calcined RSS are depicted in Figure 4.16 (a & b). The elemental

compositions are presented in Table H.1 of Appendix H. The pH for raw and calcined (800 oC)

RSS was found to be 7 and 12.5, respectively. This confirms the basic nature of the derived

catalyst. Background adjustment of peaks for trace metals such as niobium, molybdenum, tin,

and antimony brought them within the range. The results show that the major elements present

are calcium, silicon, potassium, aluminum, and iron. Trace metals such as zinc, tungsten,

cobalt, manganese chromium, tin and antimony are present in minute quantities after

calcination. The absence of heavy metals such as cadmium, arsenic, titanium, and lead confirms

that the catalyst synthesized is non-toxic and environmentally friendly.

171
Figure 4.16: XRF analysis of RSS: (a) raw and (b) calcined at 800 oC (Adapted from Onoji

et al., 2017b)

4.8.3 XRD analysis of the rubber seed shell (RSS) catalyst

The XRD diffractograms of the raw and calcined (800 oC) RSS are depicted in Figure 4.17 (a

& b). As shown in Figure 4.17 (a), a wide diffraction peak centered at 2θ = 23o was observed

in the XRD pattern of the raw RSS that suggests its amorphous structure (D'Cruz et al., 2007).

The diffractograms of the calcined RSS show that the catalytic property was improved by the

presence of lime (CaO) and quartz (SiO2), which enhances its thermal stability. The analysis
172
of the crystal using XRD illustrates that the raw RSS is made of CaCO3 and quartz. During

calcination, CaCO3 present in the raw RSS decomposes to CaO, and quartz to SiO2. The

diffractogram patterns of the calcined RSS are similar to the diffractograms CaO nano-particles

reported in the literature (Tang et al., 2008). The CaO generated from RSS during calcination

acts as a solid base catalyst for transesterification of rubber seed oil to biodiesel. CaO has been

utilized as a base catalyst for soybean oil transesterification to biodiesel (Liu et al., 2008). The

mechanisms of CaO as a base catalyst for oil transesterification are well described in the

literature (Issariyakul & Dalai, 2014; Guo & Fang, 2011; Liu et al., 2008). In the presence of a

little water, CaO generates methoxide anions that act as the real catalyst in transesterification

reactions (Liu et al., 2008).

173
Figure 4.17: XRD diffractograms of RSS: (a) raw, and (b) calcined at 800 oC (Adapted

from Onoji et al., 2017b)

4.8.4 SEM analysis of the rubber seed shell catalyst

Scanning electron micrographs (SEM) of the raw and calcined rubber seed shell (RSS) are

presented in Figure 4.18 (a & b). The morphology of the raw RSS looks like a mass aggregate

that has less surface area as compared to the RSS calcined at 800 oC, that shows major alteration

in its morphology. There is an observed maximum particle size reduction in the calcined RSS,

thereby exhibiting a higher surface area, an important characteristic of heterogeneous catalysts.

The SEM image of calcined RSS clearly shows regular particle shapes with clusters of well-

developed cubic crystals with obvious reduction in particle sizes, indicating higher surface area

useful for transesterification reaction (Onoji et al., 2017b).

174
a b

Figure 4.18: SEM images for (a) raw, and (b) calcined RSS catalyst at magnification

2450x

4.8.5 Surface properties of the rubber seed shell catalyst

The physical properties of the raw and calcined catalyst at 800 oC, such as specific surface area,

pore volume, and pore-size distribution, were obtained by N2 adsorption/desorption at 77.3 K

using Gold App surface area and porosity analyzer (Model: V-sorb 2800p). All samples were

outgassed prior to the measurement at 250 oC and 10-4 mbar for 11 h. The adsorption-desorption

isotherms for the raw and calcined RSS are depicted in Figure 4.19 (Onoji et al., 2017b). The

adsorption isotherm at low P/Po for the calcined RSS is similar to type I isotherms that are

characteristic of microporous adsorbent. At higher P/Po, there was increase in the volume

adsorbed caused by adsorption in mesopores as well as a hysteresis loop (0.5–0.9 P/Po)

developed that is typical of type IV isotherms caused by a weak capillary condensation below

the expected condensation pressure of nitrogen (Storck et al., 1998).

175
140
Quantity adsorbed (cm3/g, STP)

120

100

80 Adsorption (calcined)

Desorption (calcined)
60
Adsorption (raw)

40 Desorption (raw)

20

0
0.0 0.2 0.4 0.6 0.8 1.0 1.2

P/Po

Figure 4.19: Nitrogen adsorption/desorption isotherms for raw and calcined RSS

The nitrogen adsorption/desorption isotherm log plot (Figure H.1 of Appendix H) presents an

"S" shaped isotherm in the low P/Po region, reaffirming the presence of micropores and

mesopores in the calcined RSS. The nitrogen adsorption/desorption isotherm log plot for the

raw RSS displayed a type III isotherm which is typical of multilayer adsorption in non-porous

solids as shown in Figure H.2 of Appendix H. Brunauer-Emmett-Teller (BET) estimated the

surface area of the raw and the calcined RSS as 21.83 and 352.51 m2/g, respectively. The high

surface area for the calcined RSS clearly shows its superiority over the raw RSS in catalyst

formulation. The pore size for the raw and calcined RSS is 3.61 and 2.29 nm, respectively. The

pore volume for the calcined RSS was estimated at 0.1208 cm3/g, while the raw RSS was a

non-porous material (Onoji et al., 2017b). The increase in BET surface area and pore volume

correlate with the mass loss steps in the TGA curve and probably due to a modification of the
176
sample composition during calcination and crystal growth of calcium oxide (Syazwani et al.,

2015). The derived catalyst has an estimated micropore and external surface areas of 250.71

and 101.8 m2/g, respectively. The surface area of the calcined RSS catalyst, which depends on

the calcination temperature, relates to its catalytic activity. The mesopore-size distribution

analyzed by the Barrett-Joyner-Halenda (BJH) model is depicted in Figure H.3 of Appendix

H. The summary of the nitrogen adsorption/desorption results for the raw and calcined RSS are

shown in Tables 4.12 and 4.13, respectively (Onoji et al., 2017b). The single point surface area

at P/Po = 0.3139 was estimated at 334.23 m2/g for the calcined catalyst, and 8.53 m2/g for raw

RSS at P/Po = 0.2589. The results of the N2 adsorption-desorption listed in Table 4.13 indicate

that RSS catalyst provided a relatively large contact area with reactants that enhances the yield

of the biodiesel produced.

177
Table 4.12: Summary report of N2 adsorption/desorption analysis for raw RSS

S/N Items Description Results

Surface area
1 Single point surface area P/Po = 0.258 8.525 m2/g
2 BET surface area Point range: 0.053–0.206 21.838 m2/g
3 Langmuir surface area Monolayer adsorption model calculations 62.401 m2/g
4 t-plot micropore area A: Harkins-Jura adsorbed thickness (nm): 0.000 m2/g
0.351–0.540
5 t-plot external surface area SBET - Smicro 21.837 m2/g
6 BJH adsorption cumulative Pore width (nm): 2.092–587.658 18.965 m2/g
surface area
7 BJH desorption cumulative Pore width (nm): 2.053–587.658 24.832 m2/g
surface area

Pore volume
1 Single point adsorption P/Po = 0.996, total pore volume of the critical pore 0.019 cm3/g
total pore volume width less than 587.658 nm
2 t-plot micropore volume - 0.000 cm3/g
3 SF micropore volume P/Po = 0.080, total pore volume of the critical pore 0.001 cm3/g
width less than 1.914 nm
4 BJH adsorption cumulative Pore width range (nm): 2.092–587.658 0.024 cm3/g
volume
5 BJH desorption cumulative Pore width range (nm): 2.053–587.658 0.026 cm3/g
volume

Pore size
1 Total adsorption average pore By 4V/A, A: adsorption BET specific surface area 3.607 nm
width value
2 BJH adsorption average pore By 4V/A, A: adsorption cumulative pore surface 5.070 nm
width area
3 BJH desorption average pore By 4V/A, A: desorption cumulative pore surface 3.872 nm
width area
4 BJH median pore width Pore width range (nm): 2.226–587.658 2.500 nm
5 SF median pore width Pore width range (nm): 1.119–1.914 1.510 nm

178
Table 4.13: Summary report of N2 adsorption/desorption analysis for calcined (800 oC)

rubber seed shell catalyst

S/N Items Description Results

Surface area
1 Single point surface area P/Po = 0.313 334.229 m2/g
2 BET surface area Point range: 0.061–0.242 352.518 m2/g
3 Langmuir surface area Monolayer adsorption model calculations 489.503 m2/g
4 t-plot micropore area A: Harkins-Jura adsorbed thickness (nm): 250.713 m2/g
0.382–0.549
5 t-plot external surface area SBET - Smicro 101.804 m2/g
6 BJH adsorption cumulative Pore width (nm): 1.959–675.818 62.630 m2/g
surface area
7 BJH desorption cumulative Pore width (nm): 2.104–587.658 60.893 m2/g
surface area

Pore volume
1 Single point adsorption P/Po = 0.997, total pore volume of the critical pore 0.201 cm3/g
total pore volume width less than 675.818 nm
2 t-plot micropore volume - 0.120 cm3/g
3 SF micropore volume P/Po = 0.096, total pore volume of the critical pore 0.156 cm3/g
width less than 3.053 nm
4 BJH adsorption cumulative Pore width range (nm): 1.959–675.818 0.060 cm3/g
volume
5 BJH desorption cumulative Pore width range (nm): 2.104–675.818 0.058 cm3/g
volume

Pore size
1 Total adsorption average pore By 4V/A, A: adsorption BET specific surface area 2.290 nm
width value
2 BJH adsorption average pore By 4V/A, A: adsorption cumulative pore surface 3.855 nm
width area
3 BJH desorption average pore By 4V/A, A: desorption cumulative pore surface 3.965 nm
width area
4 BJH median pore width Pore width range (nm): 2.226–587.658 2.115 nm
5 SF median pore width Pore width range (nm): 1.119–1.914 0.654 nm

179
4.9 Biodiesel production from the extracted seed oil and analysis

This section discusses the characterization of the produced biodiesel, optimization of the

biodiesel process variables and testing of the biodiesel in an internal compression ignition

engine.

4.9.1 Characterization of the rubber seed oil biodiesel

The biodiesel was characterized by following the procedures outlined by ASTM, AOAC

(1990), and the results are presented in Table 4.14. The properties compared favourably with

similar reports on biodiesel synthesized from Hevea brasiliensis oil (rubber seed) (Dhawane et

al., 2015), and they were in agreement with the ASTM D 6751 and EN 14214 biodiesel

standards. From the results, the methyl ester content of the biodiesel determined by GC-MS

was 96.7%, while the calculated value using the correlation developed by Felizardo et al.

(2006) was 96.9%. The results show that biodiesel produced from rubber seed oil using waste

rubber seed shell as catalyst could be used in modern diesel engines without technical

modifications.

180
Table 4.14: Fuel properties of biodiesel produced from rubber seed oil
Properties Methods ASTM D 6751 standards EN 14214 standards Literature values This study

Density @ 15 oC (kg/m3) ASTM D 1298 870–900 860–900 885a, 842b, 880c 876

Water and sediment (vol %) ASTM D 2709 < 0.05 < 0.05 0.042a 0.0062

Acid value (mg KOH/g) ASTM D 664 < 0.8 < 0.5 0.42a, 0.12b, 0.06c 0.56

Iodine value (g I2/100 g) Wijs' 120 maximum 85.34

Saponification value (mg KOH/g) 182.53

Kinematic viscosity @ 40 oC (mm2/s) ASTM D 445 1.9–6.0 3.5–5.0 3.89a, 5.642b, 4.91c 4.32

Flash point (oC) ASTM D 93 93 minimum 120 minimum 152a, 183b, 151c 158

Fire point (oC) 189b 172

Cloud point (oC) ASTM D 2500 –3 to 12 3.2a, 3b, –7c 4.8

Pour point (oC) ASTM D 97 –15 to 10 0 –2a, –6b, –10c –8

Cold filter plugging point (oC) ASTM D 6371 –2a –0.62

Calorific value (MJ/kg) ASTM D 240 39.70a, 42.372b 40.67

Oxidation stability: @ 110 oC (h) Rancimatd ≥3 ≥6 8.54a 7.8

@ 140 oC (min) PetroOXYe ≥ 17 21.55

Cetane number ASTM D 613 47 minimum 51 minimum 54a, 53b 57

Metals: Group II (Ca-ppm) EN 14538 5 3.26

Ester content (%) EN 14103 ≥ 96.5 96.8a 96.7

aAhmad et al. (2014b); b Dhawane et al. (2015); c Dhawane et al. (2016); d EN 14112; e ASTM D 7545-1

181
4.9.2 Reusability of the catalyst for biodiesel production

As shown in Figure 4.20 and Table H.2 of Appendix H, the biodiesel yield (%) at optimum

catalyst loading of 2.2 g was higher than 80% when the catalyst was used in the fourth cycle

and the value decreased to 77.3% in the fifth cycle of usage. The possible reason for the loss

of catalytic activity was the blockage of active sites by the deposition of organic matter on the

catalyst surface, and the leaching of the active metals from the catalyst into the reaction medium

(Dhawane et al., 2016). This was facilitated by the breakage of bonds and the creation of CH3O-

and Ca2+ ions. The AAS analysis of the biodiesel shows that the concentration of Ca2+ ion was

3.26 ppm (mg/kg). This value is within the limits of 5 mg/kg set by ASTM D 6751 and EN

14214 biodiesel standards. The effects of leaching of Ca2+ ions from the catalyst on the

biodiesel are elucidated in Figure 4.21.

100
2 2.2 2.5 3 3.5
90

80

70
Biodiesel yield (%)

60

50

40

30

20

10

0
Fresh 1st cycle 2nd cycle 3rd cycle 4th cycle 5th cycle
Catalyst cycle

Figure 4.20: Effects of reuse of different catalyst loadings (2, 2.2, 2.5, 3, and 3.5 g)

on biodiesel yield (%) at optimum conditions

182
86 400
85 350

Ca2+ content of catalyst (ppm)


84
300
Biodiesel yield (%)

83
82 250

81 200
80 150
79
100
78 Biodiesel yield (%)
Ca content (ppm)
77 50

76 0
0 1 2 3 4 5 6
Catalyst cycle

Figure 4.21: Effects of Ca2+ ion leaching from reused catalyst on biodiesel yield (%) at

optimum conditions

4.10 Parametric effects and optimization of biodiesel production process variables

The optimization of the process variables for the biodiesel production, and the parametric

effects of these variables on the yield of biodiesel as described by RSM and ANN models are

discussed in this section.

4.10.1 Modeling and optimization of biodiesel production by RSM

The results of the transesterification step based on CCD as depicted in Table 4.15 were

statistically analyzed by RSM. Design-Expert® dx8 software version 8.0.7.1(Design Ease Inc.,

USA) generated the quadratic mathematical model. The full quadratic regression model for

rubber seed oil methyl ester (RSOME) production in coded process variables is given in

Equation (4.13).

183
Biodiesel yield (%) = 83.46 + 0.008629x1 + 0.21x2 – 0.34x3 + 0.26x1x2+ 0.21x1x3 +

0.16x2x3 – 0.11x12 – 0.34x22 – 0.30x32 (4.13)

A good correlation exists between the predicted yields and the experimental yields as shown

in Figure 4.22. The coefficient of determination (R2) given as 0.918 indicates a well-fitted and

reliable model.

184
Table 4.15: Central composite design for optimization of biodiesel production process by RSM and ANN

Std. x1 x2 x3 biodiesel Predicted biodiesel yield (%) Residual


run yield (%) RSM ANN RSM ANN
1 –1 –1 –1 83.50 83.50 83.50 0.033 0.00000046
2 1 1 –1 82.50 82.50 82.50 –0.044 0.00000005
3 –1 1 –1 83.00 83.04 83.00 –0.044 0.00000001
4 1 1 –1 83.20 83.16 83.20 0.039 0.00000012
5 –1 –1 1 82.00 82.04 82.00 –0.044 0.00000004
6 1 –1 1 82.00 81.96 82.00 0.039 0.00000001
7 –1 1 1 82.30 82.26 82.30 0.039 0.00000033
8 1 1 1 83.18 83.22 83.18 –0.039 0.00000055
9 –2 0 0 83.00 83.00 83.00 0.0023 0.00000021
10 –2 0 0 83.00 83.00 83.00 0.0023 0.00000021
11 0 –2 0 82.00 81.70 81.70 0.30 0.30000000
12 0 –2 0 81.40 81.70 81.70 –0.30 –0.30000000
13 0 0 –2 82.90 82.95 82.95 –0.047 –0.05000000
14 0 0 –2 83.00 82.95 82.95 0.053 0.05000000
15 0 0 0 83.19 83.46 83.458 –0.27 –0.26833000
16 0 0 0 83.56 83.46 83.458 0.10 0.10167000
17 0 0 0 83.70 83.46 83.458 0.24 0.24167000
18 0 0 0 83.20 83.46 83.458 –0.26 –0.25833000
19 0 0 0 83.20 83.46 83.458 –0.26 –0.25833000
20 0 0 0 83.90 83.46 83.458 0.44 0.44167000

Coded values x1: reaction time (min), x2: methanol/oil ratio (v/v), x3: RSS catalyst (g)
185
84.0

83.5

83.0
Predicted (%)

82.5
y = 0.923x + 6.355
82.0 R2 = 0.918

81.5

81.0
81.0 81.5 82.0 82.5 83.0 83.5 84.0 84.5
Actual yield (%)

Figure 4.22: Correlation of predicted vs. actual biodiesel yields (%) for the RSM model

The residual values for RSM indicated in Table 4.15 were in each case, less than unity, and this

confirms the closeness of R2 to unity. The analysis of regression coefficients (Table 4.16) shows

that the linear terms of x1 and x2 are positive, while x3 is negative. The yield of RSOME (biodiesel)

is enhanced by increase in X1 and X2 and reduction in X3. The negativity of the coefficients of

quadratic terms x12, x22, and x32 suggests that the curvature effect of the surface response is

determined by these terms.

The derived model predicted an optimal rubber seed oil methyl ester yield of 83.11% at optimum

conditions of 60 min reaction time (X1), 0.20 vol/vol methanol/oil ratio (X2), and 2.50 g catalyst

(X3). The experiment was validated in triplicate using the optimum conditions to obtain an average

RSOME yield of 83.06±0.013%. This value is close to the model predicted value, and within

±0.06% relative error based on Equation (3.27). The absolute average deviation (AAD) calculated

from Table 4.15 and based on Equation (3.28) was ±0.022%. Hence, the developed model gave a

good estimate of the response of the process to the process variables for low values obtained.

186
Table 4.16: RSM regression coefficients analysis for the biodiesel

Factor Coefficient of Degree of freedom standard error 95%confidence 95%confidence Variance inflation
Estimate level (low) level (high) factor
Intercept 83.46 1 0.10 83.23 83.69 –

x1 0.008629 1 0.087 –0.18 0.20 1.72

x2 0.21 1 0.087 0.015 0.40 1.72

x3 –0.34 1 0.087 –0.53 –0.15 1.72

x1x2 0.26 1 0.091 0.057 0.46 1.00

x1x3 0.21 1 0.091 0.007 0.41 1.00

x2x3 0.16 1 0.091 –0.043 0.36 1.00

x12 –0.11 1 0.066 –0.26 0.035 1.75

x22 –0.34 1 0.066 –0.48 –0.19 1.75

x32 –0.30 1 0.066 –0.44 –0.15 1.75

187
4.10.2 Analysis of variance (ANOVA) for biodiesel production

The statistical significance and the goodness of fit of the developed second-order quadratic

model, as well as the effect of individual terms and their interactions were analyzed by

ANOVA, and the results are summarized in Table 4.17. The quality of the model fit was

evaluated by the Fisher's test (F-value), the probability value (p-value), the "lack of fit", the

coefficient of determination (R2), adjusted R2 (R2adj), and predicted R2 (R2pred). The p-value

should be less than 0.05 for the variables to have significant effect on the response (RSOME

yield). The polynomial model has an F-value of 12.38 and p-value <0.0003, which indicates

that the developed model is statistically significant at the 95% confidence level. The higher the

F-value for the specific independent process variable, the higher will be the effect of that

variable. Two main effects, i.e. x2 and x3 are significant based on their estimated p-values

(<0.05). This shows that biodiesel produced from rubber seed oil could be effectively catalyzed

using indigenously prepared RSS catalyst. From the ANOVA analysis, the interactive effects

of x1x2 and x2x3 on the biodiesel yield are significant (p<0.05) even though x1 appears

insignificant. The effects of individual process variable as shown in Table 4.16 indicate that

the contribution of reaction time is low compared to other variables. However, there is a strong

interaction between reaction time and other variables as presented in Tables 4.16 and 4.17.

In the present study, the "lack of fit" of low F-value of 0.079 and high p-value of 0.9246 implied

that the model was not significant relative to pure error (0.65). The p-value of the "lack of fit"

parameter is greater than 0.05, indicating that there is good fit between the second-order

quadratic model and the experimental esters yield obtained (Dharma et al., 2016).

The coefficient of determination (R2) reflects the variability of the dependent response variable,

which is explained by its relationship with the independent process variables (Dharma et al.,

2016).

188
In the present study, the R2 value is 0.9187, which indicates that 91.87% of the variability in

the biodiesel yield is attributed to the independent factors considered (reaction time,

methanol/oil ratio, and catalyst amount), and only 8.13% was not explained by the model. The

R2 value is close to unity and signifies that the experimental data obtained linearly fits in the

chosen model equation. The adjusted and predicted R2 for the present study was obtained as

0.8436 and 0.7887, respectively. The difference between the adjusted R2 and predicted R2

values, which is 0.0549, is below the maximum allowable difference of 0.2 (Dhawane et al.,

2016). This explains that the model is capable of predicting the response in a reasonable

acceptable range. The adequate precision value measures the signal-to-noise ratio, and the

present study gave adequate precision of 9.83, which was higher than the critical value of 4,

thereby indicating that the model has strong signal to navigate the design space for the

optimization purpose (Kostićet al., 2016; Betiku & Ajala, 2014; Cerino Córdovaet al., 2011).

The coefficient of variance (C.V.) obtained for the present study was found to be low (0.313%),

which indicates that the experimental data are accurate and reliable to suggest that the quadratic

model is capable of optimizing the process.

189
Table 4.17 Analysis of variance (ANOVA) for the biodiesel production using RSM model

Source of variance Sum of squares degree of freedom Mean square F-value p-value

Model 7.44 9 0.83 12.38 0.0003a

x1 0.0006595 1 0.0006595 0.009884 0.9228

x2 0.39 1 0.39 5.78 0.0371a

x3 1.03 1 1.03 15.47 0.0028a

x1x2 0.54 1 0.54 8.11 0.0173a

x1x3 0.35 1 0.35 5.29 0.0443a

x2x3 0.20 1 0.20 3.07 0.1103

x12 0.19 1 0.19 2.85 0.1224

x22 1.75 1 1.75 26.18 0.0005a

x32 1.38 1 1.38 20.66 0.0011a

Residual 0.67 10 0.067

Lack of fit 0.013 2 0.006478 0.079 0.9246

Pure error 0.65 8 0.082

Corrected total 8.10 19

Statistical parameters:

Standard deviation: 0.26 R2: 0.9187

Mean of response: 82.89 Adjusted R2: 0.8436

Coefficient of variance (%): 0.313 Predicted R2: 0.7887

Adequate precision: 9.83

x1: reaction time (min), x2: methanol/oil ratio (vol/vol), x3: catalyst amount (g)
a
statistically significant at the confidence level of 95%

190
4.10.3 Parametric effects of interactive factors on biodiesel yield by RSM

In design of experiments (DOEs), the 3D model equation facilitates examination of the effects

of process variables on the response variable. The surface plot depicted in Figure 4.23 (a) shows

the interaction between reaction time (min) and methanol/oil ratio (vol/vol), with the catalyst

amount constant at the center point. Rubber seed oil methyl ester yield, marginally (0.4%)

increased with increase in reaction time from 60 to 70 min as shown in Figure 4.23 (a). At this

condition, 0.6 vol/vol (coded) that translates to 0.28 vol/vol (actual) methanol/oil ratio was

found to be the optimum for attaining 83.4% of yield when methanol/oil ratio varies from 0.20

to 0.30 vol/vol (actual). This shows that the interactive effect is not significant enough at these

conditions of low range of reaction time (Onoji et al., 2017b).

Figure 4.23 (a): RSM 3D plot for the interactive effect between reaction time and

methanol/oil ratio on biodiesel (RSOME) yield

191
Figure 4.23 (b) depicts the interaction effect between reaction time and catalyst amount on

rubber seed oil methyl ester yield, with methanol/oil ratio constant at the center point. It was

evident that the reaction time's effect had the same trend as in Figure 4.23 (a). The effect of the

reaction time alone on the yield as shown in Tables 4.16 and 4.17 is not significant, although

it has significant interactions with methanol/oil ratio and catalyst amount as previously

observed.

Figure 4.23 (b): RSM 3D plot for the interactive effect between reaction time and catalyst

amount on biodiesel (RSOME) yield

Similar observations, but with slight reduction in yield (83%), were noticed for the plot of

Figure 4.23 (c) that represents the interaction between methanol/oil ratio and catalyst amount,

when the reaction time is constant at the center point. Marginal reduction in yield was observed

when methanol/oil ratio varies from 0.20 to 0.30 vol/vol, with catalyst optimum value of 2.5 g.

It is worthy to note that each of the interactions in the surface plots exhibited a curvature that

indicates good interactive effects between the process factors involved.

192
Figure 4.23 (c): RSM 3D plot for the interactive effect between methanol/oil ratio and

catalyst amount on biodiesel (RSOME) yield

4.10.4 Optimization and interactive effects of process variables on biodiesel yield by ANN

model

Iteratively, three neurons were obtained in the hidden layer that gave a good prediction of the

outputs of training and validation sets in the present study. A network of 3-3-1 topology was

developed, and the genetic algorithm of the model predicted optimal biodiesel (RSOME) yield

of 85.07% at the following conditions: reaction time (X1=56.70 min), methanol/oil ratio

(X2=0.21 vol/vol), and catalyst amount (X3=2.20 g). The experiment was validated in triplicates

using the optimum conditions, and the average optimal biodiesel yield was 85.03±0.0012%.

This value is close to the model predicted value, and within ±0.051% relative error based on

Equation (3.27). The % absolute average deviation (AAD) calculated from Table 4.15, and

based on Equation (3.28) was ±0.000473%. Hence, the developed model gave a good estimate

of the response of the process for the low values obtained. The model has a coefficient of
193
determination (R2) value of 0.9198, R2adj value of 0.9182, and R2pred value of 0.9168. These

values indicate a good and reliable model as was obtained for RSM.

Figure 4.24 compares the predicted and experimental values of the biodiesel yield, and the

coefficient of determination (R2 = 0.9198) shows a good correlation between the data obtained.

This confirms that the model is suitable for adequate representation of the actual relationship

among the selected factors.

84.0

83.5
R² = 0.919
83.0
Predicted

82.5

82.0

81.5

81.0
81.0 81.5 82.0 82.5 83.0 83.5 84.0 84.5
Actual
Figure 4.24: Correlation of predicted vs. actual biodiesel yields for ANN model

Figures 4.25, 4.26, and 4.27 depict the interactive effects on biodiesel (RSOME) yield between

two process variables while maintaining the third variable constant at center point. The

curvatures nature of the 3D surfaces suggested significant interactions of methanol/oil ratio

with reaction time, catalyst amount with reaction time, and methanol/oil ratio with catalyst

amount. Figure 4.26 shows steep decrease in the biodiesel yield as catalyst loading increases.

This could be observed as excess catalyst negates the mixing of methanol, oil and catalyst,

which led to phase separation and subsequent reduction in biodiesel yield. The observation is

similar to the reports of other researcher that as catalyst loading increases, biodiesel yield

194
increases until optimum catalyst loading is attained (Syazwani et al., 2015). The level of

importance of the process variables considered for the RSOME synthesis is depicted in Figure

4.28. It is apparent that catalyst amount (45.62%) is the most significant variable, followed by

methanol/oil ratio (36.87%) and lastly, reaction time (17.51%).

Figure 4.25: ANN Surface plot for the interactive effect of methanol/oil ratio with reaction

time on (biodiesel) RSOME yield

195
Figure 4.26: ANN Surface plot for the interactive effect of catalyst amount with

reaction time on (biodiesel) RSOME yield

Figure 4.27: ANN Surface plot for the interactive effect of methanol/oil ratio with

catalyst amount on biodiesel (RSOME) yield


196
Figure 4.28: Percentage contribution of process variables to biodiesel production via

ANN model (catalyst amount 45.62%, methanol/oil ratio 36.87%, and reaction time

17.51%)

4.10.5 Comparison of RSM and ANN models for the biodiesel production

The accuracies of both RSM and ANN models were evaluated based on the values obtained

for R2, % AAD, % relative error and the yields obtained at optimum conditions for the models.

The RSM model equation at its optimum conditions (X1 = 60 min, X2 = 0.20 vol/vol and X3 =

2.5 g) gave a yield of 83.11%. The ANN model at its optimum conditions (X1 = 56.7 min, X2 =

0.21 vol/vol and X3 = 2.2 g) gave a higher biodiesel yield of 85.07%. The supremacy of the

ANN model over RSM was further attested to by the lower values of relative error (±0.051%)

and AAD (±0.000473%) compared to the values of RSM relative error (±0.06%) and AAD

(±0.022%).

However, both optimization techniques gave good prediction due to the values of R2 that are

relatively close to unity (i.e. 0.9187 and 0.9198 for RSM and ANN, respectively). The non-

linearity of the ANN may contribute to its superiority over RSM (Ma et al., 2016).

197
4.11 Performance and emissions analysis of test engine using biodiesel-diesel blends

The experimental analyses of the data for engine performance and emission characteristics

obtained from the tested fuels (B00, B10, B20, B30, B50, and B100) used in this study are

presented in this section. The properties of the tested fuels needed for the computation of

combustion parameters are listed in Table 4.18.

Table 4.18: Properties of the test fuels

Fuel type B00 B10 B20 B30 B50 B100

Density (kg/m3) 850 853 855 858 863 876

LHV (kJ/kg) 42,900 42,400 41,900 41,500 40,500 38,200

Kinematic viscosity (mm2/s) 3.21 3.32 3.43 3.54 3.76 4.32

Cetane number 52 52.5 53 53.5 54.5 57

4.11.1 Air/Fuel ratio (AFR)

The air-fuel mixture formation in the combustion chamber affects the combustion performance

and emission characteristics of diesel engines (Li et al., 2016). The air-box differential pressure

(∆P) and the ambient air pressure (PA) are important parameters for the computation of the

flowrate of air. The flowrate of air and the air/fuel ratio were calculated using Equations (3.31)

and (3.32), respectively. Figure 4.29 depicts the variation of air/fuel ratio with engine speed for

the fuels tested (B00, B10, B20, B30, B50, and B100). The plots data were obtained from

Tables I.1 through I.6 of Appendix I. The Figure depicts that B20 has the best air/fuel ratio,

and biodiesel the least. A higher air/fuel ratio means lower amount of fuel was consumed for

the given air flowrate. The lower value for biodiesel is attributed to its high density and

198
viscosity compared to other blends, and higher fuel consumption rate for biodiesel for same

power output.

45

40 B00

35
B10
Air/Fuel ratio

30
B20
25

20 B30

15 B50

10
B100
5
1000 1500 2000 2500 3000 3500 4000
Engine speed (rpm)

Figure 4.29: Variation of air/fuel ratio with speed for diesel and biodiesel blends

4.11.2 Brake mean effective pressure (BMEP)

The calculated BMEPs for the tested fuels are presented in Tables I.1 through I.6 of Appendix

I, and were computed using Equation (3.33). The BMEP is a useful tool for comparing and

evaluating similar types of engines. The maximum BMEP developed by the engine for all fuels

tested was 4.75 bar for pure diesel (B00) at 2500 rpm and it was within the range of values

reported for this engine using diesel fuel (TecQuipment Ltd, 2011). It was also observed that

for all fuels tested, their maximum BMEP was obtained at 2500 rpm, which appears to be the

ideal speed to operate the test engine for maximum efficiency.

199
4.11.3 Brake thermal efficiency (BTE%)

Brake thermal efficiency is the ratio between the effective power developed and the rate of

energy introduced by fuel injection (Aldhaidhawi et al., 2017). The variation of BTE% of the

engine with speeds for the biodiesel blends used in this study in relation to diesel fuel is

depicted in Figure 4.30 with plot data obtained from Table I.7 of Appendix I. It was observed

that BTE% decreases as the percentage of biodiesel in the blends increases at all speeds due to

poor atomization of the blends coupled with high viscosity. Among the blends, B50 was found

to have the maximum thermal efficiency of 42.8% at a brake power of 1.88 kW, while for

diesel oil, it was 56% and biodiesel 28.2% at 2500 rpm. It can be shown that beyond the speed

limit of 2500 rpm, the brake thermal efficiency trend is reverted and started decreasing. The

lower BTE% obtained for B100 could be due to the low calorific value of biodiesel and increase

in fuel consumption as compared to B50. Ramadhas et al. (2005a) reported this trend for a

diesel engine fueled with rubber seed oil methyl esters (biodiesel).

60 B00 (diesel)
55
Brake thermal efficiency (%)

B10
50
45 B20
40
35 B30

30 B50
25
20 B100 (pure
biodiesel)
15
10
1000 1500 2000 2500 3000 3500 4000
Engine speed (rpm)

Figure 4.30: Variation of brake thermal efficiency with speed for diesel and biodiesel

blends

200
4.11.4 Brake specific fuel consumption (BSFC)

The term brake specific fuel consumption (BSFC) refers to fuel consumption quantities which

have been normalized by the engine power representing actual fuel consumed to produce 1 kW

of power output (Aldhaidhawi et al., 2017). Figure 4.31 depicted below, and obtained from

data presented in Table I.8 of Appendix I shows the variation of BSFC of diesel and biodiesel

blends at full load conditions with engine speeds. It was found that the BSFC for the blend B20

was close to that of diesel. For higher blends, the BSFC was found to be much higher than

diesel at all speeds. This is the result of the combined effects of lower heating value and the

high fuel flow rate due to high viscosity of the blends as reported by other researchers (Celikten

et al., 2012). The finding of this study is consistent with findings of past studies (Labeckas &

Slavinskas, 2006; Raheman & Ghadge, 2008).

0.6
Brake specific fuel consumption (kg/kWh)

B00 (diesel)
0.5
B10
0.4
B20

0.3 B30

B50
0.2

B100 (pure
0.1 biodiesel)

0
1000 1500 2000 2500 3000 3500 4000

Engine speed (rpm)

Figure 4.31: Variation of brake specific fuel consumption with speed for diesel and

biodiesel blends

201
4.11.5 Engine brake power

Figure 4.32 obtained from data presented at Table I.9 of Appendix I shows that the brake power

(BP) increases progressively for all types of fuels tested (B00, B10, B20, B30, B50 and B100)

all speeds at full load conditions. The value of BP recorded followed similar trends for all fuels

blends. The BP value increment was related to the increase in engine speed, since at high speeds

the combustion inside the cylinder was close to completion, which produced increasing brake

power. The lower value of engine power for biodiesel and its blends is related to its lower

calorific value compared to fossil diesel, and it is a natural phenomenon (Mohsin, et al., 2014).

However, since the difference for biodiesel in each fuel blend was 10–20%, the data showed

similar results, but B20 and B50 showed characteristics that are similar to pure diesel with

respect to brake power. It may be argued that the higher brake specific fuel consumption of

rubber seed biodiesel and its blends will compensate this power loss and maintain the fuel

quantity, finally yielding the same engine output. Mohsin et al. (2014) reported a similar trend

at full load conditions using biodiesel blends in a test engine.

202
2.5

2
Engine power (kW)

1.5 B00 (diesel)

B10

B20
1
B30

B50
0.5
B100 (pure
biodiesel)

0
1000 1500 2000 2500 3000 3500 4000

Engine speed (rpm)

Figure 4.32: Variation of engine power with speed for diesel and biodiesel blends

4.11.6 Engine torque

As shown in Figure 4.33 below, there was reduction in the engine torque for each type of fuels

blends used in the present study beyond 2500 rpm. The data for the plots are recorded in Table

I.10 of Appendix I. The value of torque decreased as the engine speed increased beyond 2500

rpm, because of less force needed to push the engine piston at high speed. This decreasing

value of torque for high speed is considered normal for every type of engine (Mohsin et al.,

2014). Lower torque value at a higher engine speed is one of the criteria that ensure positive

reaction of the fuel inside the cylinder.

For all fuels tested, maximum torque was achieved at 2500 rpm, but biodiesel fuel gave the

lowest torque value at all speeds. A similar trend was reported by Mohsin et al. (2014), which

203
showed maximum torque for all blends at 2400 rpm but tended to decrease with further increase

in speed.

10
9 B00 (diesel)

8 B10
Engine torque (Nm)

7
B20
6
5 B30

4 B50
3
B100 (pure
2 biodiesel)
1
0
1000 1500 2000 2500 3000 3500 4000
Engine speed (rpm)

Figure 4.33: Variation of engine torque with speed for diesel and biodiesel blends

4.11.7 Exhaust gas temperature (EGT)

Figure 4.34 shows the plots of exhaust gas temperature (EGT) with engine speeds for biodiesel

blends and diesel obtained from Table I.11 of Appendix I. The results show that the EGT

increases with increase in engine speed for all fuels tested. At all speeds, fossil diesel was found

to have the lowest temperature and the temperature for various blends show an upward trend

with increasing volume of biodiesel in the blends. The ester molecule in the biodiesel contains

oxygen atoms that enhance the combustion process, which results to higher EGTs. In addition,

air-coded engines run hotter resulting to higher EGTs. The high EGT for biodiesel resulted to

higher concentration of NOx in the exhaust gas emission. These findings are consistent with

those reported by other researchers (Ramadhas et al., 2005a; Elango & Senthilkumar, 2011).

204
350

B00 (diesel)
Exhaust gas temperature (o C)

B10
300
B20

B30
250
B50

B100 (pure
biodiesel)
200

150
1000 1500 2000 2500 3000 3500 4000

Engine speed (rpm)

Figure 4.34: Variation of exhaust gas temperature with speed for diesel and biodiesel

blends

4.11.8 CO emissions

Carbon monoxide (CO) is the result of incomplete combustion of fuel such as fossil diesel that

lacks oxygen molecule in its molecular structure. Several factors such as air-fuel ratio, engine

speed, injection timing, injection pressure, and type of fuels affect CO emissions. The different

CO emission values of diesel and biodiesel blends with their trend lines at various engine

speeds are depicted in Figure 4.35 with data obtainable from Table I.12 of Appendix I. In all

causes considered, CO emission at 2500 rpm recorded the lowest values and it could be

considered as the ideal speed for the engine operation. It is interesting to note that the engine

emits more CO using diesel as compared to that of biodiesel and blends at all speeds. This

result can be attributed to the higher oxygen content and cetane number of the biodiesel fuel

than diesel fuel. As the percentage of biodiesel increased in the blend, the higher oxygen

205
content to biodiesel allows more carbon molecules to burn and facilitates the completion of

combustion process (Mofijur et al., 2014).

170

160
B00
(diesel)
150 B10

140 B20
CO (ppm)

B30
130
B50
120
B100 (pure
biodiesel)
110

100
1000 1500 2000 2500 3000 3500 4000
Engine speed (rpm)

Figure 4.35: Variation of carbon monoxide emissions with speed for diesel and biodiesel

blends

4.11.9 CO2 emissions

Figure 4.36 compares the CO2 emissions of various fuel blends tested in the diesel engine at

different speeds. Naturally, emission of CO2 from the engine is due to complete combustion of

the fuel in the combustion chamber. The release of more CO2 indicates the closeness to

complete combustion of fuel. In the present study as shown in Figure 4.36, and data obtained

from Table I.13 of Appendix I, and the blend of B20, gave the highest amount of CO2 emission

starting from 1500 through 3500 rpm. This supports the higher values of exhaust gas

temperature. In all the cases considered, diesel (B00) produces the lowest amount of CO2 due

to incomplete combustion. Although, CO2 is considered a greenhouse gas that causes increased

206
global warming, from the engine performance point of view, it achieved complete combustion

and that the CO2 emitted are readily absorbed by plants for food production process

(photosynthesis) that keeps the level of CO2 in the atmosphere balance (Mohsin et al., 2014).

310

290 B00 (diesel)


270
B10
250
CO2 (ppm)

230 B20

210
B30
190

170 B50

150 B100 (pure


130 biodiesel)

110
1000 1500 2000 2500 3000 3500 4000
Engine speed (rpm)

Figure 4.36: Variation of carbon dioxide emissions with speed for diesel and biodiesel

blends

4.11.10 NOx emissions

The variation of NOx emissions of biodiesel-diesel blends with speeds obtained from Table

I.14 of Appendix I is depicted in Figure 4.37. The NOx emission for biodiesel and the blends

followed are increasing trend with respect to engine speed. In all the fuels tested, NOx emission

was higher than the neat diesel (B00) with biodiesel and B50 exhibiting higher values than

other blends. NOx emission is directly related to the engine combustion chamber temperatures

which in turn indicated by the prevailing exhaust gas temperature. With increase in the value

of exhaust gas temperature, NOx emission also increases (Ramadhas et al., 2005a). The results
207
of the present study are in agreement with the reports of researchers cited in the literature

(Özener et al., 2014). Most researchers proposed that oxygen content of biodiesel, unexpected

advance in the fuel injection timing, extended combustion timing, and higher combustion

temperature trigger higher formation of NOx (Ozsezen & Canakci, 2011; Murillo et al., 2007).

60

B00 (diesel)
50
B10
40
B20
NOx (ppm)

30
B30

20 B50

10 B100 (pure
biodiesel)

0
1000 1500 2000 2500 3000 3500 4000
Engine speed (rpm)

Figure 4.37: Variation of nitrogen oxides emissions with speed for diesel and biodiesel

blends

4.11.11 Total hydrocarbons (THCs)

Figure 4.38 with plot data obtained from Table I.15 of Appendix I shows that the amount of

THCs decreased with the increasing amount of biodiesel in the blends. This indicates that most

of the total hydrocarbons emissions present in the combustion chamber were generated from

the pure diesel used since the amount of fuel injected was constant for each type of blend tested.

In all cases considered, engine speed at 3000 rpm produces the lowest amount of THCs, after

which there was sudden increase in the content of THCs emitted at higher speeds.

208
130
B00 (diesel)
120
B10
110
B20
THC (ppm)

100 B30

90 B50

80 B100 (pure
biodiesel)
70

60

50
1000 1500 2000 2500 3000 3500 4000
Engine speed (rpm)

Figure 4.38: Variation of total hydrocarbons emissions with speed for diesel and biodiesel

blends

4.11.12 Smoke opacity

The variation of smoke density with respect to different fuels considered is depicted in Figure

4.39, with plot data presented in Table I.16 of Appendix I. The smoke density for biodiesel and

blends is generally lowered than that of diesel. B20 blend gave the lowest smoke intensity

compared to other blends. Higher cetane number and lower compressibility of the biodiesel

resulted in a lower ignition delay compared to that of diesel. A lower ignition delay reduces the

accumulation of fuel in the combustion chamber and reduces smoke emissions (Geo et al.,

2009). The formation of smoke primarily results from the unburnt molecules of fuels because

of partial reaction of carbon in the liquid fuel (Özener et al., 2014). As shown in Figure 4.39,

the smoke opacity decreases as the content of biodiesel in the blends increases. The reduction

209
in the smoke opacity with increase in biodiesel content can also be attributed to the decrease in

the carbon content, and increase in the concentration of oxygen in biodiesel fuel.

7
B00 (diesel)
6
Smoke opacity (BSU)

B10
5
B20
4
B30
3
B50
2
B100 (pure
1 biodiesel)

0
1000 1500 2000 2500 3000 3500 4000
Engine speed (rpm)

Figure 4.39: Variation of smoke opacity with speed for diesel and biodiesel blends

210
CHAPTER 5: CONCLUSION AND RECOMMENDATIONS

5.1 Conclusion

Energy security has been the major hindrance to the development of African countries. Biomass

in the form of charcoal and fuelwood continues to be the main fuel source in these countries.

Biodiesel, a sustainable energy whose properties are similar to petro-diesel could reduce the

dependency on fossil diesel. Over 90% of biodiesel produced is sourced from edible oils that

compete with food uses. Food is a necessity of man after shelter; and the use of edible vegetable

oils as biodiesel source materials is counterproductive. Non-edible low-cost rubber seed oil

obtained from abundant waste rubber seeds in Nigeria was exploited for biodiesel production

in this study. The oil was chemically extracted from the Nigerian NIG800 clonal seeds using

n-hexane solvent in a soxhlet extractor, and the physicochemical properties determined were

in the range previously reported by researchers for rubber seed oil.

The oil extraction process variables were optimized using the response surface methodology

and artificial neural network techniques via Box-Behnken experimental design. The developed

models were effective in describing the parametric effect of the considered operating variables

on the extraction of oil from rubber seeds. The high oleic acid content and the presence of cis-

erucic acid indicate the potential use of the oil as a source material for oleo chemical and steel

industries in the future. The kinetics of thermo-oxidative degradation of the oil follows first-

order kinetics with activation energy of 13.07 kJ/mol and thermally stable up to 250 oC.

The applicability of the waste rubber seed shell obtained during oil extraction was investigated

as a base catalyst for the oil transesterification to biodiesel.

The lower energy content of the produced rubber seed oil biodiesel affects the brake specific

fuel consumption, but it was compensated by the improved combustion efficiency and lower

environmental pollution as observed for total hydrocarbon, carbon monoxide, and smoke

opacity emissions compared to fossil diesel. Optimal performance of the fuel was observed
211
with the B20 blend followed by B50 and the engine operated at 2500 rpm. To ensure timely

dissemination of the findings in this study, results have been communicated to colleagues and

researchers in the field via four published journal articles and one communication in an

international conference.

5.2 Recommendations

It is recommended that further studies should be undertaken on such as:

1. The developed basic heterogeneous rubber seed shell catalyst doped with sulphuric acid via

wet impregnation method be investigated for biodiesel production. This will ensure that both

esterification and transesterification reactions are carried out in a single stage expected to

generate a higher yield of biodiesel at a lower processing time;

2. The mechanisms of the developed catalyst used in this thesis should be investigated further

to offer insight into its anticipated behaviour during biodiesel production;

3. A zero-dimensional simulation model should be employed to predict performance

characteristics of the engine using biodiesel produced from rubber seed oil catalyzed by rubber

seed shell catalyst;

4. Development of a differential thermodynamic equation system to obtain the pressure inside

the cylinder as a function of the crank angle (θ) for differential engine conditions is

recommended as future work;

5 Effects of the biodiesel on the engine components such as plunger, injectors, cylinder head,

valves, piston and rings should be investigated in an endurance test to determine the level of

components wear.

212
5.3 Contributions to knowledge

1. The outcomes of this research will serve as a pool of information for the relevant

Government Ministries, Parastatals, Departments, Research Institutes, Industries, and

Agencies that will serve as advocates for the implementation of the National Biofuels Policy

across Nigeria's business strata with respect to biodiesel production from rubber seed oil and

other African nations where rubber trees are cultivated;

2. This research provides valuable information to the global community that the genetic

engineered rubber seed obtained from Nigerian NIG800 series contains high content of oleic

acid and cis-erucic acid as a source material for oleo chemicals currently sourced from non-

renewable fossil materials;

3. The performance and emission analyses of diesel engine show that the biodiesel produced

from rubber seed oil via calcined waste rubber seed shell as catalyst could be used in the modern

diesel engines without technical modifications.

213
REFERENCES

Abassi, A., Khalilarya, Sh., & Jafarmadar, S. (2010). The influence of the inlet charge

temperature on the second law balance under the various operating engine speed in DI

diesel engine. Fuel, 89, 2425–2432.

Abdulkadir, B.A., Danbature, W., Yirankinyuki, F.Y., Magaji, B., & Muzakkir, M. M. (2014).

Insitu transesterification of rubber seeds (Hevea brasiliensis). Greener Journal of

Physical Sciences, 4 (3), 038–044.

Abila, N. (2012). Biofuels development and adoption in Nigeria: Synthesis of drivers,

incentives and enablers. Energy Policy, 43, 387–395.

Ahmad, J., Yusup, S., Bokhari, A., & Kamil, R. N. M. (2014a). Biodiesel production from the

high free fatty acid “Hevea brasiliensis” and fuel properties characterization. Applied

Mechanics and Materials, 625, 897–900.

Ahmad, J., Yusup, S., Bokhari, A., & Kamil, R. N. M. (2014b). Study of fuel properties of

rubber seed oil based biodiesel. Energy Conversion and Management, 78, 266–275.

Ahmad, J., Bokhari, A., & Yusup, S. (2014c). Optimization and parametric study of free fatty

acid (FFA) reduction from rubber seed oil (RSO) by using response surface

methodology (RSM). Australian Journal of Basic and Applied Sciences, 8(5), 299–303.

Ahmad, M., Teong, L.K., Zafar, M., Sultana, S., Sadia, H., & Khan, M. A. (2013). Prospects

and potential of green fuel from some non-traditional seed oils used as biodiesel. In: Z.

Fang (Ed.),. (Zheng fang, Ed.). https://doi.org/10.5772/52031

Aigbodion, A.I., & Pillai, C. K. S. (2000). Preparation, analysis and applications of rubber seed

oil and its derivatives in surface coatings. Progress in Organic Coatings, 38, 187–192.

Ajiwe, V.I.E., Okeke, C.A., & Agbo, H. U. (1995). Extraction and utilization of Afzelia

africana seed oil. Bioresource Technology, 53, 89–90.

Aldhaidhawi, M., Chirac, R., & Badescu, V. (2017). Ignition delay, combustion and emission

214
characteristics of diesel engine fueled with rapeseed biodiesel-A literature review.

Renewable and Sustainable Energy Reviews, 73, 178–186.

Ali, M., & Watson, I. A. (2014). Comparison of oil extraction methods, energy analysis and

biodiesel production from flax seeds. International Journal of Energy Research, 38,

614–625.

Alkida, A. C. (1988). The application of availability and energy balances to a diesel engine.

Journal of Engineering for Gas Turbines and Power, 110 (3), 462–469.

Amigun, B., Musango, J.K., & Stafford, W. (2011). Biofuels and sustainability in Africa.

Renewable and Sustainable Energy Reviews, 15, 1360–1372.

Amigun, B., Sigamoney, R., & Blottnitz, H. von. (2008). Commercialization of biofuel

industry in Africa: a review. Renewable and Sustainable Energy Reviews, 12, 690–711.

Anastopoulos, G., Zannikou, Y., Stournas, S., & Kalligeros, S. (2009). Transesterification of

vegetable oils with ethanol and characterization of the key fuel properties of ethyl

esters. Energies, 2, 362–376.

Aniamaka, E.E., & Uriah, O. B. C. (1990). Collection of preparation of rubber (Hevea

brasiliensis) seeds for seedling production In: Proceedings of a National Workshop on

Fruit/tree production. (p. 75–79.). NAERLS Zaria, Nigeria.

Antczak, M.S., Kabiak, A., Antczak, T., & Bielecki, S. (2009). Enzymatic biodiesel synthesis-

Key factors affecting efficiency of the process. Renewable Energy, 34 (5), 1185–1194.

Aransiola, E.F., Ojumu, T.V., Oyekola, O.O., Madzimbamuto, T.F. & Ikhu-Omoregbe, D. I.

O. (2014). A review of current technology for biodiesel production: State of the art.

Biomass and Bioenergy, 61, 276–297.

Aravind, A., Joy, M.L., & Nair, K. P. (2015). Lubricant properties of biodegradable rubber tree

seed (Hevea brasiliensis Muell. Arg) oil. Industrial Crops and Products, 74, 14–19.

Ashraful, A.M., Masjuki, H.H., Kalam, M.A., Rizwanul Fattah, I.M., Imtenan, S., Shahir, S.

215
A., & & Mobarak, H. M. (2014). Production and comparison of fuel properties, engine

performance and emission characteristics of biodiesel from various non-edible

vegetable oils: a review. Energy Conversion and Management, 80, 202–228.

Association of Official Analytical Chemists. (1990). AOAC. Official methods of analysis, 15th

ed., (15th ed.). Washington D.C.

Asuquo, J.E., Anusiem, A.C.I., & Etim, E. E. (2012). Extraction and characterization of rubber

seed oil. International Journal of Modern Chemistry, 1 (3), 109–115.

Atabani, A.E., Silitonga, A.S., Ong, H.C., Mahlia, T.M.I., Masjuki, H.H., Badruddin, I.A., &

Fayaz, H. (2013). Non-edible vegetable oils: A critical evaluation of oil extraction, fatty

acid compositions, biodiesel production, characteristics, engine performance and

emissions production. Renewable and Sustainable Energy Reviews, 18, 211–245.

Awolu, O.O., & Layokun, S. K. (2013). Optimization of two-step transesterification production

of biodiesel from neem (Azadirachta indica) oil. International Journal of Energy and

Environment Engineering, 4 (39), 1–9.

Aylmore, L. A. G. (1974). Hysteresis in gas sorption isotherms. Journal of Colloid and

Interface Science, 46 (3), 410–416.

Azoumah, Y., Blin, J., & Daho, T. (2009). Exergy efficiency applied for the performance

optimization of a direct injection compression ignition (CI) engine using biofuels.

Renewable Energy, 34, 1494–1500.

Bakare, I.O., Pavithran, C., Okieimen, F.E., & Pillai, C. K. . (2006). Polyesters from renewable

resources: preparation and characterization. Journal of Applied Polymer Science, 100,

3748–3755.

Banković-Ilić, I.B., Stamenković, O.S., & Veljković, V. B. (2012). Biodiesel production from

non-edible plant oils. Renewable and Sustainable Energy Reviews, 16, 3621–3647.

Baş, D., & Boyaci, I. H. (2007a). Modeling and optimization II: comparison of estimation

216
capabilities of response surface methodology with artificial neural networks in a

biochemical reaction. Journal of Food Engineering, 78 (3), 846–854.

Baş, D., & Boyaci, I. H. (2007b). Modeling and optimization I: usability of response surface

methodology. Journal of Food Engineering, 78 (3), 836–845.

Bayindir, H., Ișik, M.Z., Argunhan, Z., Yücel, H.L., & Aydin, H. (2017). Combustion,

performance and emissions of a diesel power generator fueled with biodiesel-kerosene

and biodiesel-kerosene-diesel blends. Energy, 123, 241–251.

Bayraktar, H. (2005). Experimental and theoretical investigation of using gasoline-ethanol

blends in spark-ignition engines. Renewable Energy, 30, 1733–1747.

Betiku, E., & Ajala, S. O. (2014). Modeling and optimization of Thevetia peruviana (yellow

oleander) oil biodiesel synthesis via Musa paradisiacal (plantain) peels as

heterogeneous base catalyst: A case of artificial neural network vs. response surface

methodology. Industrial Crops and Products, 53, 314–322.

Bhuiya, M.M.K., Rasul, M.G., Khan, M.M.K., Ashwath, N., Azad, A.K., & Mofijur, M.

(2015). Optimization of oil extraction process from Australian native beauty leaf seed

(Calophyllum inophyllum). Energy Procedia, 75, 56–61.

Bohre, A. (2013). Immobolization of radioactive waste in ceramic-based hosts: Radioactive

waste immobolization. Hamburg: Anchor Academic Publishers.

Bokhari, A., Yusup, S., & Ahmad, M. M. (2012). Optimization of the parameters that affects

the solvent extraction of crude rubber seed oil using response surface methodology

(RSM). In C. C. L. Dmitriev A. (Ed.), Recent Advance in Engineering (pp. 28–30).

Paris: WSEAS.

Bokhari, A., Yusup, S., Kamil, R.N.M., & Ahmad, J. (2014). Blending study of palm oil methyl

esters with rubber seed oil to improve biodiesel blending properties. Chemical

Engineering Transactions, 37, 571–576.

217
Bora, M.M., Gogoi, P., Deka, D.C., & Kakati, D. K. (2014). Synthesis and characterization of

yellow oleander (Thevetia peruviana) seed oil-based alkyd resin. Industrial Crops and

Products, 52, 721–728.

Boro, J., Deka, D., & Thakur, A. J. (2012). A review on solid oxide derived from waste shells

as catalyst for biodiesel production. Renewable and Sustainable Energy Reviews, 16,

904– 910.

Borugadda, V.B., & Goud, V. V. (2013). Comparative studies of thermal, oxidative and low

temperature properties of waste cooking oil and castor oil. Journal of Renewable and

Sustainable Energy, 5 (63104), 1–15.

Botella, L., Bimbela, F., Martin, L., Arauzo, J., & Sánchez, J. L. (2014). Oxidation stability of

biodiesel fuels and blends using the Rancimat and PetroOXY methods. Effect of 4-

allyl-2, 6-dimethoxyphenol and catechol as biodiesel additives on oxidation stability.

Frontiers in Chemistry, 2 (43), 1–9.

Caton, J. A. (2000). On the destruction of availability (exergy) due to combustion processes-

with specific application to internal combustion engines. Energy, 25, 1097–1117.

Celikten, I., Mutlu, E., & Solmaz, H. (2012). Variation of performance and emission

characteristics of a diesel engine fueled with diesel, rapeseed oil and hazelnut oil methyl

ester blends. Renewable Energy, 48, 122–126.

Ҫengel, Y.A., & Boles, M. A. (1994). Thermodynamics: an engineering approach (2nd ed.).

New York: McGraw-Hill, Inc.

Cerino Córdova, F.J., Garcia León, A.M., Garcia Reyes, R.B., Garza González, M.T., Soto

Regalado, E., Sánchez González, M.N., & Quezada Lόpez, I. (2011). Response surface

methodology for lead biosorption on Aspergillus terreus. International Journal

Environmental Science Technology, 8 (4), 695–704.

Ceylan, S., & Goldfarb, J. L. (2015). Green tide to green fuels: TG-FTIR analysis and kinetic

218
study of Ulva prolifera pyrolysis. Energy Conversion and Management, 101, 263–270.

Charles, S. (2002). Noble obsession, Charles Goodyear, Thomas Hancock, and the race to

unlock the greatest industrial secret of the 19th century. Hyperion.

Choe, E., & Min, D. B. (2007). Chemistry of deep-fat frying oils. Journal of Food Science, 72

(5), 77–86.

Christmann, K. (2012). Thermodynamics and Kinetics of Adsorption. Institut für Chemie und

Biochemie. IMPRS-Lecture Series, 1–58.

Chuahy, F.D.F., & Kokjohn, S. L. (2017). High efficiency dual-fuel combustion through

thermochemical recovery and diesel reforming. Applied Energy, 195, 503–522.

Claude, V., Courson, C., Kӧhler, M., & Lambert, S. D. (2016). Overview and essentials of

biomass gasification technologies and their catalytic cleaning methods. Energy &

Fuels, 30 (11), 8791–8814.

Coats, A.W., & Redfen, J. P. (1963). Thermogravimetric analysis: A review. Analyst, 88

(1053), 906–924.

Cullity, B. D. (1978). Elements of X-ray Diffraction (2nd ed.). Reading, Massachusetts:

Addison-Wesley Publishing Company Inc.

D’Cruz, A., Kulkarni, M.G., Meher, L.C., & Dalai, A. K. (2007). Synthesis of biodiesel from

canola oil using heterogeneous base catalyst. Journal of American Chemists’ Society,

84 (10), 937–943.

De Boer, J.H., Lippens, B.C., Lisen, B.G., Broekhoff, J.C.P., van den Heuvel, A., & Osinga,

T. J. (1966). The t-curve of multimolecular N2-adosorption. Journal of Colloid and

Interface Science, 21, 405–414.

Demirbaş, A. (2009). Political, economic and environmental impacts of biofuels: A review.

Applied Energy, 86, S108–S117.

Demirbaş, A. (1998). Fuel properties and calculation of higher heating values of vegetable oils.

219
Fuel, 77 (9/10), 1117–1120.

Demirbaş, M.F., Balat, M., & Balat, H. (2009). Potential contribution of biomass to the

sustainable energy development. Energy Conversion and Management, 50, 1746–1760.

Desjardins R. (1997). Water treatment. Editions of University of Montreal, Canada: Presses

inter Polytechnique.

Dharma, S., Masjuki, H.H., Ong, H.C., Sebayang, A.H., Silitonga, A.S., & Kusumo, F. (2016).

Optimization of biodiesel production process for mixed Jatropha curcas–Ceiba

pentandra biodiesel using response surface methodology. Energy Conversion and

Management, 178–190.

Dhawane, S.H., Kumar, T., & Halder, G. (2015). Central composite design approach towards

optimization of flamboyant pods derived steam activated carbon for its use as

heterogeneous catalyst in transesterification of Hevea brasiliensis oil. Energy

Conversion and Management, 100, 277–287.

Dhawane, S.H., Kumar, T., & Halder, G. (2016). Parametric effects and optimization on

synthesis of Iron (II) doped carbonaceous catalyst for the production of biodiesel.

Energy Conversion and Management, 122, 310–320.

Di Serio, M., Tesser, R., Pengmei, L., & Santacesaria, E. (2008). Heterogeneous catalysts for

biodiesel production. Energy & Fuels, 22 (1), 207–217.

Dizge, N. & Keskinler, B. (2008). Enzymatic production of biodiesel from canola oil using

immobilized lipase. Biomass and Bioenergy, 32 (12), 1274–1278.

Dwivedi, G., Jain, S., & Sharma, M. P. (2013). Diesel engine performance and emission

analysis using biodiesel from various oil sources-reviews. Journal of Materials and

Environmental Science, 4 (4), 434–447.

Dwivedi, G., & Sharma, M. P. (2015). Application of Box-Behnken design in optimization of

biodiesel yield from Pongamia oil and its stability analysis. Fuel, 145, 256–262.

220
Ebewele, R.O., Iyayi, A.F., & Hymore, F. K. (2010a). Considerations of the extraction process

and potential technical applications of Nigerian rubber seed oil. International Journal

of Physical Sciences, 5 (6), 826–831.

Ebewele, R.O., Iyayi, A.F., & Hymore, F.K. (2010b). Deacidification of high acidic rubber

seed oil by re-esterification with glycerol. International Journal of Physical Sciences,

5 (6), 841–846.

Eka, H.D., Tajul Aris, Y., & Wan Nadiah, W. A. (2010). Potential use of Malaysia rubber

(Hevea brasiliensis) seed as food, feed and biofuel. International Food Research

Journal, 17, 527–534.

El-Seesy, A.I., Abdel-Rahman, A.K., Bady, M., & Ookawara, S. (2017). Performance,

combustion, and emission characteristics of a diesel engine fueled by biodiesel-diesel

mixtures with multi-walled carbon nanotubes additives. Energy Conversion and

Management, 135, 373–393.

Elango, T., & Senthilkumar, T. (2011). Performance and emission characteristics of CI engine

fuelled with non-edible vegetable oil and diesel blends. Journal of Engineering Science

and Technology, 6 (2), 240–250.

Elkhaleefa, A., & Shigidi, I. (2015). Optimization of sesame oil extraction process conditions.

Advances in Chemical Engineering Science, 5, 305–310.

Encinar, J.M., Susan, F., Gonzalez, J.F., & Rodriguez-Reinares, A. (2007). Ethanolysis of used

frying oils: Biodiesel preparation and characterization. Fuel Processing Technology, 88

(5), 513–522.

Felizardo, P., Correia, M.J.N., Raposo, I., Mendes, J.F., Berkemeier, R., & Bordado, J. M.

(2006). Production of biodiesel from waste frying oil. Waste Management, 26, 487–

494.

Fillières, R., Benjelloun-Mlayal, B., & Delmas, M. (1995). Ethanolysis of rapeseed oil:

221
Quantitation of ethyl esters, mono-, di-, and triglycerides and glycerol by high-

performance size-exclusion chromatography. Journal of American Oil Chemists’

Society, 72(4), 427–432.

Fogler, H. S. (2008). Elements of Chemical Reaction Engineering (4th ed.). singapore: John

Wiley & Sons (Asia).

Food and Agriculture Organization of the United Nations (2014). Natural rubber plantation

areas harvested (hectares) for all countries. http://www.factfish.com/catalog/crop%3E

[accessed 18.05.15].

Freedman, B., Butterfield, R.O., & Pryde, E. H. (1986). Transesterification kinetics of soybean

oil. Journal of American Oil Chemists’ Society, 63, 1375–1380.

Gan, M., Pan, D., Ma, L., Yue, E., & Hong, J. (2009). The kinetics of the esterification of free

fatty acids in waste cooking oil using Fe2 (SO4)3/C catalyst. Chinese Journal of

Chemical Engineering, 17 (1), 83–87.

Geo, V.E., Nagarajan, G., & Nagalingam, B. (2010). Studies on improving the performance of

rubber seed oil fuel for diesel engine with DEE port injection. Fuel, 89, 3559–3567.

Geo, V.E., Nagarajan, G., Kamalakannan, J., & Nagalingam, B. (2009). Experimental

investigations to study the characteristics of rubber-seed-oil-fueled diesel engine

supplemented with diethyl ether. Energy & Fuels, 23, 533–538.

Gimbun, J., Ali, S., Kanwal, C.C.S.C., Shah, L.A., Gazali, N.H.M., Cheng, C.K., & Nurdin, S.

(2012). Biodiesel production from rubber seed oil using a limestone based catalyst.

Advances in Materials Physics and Chemistry, 2 (4), 138–141.

Gog, A., Roman, M., Toşa, M., Paizs, C., & Irimie, F. . (2012). Biodiesel production using

enzymatic transesterification-Current state and perspectives. Renewable Energy, 39 (1),

10–16.

Gouveia de Souza, A., Oliveira Santos, J.C., Conceiçáo, M.M., Dantas Silva, M.C., & Prasad,

222
S. (2004). A thermoanalytic and kinetic study of sunflower oil. Brazilian Journal of

Chemical Engineering, 21 (2), 265–273.

Guardiola, C., Martín, J., Pla. B., & Bares, P. (2017). Cycle by cycle NOx model for diesel

engine control. Applied Thermal Engineering, 110, 1011–1020.

Gueguim Kana, E.B., Oloke, J.K., Lateef, A., & Adesiyan, M. O. (2012). Modeling and

optimization of biogas production on saw dust and other co-substrates using Artificial

Neural network and Genetic Algorithm. Renewable Energy, 46, 276–281.

Guillèn, M.D., & Cabo, N. (2002). Fourier transform infrared spectra data versus peroxide and

anisidine values to determine oxidative stability of edible oils. Food Chemistry, 77,

503–510.

Guillèn, M.D., & Ruiz, A. (2003). 1H nuclear magnetic resonance as a fast tool for determing

the composition of acyl chains in acylglycerol mixtures. European Journal of Lipid

Science and Technology, 105, 502–507.

Guo, F., & Fang, Z. (2011). Biodiesel production with solid catalyst, Biodiesel-feedstocks and

processing technologies, In: M. Stoytcheva (Ed.), ISBN: 978-953-307-713-0, InTech.

www.intechopen.com/books/biodiesel-feedstocks-and-processing-

technologies/biodiesel-production-with-solid. In M. Stoytcheva (Ed.).

Guo, F., Peng, Z.-G., Dai, J.-Y., & Xiu, Z.-L. (2010). Calcined sodium silicate as solid base

catalyst for biodiesel production. Fuel Processing Technology, 9, 322–328.

Hago, W., & Morin, A. (2010). An internal combustion engine with CO2 capture. Efficient

Hydrogen Motors, 7–9.

Harris, L. B. (1969). Adsorption of a patchwise heterogeneous surface II. Heats of adsorption

from the condensation approximation. Surface Science, 13, 377–392.

Hassan, S.N.A.M., Ishak, M.A.M., Ismail, K., Ali, S.N., & Yusop, M. F. (2014). Comparison

study of rubber seed shell and kernel (Hevea brasiliensis) as raw material for bio-oil

223
production. Energy Procedia, 52, 610–617.

Heywood, J. B. (1988). Internal Combustion Engine Fundamentals (Int. ed.). New York:

McGraw-Hill Book Company.

Hincapie, G.M., Valange, S., Barrault, J., Moreno, J.A., & Lopez, D. P. (2014). Effect of

microwave assisted system on transesterification of castor oil with ethanol. Universitas

Scientiarum, 19 (3), 193–200.

Hosamani, K.M., & Katagi, K. S. (2008). Characterization and structure elucidation of 12-

hydroxyoctadec-cis-9-enoic acid in Jatropha gossypifolia and Hevea brasiliensis seed

oils: a rich source of hydroxy fatty acid. Chemistry and Physics of Lipids, 152, 9–12.

Hosmani, S. S. (2014). Introduction to Surface Alloying of Metals. new york: Springer Science

& Business.

Hrušovský, I., Martinka, J., & Chrebet, T. (2013). Evaluation of thermal oxidation of vegetable

oils by means of safety calorimeter SEDEX. European Journal of Environmental and

Safety Sciences, 1 (1), 13–17.

Ibrahim, A.M., & Pillai, B. C. (2011). Optimization of process parameters for biodiesel

extraction from rubber seed oil using central composite design. International Journal

of Production and Management Research, 2 (1), 23–31.

Ibrahim, A. (2016). Investigating the effect of using diethyl ether as a fuel additive on diesel

engine performance and combustion. Applied Thermal Engineering, 107, 853–862.

Igeleke, C.L., & Omorusi, V. I. (2007). Review of post-harvest deterioration of rubber Seeds.

Journal of Agriculture and Social Research, 7 (2), 11–19.

Ishola, M.M., Brandberg, T., Sanni, S.A., & Taherzadeh, M. J. (2013). Biofuels in Nigeria: A

critical and strategic evaluation. Renewable Energy, 55, 554–560.

Islam, M.A., Auta, M., Kabir, G. & Hameed, B. . (2016). A thermogravimetric analysis of the

combustion kinetics of karanja (Pongamia pinnata) fruit hulls char. Bioresource

224
Technology, 200, 335–341.

Ismail, A., Van de Voort, F., Emo, G., & Sedman, J. (1993). Rapid quantitative determination

of free fatty acids in fats and oils by Fourier transform infrared spectroscopy. Journal

of American Oil Chemists’ Society, 70 (4), 335−341.

Issariyakul, T., & Dalai, A. K. (2014). Biodiesel from vegetable oils. Renewable and

Sustainable Energy Reviews, 31, 446−471.

Iyayi, A.F., Akpaka, P.O., & Ukpeoyibo, U. (2008). Rubber seed processing for value-added

latex production in Nigeria. African Journal of Agricultural Research, 3 (7), 505–509.

Iyer, K.P.D., & Kunji, A. S. (1992). Extension of Harkins-Jura adsorption isotherm to solute

adsorption. Colloids and Surfaces, 63, 235–240.

Jaya, N., & Ethirajulu, K. (2011). Kinetic studies of heterogeneously catalyzed

transesterification of cottonseed oil to biodiesel. Journal of Environmental Research

and Development, 5 (3A), 689–695.

Karger, B.L., Synder, L.R., & Horvath, C. (1973). An Introduction to Separation Science (1st

ed.). new york: John Wiley and Sons, New York.

Kawashima, A., Matsubara, K., & Honda, K. (2008). Development of heterogeneous base

catalysts for biodiesel production. Bioresource Technology, 99, 3439–3443.

Khan, F. A. (2014). Biotechnology in Medical: Sciences. Boca Raton: CRC Press.

Kim, S.J., Kim, M.Y., Jeong, S.T., Jang, M.S., & Chung, I. M. (2012). Analysis of the biomass

content of various Miscanthus genotypes for biofuel production in Korea. Industrial

Crops and Products, 38, 46–49.

Kings, A.J., Raj, R.E., Miriam, L.M.M., & Visvanathan, M. A. (2017). Cultivation, extraction

and optimization of biodiesel production from potential microalgae Euglena sanguinea

using eco-friendly natural catalyst. Energy Conversion and Management, 141, 224–

235.

225
Kittigowittana, K., Wongsakul, S., Krisdaphong, P., Jimtaisong, A., & Saewan, N. (2013).

Fatty acid composition and biological activities of seed oil from rubber cultivar RRIM

600 (Hevea brasiliensis). International Journal of Applied Research in Natural

Products, 6 (2), 1–7.

Knothe, G. (1999). Rapid monitoring of transesterification and assessing biodiesel fuel quality

by near-infrared spectroscopy using a fiber-optic probe. Journal of American Oil

Chemists’ Society, 76 (7), 795–800.

Koh, M.Y., & Mohd. Ghazi, T. I. (2011). A review of biodiesel production from Jatropha

curcas L. oil. Renewable and Sustainable Energy Reviews, 15 (5), 2240–2251.

Kostić, M.D., Joković, N.M., Stamenković, O.S., Rajković, K.M., Milić, P.S., & Veljković, V.

B. (2013). Optimization of hampseed oil extraction by n-hexane. Industrial Crops and

Products, 48, 133–143.

Kostić, M.D., Veličković, A.V., Joković, N.M., Stamenković, O.S., & Veljković, V. B. (2016).

Optimization and kinetic modeling of esterification of the oil obtained from waste plum

stones as a pretreatment step in biodiesel production. Waste Management, 48, 619–629.

Krishnakumar, U., Sivasubramanian, V., & Selvaraju, N. (2013). Physico-chemical properties

of the biodiesel extracted from rubber seed oil using solid metal oxide catalysts.

International Journal of Engineering Research and Applications, 3 (4), 2206–2209.

Kumar, A., & Sharma, S. (2011). Potential non-edible oil resources as biodiesel feedstock: An

Indian perspective. Renewable and Sustainable Energy Reviews, 15, 1791–1800.

Kumar, A., Jones, D.D., & Hanna, M. . (2009). Thermochemical biomass gasification: A

review of the current status of the technology. Energies, 2 (3), 556–581.

Kumar, S.S., & Purushothaman, K. (2010). High FFA rubber seed oil as an alternative fuel for

diesel engine-an overview. International Journal of Engineering and Science, 10 (1),

16–24.

226
Labeckas, G., & Slavinskas, S. (2006). Performance of direct-injection off-road diesel engine

on rapeseed oil. Renewable Energy, 31, 849–863.

Lee, A., Chaibakhsh, N., Rahman, M.B.A., Basri, M., & Tejo, B. A. (2010). Optimized

enzymatic synthesis of levulinate ester in solvent-free system. Industrial Crops and

Products, 32, 246–251.

Levenspiel, O. (2005). Chemical Reaction Engineering (4th ed.). New Delhi: Prentice-Hall of

India, New Delhi.

Li, .H, Pordesimo, L.O., Weiss, J., & Wilhelm, L. R. (2004). Microwave and ultrasound

assisted extraction of soybean oil. Transactions of the ASAE, 47 (4), 1187–1194.

Li, J., Zu, Y.-G., Luo, M., Gu, C.-B., Zhao, C.-J., Efferth, T., & Fu, Y.-J. (2013). Aqueous

enzymatic process assisted by microwave extraction of oil from yellow horn

(Xanthoceras sorbifolia Bunge) seed kernels and its quality evaluation. Food

Chemistry, 138, 2152–2158.

Li, X.R., Zhou, H., Su, L., Chen, Y., Qiao, Z., & Liu, F. (2016). Combustion and emission

characteristics of a lateral swirl combustion system for DI diesel engines under low

excess air ratio conditions. Fuel, 84, 672–680.

Lior, N., & Ruddy, G. J. (1988). Second law analysis of an ideal Otto cycle. Energy Conversion

and Management, 28 (4), 327–334.

Liu, X.J., He, H., Wang, Y.J., Zhu, S.L., & Piao, X. (2008). Transesterification of soybean oil

to biodiesel using CaO as a solid base catalyst. Fuel, 87, 216–221.

Lowell, S., Shields, E., Thomas, M., & Thommes, M. (2004). Characterization of Porous

Materials and Powders: Surface Area, Pore size and Density (1st ed.). Norwell: Kluwer

Academic Publishers.

Ma, F., & Hanna, M. A. (1999). Biodiesel production: a review. Bioresource Technology, 70,

1–15.

227
Ma, L., Lv, E., Du, L., Lu, J., & Ding, J. (2016). Statistical modeling/optimization and process

intensification of microwave-assisted acidified oil esterification. Energy Conversion

and Management, 122, 411–418.

Madani, M. (2004). Contribution to the study of the adsorption of imazethapyr and catalyzed

photodegradation of imazethapyr and diuron. PhD Thesis, University Mohammed V.

Faculty of Sciences, Rabat (Morocco).

Malik, U.S., Ahmed, M., Sombilla, M.A., & Cueno, S. L. (2009). Biofuels production for

smallholder producers in the greater Mekong sub-region. Applied Energy, 86, S58–68.

Man, Y. C., & Setiowaty, G. (1999). Determinations of anisidine value in thermally oxidized

palm olein by Fourier transform infrared spectroscopy. Journal of American Oil

Chemists’ Society, 76 (2), 243−247.

Marchetti, J.M., & Errazu, A. F. (2008). Esterification of free fatty acids using sulfuric acid as

catalyst in the presence of triglycerides. Biomass and Bioenergy, 32 (9), 892−895.

Mazumdar, P., Borugadda, V.B., Goud, V.V., & Sahoo, L. (2012). Physico-chemical

characteristics of Jatropha curcas L. of North East India for exploration of biodiesel.

Biomass Bioenergy, 46, 546−554.

Mechrafi, E. (2002). Adsorption, desorption and mobility of herbicides in contact with organic

and inorganic adsorbents. PhD Thesis. University Mohammed V, Faculty of Sciences,

Rabat (Morocco).

Meisami, F., Ajam, H., & Tabasizadeh, M. (2017). Thermo-economic analysis of diesel engine

fueled with blended levels of waste cooking oil biodiesel in diesel engine. Biofuels, 1–

10. doi:10.1080/17597269.2017.1284475.

Menkiti, M.C., Agu, C.M., & Udeigwe, T. K. (2015). Extraction of oil from Terminalia catappa

L.: Process parameter impacts, kinetics, and thermodynamics. Industrial Crops and

Products, 77, 713–723.

228
Misra, D. N. (1969). Adsorption on heterogeneous surfaces: A Dubinin-Radushkevich

equation. Surface Science, 18 (2), 367–372.

Mitropoulos, A. C. (2008). The Kelvin equation. Journal of Colloid and Interface Science, 317,

643–648.

Modi, M.K., Reddy, J.R.C., Rao, B.V.S.K., & Prasad, R. B. N. (2007). Lipase-mediated

conversion of vegetable oils into biodiesel using ethyl acetate as acyl acceptor.

Bioresource Technology, 98 (6), 1260–1264.

Mofijur, M., Masjuki, H.H., Kalam, M.A., Atabani, A.E., Arbab, M.I., Cheng, S.F., & G., &

S.W. (2014). Properties and use of moringa oleifera biodiesel and diesel fuel blends in

a multi cylinder diesel engine. Energy Conversion and Management, 82, 169–176.

Mohsin, R., Majid, Z.A., Shihnan, A.H., Nasri, N.S., & Sharer, Z. (2014). Effect of biodiesel

blends on engine performance and exhaust emission for diesel dual fuel engine. Energy

Conversion and Management, 88., 821–828.

Molintas, H., & Gupta, A. K. (2011). Kinetic study for the reduction of residual char particles

using oxygen and air. Applied Energy, 88 (1), 306–315.

Morshed, M., Ferdous, K., Khan, M.R., Mazumder, M.S.I., Islam, M.A., & Uddin, M. T.

(2011). Rubber seed oil as a potential source for biodiesel production in Bangladesh.

Fuel, 90, 2981–2986.

Murga, R., Ruiz, R., Beltrán, S., & Cabezas, J. L. (2000). Extraction of natural complex phenols

and tannins from grape seeds by using supercritical mixtures of carbondioxide and

alcohol. Journal of Agriculture and Food Chemistry, 48, 3408–3412.

Murillo, S., Miguez, J.L., Porteiro, J., Granada, E., & Morán, J. C. (2007). Performance and

exhaust emissions in the use of biodiesel in outboard diesel engines. Fuel, 86, 1765–

1771.

Ng, W.P.Q., Lam, H.L., & Yusup, S. (2013). Supply network synthesis on rubber seed oil

229
utilisation as potential biofuel feedstock. Energy, 55, 82–88.

Nguyen, C., & Do, D. D. (2001). The Dubinin-Radushkevich equation and the underlying

microscopic adsorption description. Carbon, 39, 1327–1336.

NIS (1992). Nigerian Industrial Standards. Standards for edible vegetable oils.

Nisar, J., Razaq, R., Farooq, M., Iqbal, M., Khan, R.A., Sayed, M., Shah, A., & Rahman, I.

(2017). Enhanced biodiesel production from Jatropha oil using calcined waste animal

bones as catalyst. Renewable Energy, 101, 111–119.

Norazahar, N., Yusup, S., Ahmad, M.M., Abubakar, S., & Ahmad, J. (2012). Parametric

optimization of kapok (Ceiba pentandra) oil methyl ester production using Taguchi

approach. International Journal of Energy and Environment Engineering, 6 (6), 541–

548.

Noureddini, H., Gao, X., & Philkana, R. S. (2005). Immobilized pseudomonas cepecia lipase

for biodiesel fuel production from soybean oil. Bioresource Technology, 96 (7), 769–

777.

Oderinde, R.A., Ajayi, I.A., & Adewuyi, A. (2009). Characterization of seed and seed oil of

Hura crepitans and the kinetics of degradation of the oil during heating. Electronic

Journal of Environmental, Agricultural and Food Chemistry, 8 (3), 201–208.

Ogbu, I.M., & Ajiwe, V. I. E. (2016). FTIR studies of thermal stability of the oils and methyl

esters from Afzelia africana and Hura crepitans seeds. Renewable Energy, 96, 203–208.

Ohimain, E. I. (2013). A review of the Nigeria biofuel policy and incentives (2007). Renewable

Energy and Sustainable Energy Review, 22, 246–256.

Okieimen, F.E., Okieimen, C.O., & Ojokoh, F. I. (2005a). Rubber seed shell carbon as

sequestrant of heavy metals and organic compounds from acqueous solution. Indian

Journal of Chemical Technology, 12, 181–186.

Okieimen, F.E., Pavithran, C., & Bakare, I. O. (2005b). Epoxidation and hydroxiation of rubber

230
seed oil: one-pot multi-step reactions. European Journal of Lipid Science and

Technology, 107, 330–336.

Olajide, J.O., Igbeka, J.C., Afolabi, T.J., & Emiola, O. A. (2007). Prediction of oil yield from

groundnut kernels in an hydraulic press using artificial neural network (ANN). Journal

of Food Engineering, 81 (4), 643–646.

Om Tapanes, N.C., Gomes Aranda, D.A., de Mesquita Varneiro, J.W., & Ceva Antunes, O. A.

(2008). Transesterification of jatropha curcas oil glycerides: Theoretical and

experimental studies of biodiesel reaction. Fuel, 87 (10/11), 2286–2295.

Omo-Ikerodah, E.E., Omokhafe, K.O., Akpobome, F.A., & Mokwunye, M. U. (2009). An

overview of the potentials of natural rubber (Hevea brasiliensis) engineering for the

production of valuable proteins. African Journal of Biotechnology, 8 (25), 7303–7307.

Omorogbe, S.O., Ikhuoria, E.U., Aigbodion, A.I., Obazee, E.O., & Momodu, V. M. (2013).

Production of rubber seed oil based biodiesel using different catalysis. Current

Research in Chemistry, 5 (1), 11–18.

Onoji, S.E., Iyuke, S.E., Igbafe, A.I., & Nkazi, D. B. (2016a). Rubber seed oil: A potential

renewable source of biodiesel for sustainable development in sub-Saharan Africa.

Energy Conversion and Management, 110, 125–134.

Onoji, S.E., Iyuke, S.E., & Igbafe, A. I. (2016b). Hevea brasiliensis (Rubber seed) oil:

Extraction, characterization, and kinetics of thermo-oxidative degradation using

classical chemical methods. Energy & Fuels, 30 (12), 10555–10567.

Onoji, S.E., Iyuke, S.E., Igbafe, A.I., & Daramola, M. O. (2017a). Hevea brasiliensis (Rubber

seed) oil: Modeling and optimization of extraction process parameters using response

surface methodology and artificial neural network techniques. Biofuels, 1–11.

doi:10.1080/17597269.2017.1338122.

Onoji, S.E., Iyuke, S.E., Igbafe, A.I., & Daramola, M. O. (2017b). Transesterification of rubber

231
seed oil to biodiesel over a calcined waste rubber seed shell catalyst: Modeling and

optimization of process variables. Energy & Fuels, 31 (6), 6109–6119.

Oyedeji, F., & Oderinde, R. (2006). Characterization of isopropanol extracted vegetable oils.

Journal of Applied Science, 6, 2510–2513.

Özener, O., Yüksek, L., Ergenç, A.T., & Özkan, M. (2014). Effects of soybean biodiesel on a

DI diesel engine performance, emission and combustion characteristics. Fuel, 115,

875–883.

Özkan, M. (2015). A comparative study on energy and exergy analyses of a CI engine

performed with different multiple injection strategies at port load: Effect of injection

pressure. Entropy, 17 (1), 244–263.

Ozsezen, A.N., & Canakci, M. (2011). Determination of performance and combustion

characteristics of a diesel engine fueled with canola and waste palm oil methyl esters.

Energy Conversion and Management, 52, 108–116.

Peduzzi, E., Boissonnet, G., & Maréchal, F. (2016). Biomass modelling: Estimating

thermodynamic properties from the elemental composition. Fuel, 181, 201–217.

Piker, A., Tabah, B., Perkas, N., & Gedanken, A. (2016). A green and low-cost room

temperature biodiesel production method from waste oil using egg shells as catalyst.

Fuel, 182, 34–41.

Polany, M., Smisek, M., & Cerney, S. (1970). Active carbon manufacture, properties and

application. Analytical Chemistry, 42 (14), 81A–81A.

Portet-Koltalo, F., & Machour, N. (2013). Analytical methodologies for the control of particle

phase polycyclic aromatic compounds from diesel engine exhaust. In S. Bari (Ed.),

Diesel engine-combustion, emission and condition monitoring (p. 91–117.). InTech e-

Publishing company. https://doi.org/10.5772/53725

Prabhahar, M., Manohar, R.M., & Sendilvelan, S. (2012). Performance and emission studies

232
of a diesel engine with pongamia methyl ester at different load conditions. International

Journal of Engineering Research and Applications, 2 (3), 2707–2713.

Raheman, H., & Ghadge, S. (2008). Performance of diesel engine with biodiesel at varying

compression ratio and ignition timing. Fuel 87, 2659–2666.

Rahman, S.M.A., Masjuki, H.H., Kalam, M.A., Abedin, M.J., Sanjid, A., & Rahman, M. M.

(2014). Assessing idling effects on a compression ignition engine fueled with Jatropha

and Palm biodiesel blends. Renewable Energy, 68, 644–650.

Rakopoulos, C.D., & Giakoumis, E. G. (1997). Development of cumulative and availability

rate balances in a multi-cylinder turbocharged indirect injection diesel engine. Energy

Conversion and Management, 38 (4), 347–369.

Rakopoulos, C.D., & Giakoumis, E. G. (2006). Second-law analyses applied to internal

combustion engines operation. Progress in Energy and Combustion Science, 32 (1), 2–

47.

Ramadhas, A.S., Jayaraj, S., & Muraleedharan, C. (2005). Biodiesel production from high FFA

rubber seed oil. Fuel, 84 (4), 335–340.

Ramadhas, A.S., Muraleedharan, C., & Jayaraj, S. (2005a). Performance and emission

evaluation of a diesel engine fueled with methyl esters of rubber seed oil. Renewable

Energy, 30 (12), 1789–1800.

Ramadhas, A.S., Jayaraj, S., & Muraleedharan, C. (2005b). Characterization and effect of using

rubber seed oil as fuel in the compression ignition engines. Renewable Energy, 30, 795–

803.

Ramaswamy, H.S., Van De Voort, F.R., & Ghazala, S. (1989). An analysis of TDT and

Arrhenius methods for handling process and kinetic data. Journal of Food Science, 54

(5), 1322–1326.

Raventós, M., Duarte, S., & Alarcón, R. (2002). Application and possibilities of supercritical

233
CO2 extraction in food processing industry: an overview. Food Science Technology

International, 8, 269–284.

Reshad, A.S., Tiwari, P., & Goud, V. V. (2015). Extraction of oil from rubber seeds for

biodiesel application: Optimization of parameters. Fuel, 150, 636–644.

Rodríguez-Solana, R., Salgado, J.M., Domínguez, J.M., & Cortés-Diéguez, S. (2014).

Estragole quantity optimization from fennel seeds by supercritical fluid extraction

(carbon dioxide– methanol) using a Box–Behnken design. Characterization of fennel

extract. Industrial Crops and Products Crops and Products, 60, 186–192.

Roschat, W., Kacha, M., Yoosuk, B., Sudyoadsuk, T., & Promarak, V. (2012). Biodiesel

production based on heterogeneous process catalyzed by solid waste coral fragment.

Fuel, 98, 194–202.

Sabarish, C.S., Sebastian, J., & Muraleedharan, C. (2016). Extraction of oil from rubber seed

through hydraulic press and kinetic study of acid esterification of rubber seed oil.

Procedia Technology, 25, 1006–1013.

Sadowska, J., Johansson, B., Johannessen, E., Friman, R., Broniarz-Press, L., & Rosenholm, J.

B. (2008). Characterization of ozonated vegetable oils by spectroscopic and

chromatographic methods. Chemistry and Physics of Lipids, 151, 85–91.

Sajjadi, B., Raman, A.A.A., & Arandiyan, H. (2016). A comprehensive review on properties

of edible and non-edible vegetable oil-based biodiesel: Composition, specifications and

prediction models. Renewable and Sustainable Energy Reviews, 63, 62–92.

Sanjel, N., Gu, J.H., & Oh, S. C. (2014). Transesterification kinetics of waste vegetable oil in

supercritical alcohols. Energies, 7, 2095–2106.

Santos, J.C.O., Dos Santos, I.M.G., De Souza, A.G., Prasad, S., & Dos Santos, A. V. (2002).

Thermal stability and kinetic study on thermal decomposition of commercial edible oils

by thermogravimetry. Journal of Food Science, 67 (4), 1393–1398.

234
Sasmal, S., Goud, V.V, & Mohanty, K. (2012). Characterization of biomass available in the

region of North-East India for production of biofuels. Biomass and Bioenergy, 45, 212–

220.

Sayyar, S., Abidin, Z.Z., Yunus, R., & Muhammad, A. (2009). Extraction of oil from jatropha

seeds-optimization and kinetics. American Journal of Applied Science, 6, 1390–1395.

Schwab, A.W., Bagby, M.O. & Freedman, B. (1987). Preparation and properties of diesel fuels

from vegetable oils. Fuel, 66, 1372–1378.

Senthilkumar, S., & Purushothaman, K. (2012). High FFA rubber seed oil as an alternative fuel

for diesel engine-An overview. International Journal of Engineering and Science, 1

(10), 16–24.

Shanavas, S., Kunji, A.S., Varghese, H.T., & Panicker, C. Y. (2011). Comparison of Langmuir

and Harkins-Jura adsorption isotherms for the determination of surface area of solids.

Oriental Journal of Chemistry, 27 (1), 245–252.

Shokri, A., Hatami, T., & Khamforoush, M. (2011). Near critical carbon dioxide extraction of

Anise (Pimpinella Anisum L.) seed: Mathematical and artificial neural network

modeling. Journal of Supercritical Fluids, 58, 49–57.

Siboni, S., & Volpe, C. D. (2008). Some mathematical aspects of the Kelvin equation.

Computers and Mathematics with Applications, 55 (1), 51–65.

Singh, J., & Bargale, P. C. (2000). Development of a small capacity double stage compression

screw press for oil expression. Journal of Food Engineering, 43 (2), 75–82.

Skoog, D.A., Holler, F.J., & Crouch, S. R. (2007). Principle of Instrumental Analysis (6th Ed.).

Thomson Publishing, USA.

Slopiecka, K., Bartocci, P., & Fantozzi, F. (2012). Thermogravimetric analysis and kinetic

study of poplar wood pyrolysis. Applied Energy, 97, 491–497.

Smith, B. C. (2011). Fundamentals of Fourier Transform Infrared Spectroscopy (2nd ed.).

235
Boca Raton: CRC Press.

SON (2000). Standard Organization of Nigeria. Standards for edible refined palm oil and its

processed form.

Stepanov, V.S. (1995). Chemical energies and exergies of fuels. Energy 20 (3), 235–242.

Storck, S., Bretinger, H., & Maier, W. F. (1998). Characterization of micro- and mesoporous

solids by physisorption methods and pore-size analysis. Applied Catalysis A: General,

174, 137–146.

Subramaniam, D., Murugesan, A., Avinash, A., & Kumaravel, A. (2013). Bio-diesel

production and its engine characteristics: An expatiate view. Renewable and

Sustainable Energy Reviews, 22, 361–370.

Subroto, E., Manurung, R., Heeres, H.J., & Broekhuis, A. A. (2015). Optimization of

mechanical oil extraction from Jatropha curcas L. kernel using response surface

method. Industrial Crops and Products, 63, 294–302.

Sun, K., & Jiang, J. c. (2010). Preparation and characterization of activated carbon from rubber

seed shell by physical activation with steam. Biomass and Bioenergy, 34, 539–544.

Syazwani, O.N., Rashid, U., & Yap, Y. H. T. (2015). Low-cost solid catalyst derived from

waste Cyrtopleura costata (Angel Wing Shell) for biodiesel production using

microalgae oil. Energy Conversion and Management, 101, 749–756.

Takase, M., Zhao, T., Zhang, M., Chen, Y., Liu, H., Yang, L., & Wu, X. (2015). An expatiate

review of neem, jatropha, rubber and karanja as multipurpose non-edible biodiesel

resources and comparison of their fuel, engine and emission properties. Renewable and

Sustainable Energy Reviews, 43, 495–520.

Tang, Z.-X., Clavean, D., Corcuff, R., Belkacemi, K., & Arul, J. (2008). Preparation of nano-

CaO using thermal-decomposition method. Materials Letters, 62, 2096–2098.

TecQuipment Ltd. (2011). TD200 Small Engine Test Set User Guide Manual for TecQuipment

236
Test Engines. Nottingham, UK.

Thanh, L.T., Okitsu, K., Boi, L.V., & Maeda, Y. (2012). Catalytic technologies for biodiesel

fuel production and utilization of glycerol: a review. Catalysts, 2, 191–222.

Toscano, G., Riva, G., Pedretti, E.F., & Duca, D. (2012). Vegetable oil and fat viscosity

forecast models based on iodine number and saponification number. Biomass

Bioenergy, 46, 511–516.

UNFCCC (2015). United Nations Framework Convention on Climate Change. Adoption of the

Paris Agreement. http://unfccc.int/resource/dos/2015/COP21/eng/109r01.pdf

Ude, C.N., Onukwuli, O.D., Nwobi-Okoye, C., Anisiji, O.E., Atuanya, C.U., & Menkiti, M.C.

(2017). Performance evaluation of cottonseed oil methyl esters produced using CaO

and prediction with an artificial neural network. Biofuel, 1–8, doi:

10.1080/17597269.2017.1345355.

Uprety, B.K., Chaiwong, W., Ewelike, C., & Rakshit, S. K. (2016). Biodiesel production using

heterogeneous catalysts including wood ash and the importance of enhancing byproduct

glycerol purity. Energy Conversion and Management, 115, 191–199.

Valente, V.S.B., Vieira, A. dos S., & Teixeira, R. M. (2016). Physicochemical characterization

of commercial biodiesel/diesel blends and evaluation of unconventional spectroscopic

vibrational techniques in the monitoring of their oxidation and hydrolysis during

storage. Energy & Fuels, 30, 8399–8409.

Van Boekel, M. A. J. S. (1996). Statistical aspects of kinetic modeling for food science

problems. Journal of Food Science, 6 (3), 477–489.

Van Dongen, R.H. & Broekhoff, J. C. P. (1969). The isosteric heat of adsorption on

homogeneous and patchwise heterogeneous surfaces. Surface Science, 18, 462–469.

Vasudevan, M., Sakaria, P.L., Bhatt, A.S., Mody, H.M., & Bajaj, H. C. (2011). Effect of

concentration of aminopropyl groups on the surface of MCM-41 on adsorption of Cu2+.

237
Industrial & Engineering Chemistry Research, 50 (19), 11432–11439.

Venkatachalam, P., Geetha, N., Sangeetha, P., & Thulaseedharan, A. (2013). Natural rubber

producing plants: An overview. African Journal of Biotechnology, 12 (12), 1297–1310.

Verma, P., Sharma, M.P., & Dwivedi, G. (2016). Impact of alcohol on biodiesel production

and properties. Renewable and Sustainable Energy Reviews, 56, 319–333.

Vipin, V.C., Sebastian, J., Muraleedharan, C., & Santhiagu, A. (2016). Enzymatic

transesterification of rubber seed oil using rhizopus oryzae lipase. Procedia

Technology, 25, 1014–1021.

Widayat, W., & Kiono, B. F. T. (2012). Ultrasound assisted esterification of rubber seed oil for

biodiesel production. International Journal of Renewable Energy Development, 1, 1–

5.

Widayat, W., Wibowo, A.D.K., & Hadiyanto (2013). Study on production process of biodiesel

from rubber seed (Hevea brasiliensis) by in situ (trans) esterification method with acid

catalyst. Energy Procedia, 32, 64–73.

Widyarani; Ratnaningsih, E., Sanders, J.P.M., & Bruins, M. E. (2014). Biorefinery methods

for separation of protein and oil fractions from rubber seed kernel. Industrial Crops and

Products, 42, 323–332.

Willard, H.W., Merritt, Jr., Dean, J.A., & Seattle, J. (1981). Instrumental Methods of Analysis

((6th Ed.)). Belmont, CA, USA, Wadsworth Publishing Company.

Willems, P., Kuipers, N.J.M., & De Haan, A. B. (2008). Hydraulic pressing of oilseeds:

Experimental determination and modeling of yield and pressing rates. Journal of Food

Engineering, 89 (1), 8–16.

Wohlgemutgh, N. (1999). Cost benefits indicators associated with the integration of alternative

energy sources: a system approach for Carinthia, Austria. Renewable Energy, 16, 1147–

1150.

238
Wu, J. (2004). Modeling adsorption of organic compounds on activated carbon: A multivariate

approach. PhD Thesis. University of Neuchâtel, Schweiz.

Yang, L., Takase, M., Zhang, M., Zhao, T., & Wu, X. (2014). Potential non-edible oil feedstock

for biodiesel production in Africa: A survey. Renewable and Sustainable Energy

Reviews, 38, 461–477.

Yuan, X., Liu, J., Zeng, G., Shi, J., Tong, J., & Huang, G. (2008). Optimization of conversion

of waste rapeseed oil with high FFA to biodiesel using response surface methodology.

Renewable Energy, 33 (7), 1678–1684.

Zabeti, M., Daud, W.M.A.W., & Aroua, M. K. (2009). Activity of solid catalysts for biodiesel

production: A review. Fuel Processing Technology, 90, 770–777.

Zahedi, G., & Azarpour, A. (2011). Optimization of supercritical carbon dioxide extraction of

Passiflora seed oil. Journal of Supercritical Fluids, 58, 40–48.

Zhang, L., Xu, C., & Champagne, P. (2010). Overview of recent advances in thermo-chemical

conversion of biomass. Energy Conversion and Management, 51, 969–982.

Zhang, S., Zu, Y.G, Fu, Y.J., Luo, M., Zhang, D.Y., & Efferth, T. (2011). Rapid microwave-

assisted transesterification of yellow horn oil to biodiesel using a heteropolyacid solid

catalyst. Bioresource Technology 101 (3), 931–936.

Zheng, Y.L., Wiesenborn, D.P., Tostenson, K., & Kangas, N. (2003). Screw pressing of mode

and de-hulled flax seed for organic oil. Journal of the American Oil Chemists’ Society,

80 (10), 1039–1045.

Zhu, Y., Xu, J., Li, Q., & Mortimer, P. E. (2014). Investigation of rubber seed yield in

Xishuangbanna and estimation of rubber seed oil based biodiesel potential in Southeast

Asia. Energy, 69, 837–842.

239
APPENDIX

Appendix A

240
241
242
243
Appendix B

Section B.1

Rubber seed analysis:

1st Experiment:

Amount of rubber seed used = 500 g

Amount of shell present = 249.65 g

∴ Amount of kernel (solid + moisture) = 250.35 g

2nd Experiment:

Amount of rubber seed used = 500 g

Amount of shell present = 248.65 g

Amount of kernel (solid + moisture) = 251.35 g

249.65
Mass % shell (1st reading) =  100  49.93%
500

248.65
Mass % shell (2nd reading) =  100  49.73%
500

Average mass % shell =


1
49.93  49.73  49.83%
2

Standard deviation,  

 X X 2

n 1 (B.1)

Where X = mean of the data (49.93, 49.73)

n = number of data

 Standard deviation for shell = ± 0.14

Mass % of kernel and moisture calculations:

Average mass of kernel and moisture =


1
250.35  251.35  250.85g
2

250.85
Mass % (kernel + moisture) =  100  50.17%
500

244
Determination of moisture content of the kernel:

1st Experiment:

Amount of kernel used = 100 g

Amount of kernel after drying to constant mass at 105 oC for 5 h = 90.35 g

Mass of moisture = (100 – 90.35) g = 9.65 g

9.65
 Mass % moisture =  100  9.65%
100

2nd Experiment:

Amount of kernel used = 100 g

Amount of kernel after drying to constant mass at 105 oC for 5 h = 90.25 g

Mass of moisture = (100 – 90.25) g = 9.75 g

9.75
 Mass % moisture =  100  9.75%
100

 Average mass % moisture =


1
9.65  9.75  9.7 ± 0.07%
2

Therefore, average mass % kernel in the seed = 50.17 – 9.7 = 40.47± 0.07%

245
Appendix C

Section C.1

This section shows the experimental setup for rubber seed oil extraction using the n-hexane–

soxhlet extractor.

Weighing of milled kernel Milled kernel in muslin cloth

Figure C.1: Preparation of milled kernel for oil extraction

Figure C.2: Oil soxhlet extractor used for the study


246
Figure C.3: Rotary evaporator used in the study

(a) (b)

Figure C.4: Extracted rubber seed oil: (a) filtered, (b) crude

247
Appendix D

Section D.1

This section deals with the determination of the oil yield (%) using the experimental setup in

Appendix C. The oil yield was calculated based on Equation (3.1) for each particle size (0.5,

1, 1.5, 2 and 2.5 mm) considered. The experimental procedures are described in Section 3.2.2.

Calculation of oil yield for 0.5 mm particle size:

1st Experiment:

Mass of kernel charged into extractor = 50 g

Volume of n-Hexane used = 225 mL

Mass of extracted oil = 20.12 g

20.12
Oil yield = 100  40.24%
50

2nd Experiment:

Mass of kernel charged into extractor = 50 g

Volume of n-Hexane used = 225 mL

Mass of extracted oil = 20.18 g

20.18
Oil yield = 100  40.36%
50

 Average oil yield for 0.5 mm kernel size =


1
40.24  40.36  40.3 ± 0.085%
2

Similar experiments were conducted in duplicates, and the oil yields obtained for 1, 1.5, 2, and

2.5 mm particle size are 39.95, 40.25; 37.8, 37.4; 32.7, 32.3; and 29.5, 29.1%, respectively.

The average results are tabulated in Table D.1, and plotted in Figure 4.1.

248
Table D.1: Variation of oil yield with kernel particle size

Particle size (mm) 0.5 1 1.5 2 2.5

Oil yield (%) 40.3±0.085 40.1±0.212 37.6±0.283 32.5±0.283 29.3±0.283

Section D.2: Validation of optimal oil yield for RSM model

The RSM model predicted an optimum oil yield of 42.98% at the following conditions:

Rubber seed milled kernel = 60 g

Solvent volume = 250 mL

Extraction time = 45 min

60 g of milled kernel of 0.5 mm particle size was charged into the n-hexane extractor with 250

mL hexane. The experiment was conducted for a period of 45 min and repeated in triplicates.

The procedures in Section D.1 were followed and the average oil yield calculated is shown

below.

1st Experiment:

Mass of milled kernel used = 60 g

Mass of extracted oil = 25.572 g

25.572
Oil yield =  100  42.62%
60

2nd Experiment:

Mass of milled kernel used = 60 g

Mass of extracted oil = 25.59 g

25.59
Oil yield =  100  42.65%
60

249
3rd Experiment:

Mass of milled kernel used = 60 g

Mass of extracted oil = 25.59 g

25.59
Oil yield =  100  42.65%
60

Average oil yield =


1
42.62  42.65  42.65  42.64  0.017%
2

The experimental validated oil yield value of 42.64% is close to the model predicted optimum

value of 42.98% within an error of 0.8%.

Section D.3: validation of optimal oil yield for ANN model

The ANN model predicted on optimum oil yield of 43% at the following conditions:

Rubber seed milled kernel = 40 g

Solvent volume = 202 mL

Extraction time = 49.99 min

Using these optimum conditions, the experiment was conducted in triplicates as described in

Section D.1.

1st Experiment:

Mass of milled kernel used = 40 g

Mass of extracted oil = 17.148 g

17.148
Oil yield = 100  42.87%
40

2nd Experiment:

Mass of milled kernel used = 40 g

Mass of extracted oil = 17.16 g

17.16
Oil yield =  100  42.9%
40
250
3rd Experiment:

Mass of milled kernel used = 40 g

Mass of extracted oil = 17.244 g

17.244
Oil yield =  100  43.11%
40

Average oil yield =


1
42.87  42.9  43.11  42.96  0.131%
2

The experimental validated oil yield value of 42.96% is close to the model predicted optimum

value of 43% within an error of 0.09%.

Appendix E

Section E.1

This section deals with the procedures and calculations of the physicochemical properties of

the extracted rubber seed oil based on the procedures described in ASTM and AOAC methods.

Section E.1.1 Determination of the density (ρ) and specific gravity (SG) of the oil

This was determined using 25 mL capacity density bottles at 15 oC. The SG was calculated as

the ratio of the density of the oil to that of water at 15 oC.

mass
 (E.1)
volume

o
SG  (E.2)
w

Section E.1.2 Determination of the oAPI gravity

Equation (E.3) determined the oAPI of the oil as shown below.

141.5
o
API   131.5 (E.3)
SG @15o C

251
Section E.1.3 Conradson carbon residue of the seed oil

The conradson carbon residue was determined following the procedures described in ASTM

D189-IP 13. The determination of carbon residue left after evaporation and pyrolysis of oil is

intended to provide some indication of its relative coke-forming propensity when used as

source material for fuel. In this study, 3 ± 0.1 g of oil was weighed and placed in a porcelain

crucible containing 0.3149 g of glass bead to maintain uniform heating (Onoji et al., 2016b).

The crucible was placed in the center of the skidmore iron crucible (Stanhope-Seta apparatus)

with a lid. Heat was applied and pre-ignition period observed was 30 ± 5 min when smoke

appeared above the chimney. The experiment lasted for 60 ± 5 min in an air-free environment.

At the end of the specified heating period, the test crucible containing the carbonaceous residue

was cooled in a desiccator and weighed again. The residue remaining was calculated as a

percentage of the original sample, and reported as non-combustible carbon residue. The

experiment was carried out in duplicate and average value computed as shown below.

1st Experiment:

Amount of oil and glass bead placed in a porcelain crucible = 3.3149 g

Mass of oil = 3.0 g

Mass of carbon residue and bead after evaporation and pyrolysis of the oil = 0.3263 g

 Mass of carbon residue = (0.3263 – 0.3149) g = 0.0114 g

0.0114
% mass carbon residue =  100  0.38%
3

2nd Experiment:

Amount of oil and glass bead placed in a porcelain crucible = 3.3149 g

Mass of oil = 3.0 g

Mass of carbon residue and bead after evaporation and pyrolysis of the oil = 0.3275 g

 Mass of carbon residue = (0.3275 – 0.3149) g = 0.0126 g


252
0.0126
% mass carbon residue = 100  0.42%
3

 Average mass % carbon residue =


1
0.38  0.42  0.4%
2

Section E.1.4 Moisture content analysis of the seed oil

The moisture content was determined using a standard laboratory U CLEAR oven (Model:

DHG-9053A) operated at 105 oC. About 20 g of the sample was placed in a clean porcelain

crucible, and weighed. The process of heating and cooling was repeated every 10 min after

weighing until constant weight was obtained. All analyses were done in duplicates, and

moisture content was determined by calculating the weight difference of the sample before and

after oven drying as documented below.

1st Experiment:

Amount of oil used = 20 g

Amount of oil after drying to constant mass at 105 oC = 19.658 g

Amount of moisture removed = (20 – 19.658) g = 0.342 g

0.342
% moisture content =  100  1.71%
20

2nd Experiment:

Amount of oil used = 20 g

Amount of oil after drying to constant mass at 105 oC = 19.65 g

Amount of moisture removed = (20 – 19.65) g = 0.35 g

0.35
% moisture content =  100  1.75%
20

 Average mass % moisture =


1
1.711.75  1.73%
2

253
Section E.1.5 Ash content analysis of the seed oil

The ash content was determined by burning 20 g of oven-dried sample in a porcelain crucible

placed in an electric muffle furnace (Carbolite, Parson Lane, Hope Valley S33 6RB, England,

Model: RWF 12/5) maintained at 550 ± 25 oC for an initial 20 min, cooled in a desiccator and

weighed again. The heating and cooling process was repeated every 10-min interval until the

carbonaceous residue was reduced to an ash. The results were expressed on a dry weight basis

according to procedures prescribed in ASTM D 482-IP 4, and documented as shown below.

1st Experiment:

Amount of oven-dried oil placed in a porcelain crucible = 20 g

Amount of cooled ash content after combustion at 550 ± 25 oC = 0.00018 g

0.00018
% Ash content of oil = 100  0.0009%
20

2nd Experiment:

Amount of oven-dried oil placed in a porcelain crucible = 20 g

Amount of cooled ash content after combustion at 550 ± 25 oC = 0.00022 g

0.00022
% Ash content of oil = 100  0.0011%
20

 Average mass % ash =


1
0.0009  0.0011  0.001%
2

Section E.1.6 Determination of the seed oil volatile matter

The percentage of volatile matter in the oil was obtained by using Equation (E.4).

𝑉𝑜𝑙𝑎𝑡𝑖𝑙𝑒 𝑚𝑎𝑡𝑡𝑒𝑟 (%) = 100 − (𝑐𝑎𝑟𝑏𝑜𝑛 𝑟𝑒𝑠𝑖𝑑𝑢𝑒 + 𝑚𝑜𝑖𝑠𝑡𝑢𝑟𝑒 + 𝑎𝑠ℎ) (E.4)

Therefore, % volatile matter = 100 – (0.4 + 1.73 + 0.001) = 97.869%

254
Section E.1.7 Determination of the peroxide value (PV) of the oil

About 2 g of oil sample was weighed into a 250-mL pyrex flask and 40 mL of solvent mixture

(2:1 glacial acetic acid/chloroform) was added. The content was swirled until it was dissolved

completely and 2 g of KI powder was then added. The mixture was boiled briskly in a water

bath maintained at 70 oC for 1 min. The boiled mixture was added to a flask containing 40 mL

of 5% KI and the resultant mixture was then washed with 50 mL of distilled water. The content

of the flask was titrated with 0.004M Na2S2O3 solution using 1 mL of 0.5% starch solution as

indicator. The experiment was repeated and peroxide value (mequiv O2/kg oil) was calculated

as indicated in Equation (E.5).

V×M
Peroxide value = (E.5)
w

where:

V = volume of Na2S2O3 used for titration

M = molarity of Na2S2O3

w = weight of oil sample

Section E.1.8 Determination of the iodine value (IV) of the oil (Wijs' method)

About 0.26 g of oil was dissolved in 10 mL of carbon tetrachloride and agitated in a stopper

flask. 20 mL of Wijs solution was added and allowed to stand for 30 min in the dark at room

temperature. 20 mL of 10% potassium iodide solution and 100 mL distilled water was added

to the mixture in the flask and titrated with 0.1M Na2S2O3 (sodium thiosulphate) standard

solution using 1 mL of 0.5% starch solution as indicator. The iodine value was calculated using

Equation (E.6).

12.69  M  B  S 
IV  (E.6)
w

where:

255
B = volume of Na2S2O3 titrated in blank

S = volume of Na2S2O3 titrated in test

M = molarity of standard Na2S2O3

w = mass of oil used

Section E.1.9 Determination of the refractive index (RI) of the oil

A digital Abbe Refractometer (Model: 60/ED) was used to determine the refractive index of

the oil that was maintained at 20 oC in a water bath to keep the temperature of the glass slide

uniform. Then, few drop of oil sample was dropped into the glass slide of the refractometer by

means of syringe and needle. This was repeated twice and the values were converted using a

calibrated chart, and average value was recorded.

Section E.1.10 Determination of the saponification value (SV) of the oil

About 4 g of oil was weighed into a 250-mL pyrex flask and 50 mL of 0.5M ethanolic KOH

was added. The setup was connected to a reflux condenser and heated gently for 30 min. The

mixture formed (soap solution) was cooled and titrated with 0.5M HCl solution using 2 drops

of phenolphthalein indicator. The endpoint is reached when the pink colour of the indicator

disappeared. The experiment was conducted for the blank titration and repeated. The

saponification value was calculated according to Equation (E.7).

56.1  M  B  S 
SV  (E.7)
w

where:

B = volume of HCl used during blank titration

S = volume of HCl used for test

M = molarity of standard HCl

w = mass of oil used in gram; Molecular mass of KOH = 56.1


256
Section E.1.11 Determination of acid value (AV) of the oil

The acid value was determined by dissolving 5 g of oil sample in a hot mixture of 25 mL (95%

vol/vol) diethyl ether and 25 mL ethanol in a 250-mL flask. The hot solution was titrated with

0.1M KOH solution using 3 drops of phenolphthalein indicator. The acid value was calculated

using Equation (E.8). It is important to note that %FFA is half the acid value.

56.1  M  V
AV  (E.8)
w

where:

M = molarity of standard KOH

V = volume of titrant (KOH) in cm3

w = mass of oil sample

56.1 = molecular mass of KOH

Section E.1.12 Determination of the pH value of the oil

About 2 g of the oil sample was poured into a clean dry 100 mL beaker and 12.5 mL of hot

distilled water was added to the sample in the beaker and stirred gradually. The resulting

solution was cooled in a water bath at temperature of 28 oC. A standardized pH meter with

buffer solution and the electrode immersed into the sample was used to read up the pH value

of the oil sample. The experiment was repeated and the average value recorded.

Section E.2

This section deals with the effects of temperature and time parameters on the peroxide, iodine,

and refractive index values of the oil during heat treatment. At each temperature and time

indicated, the values are determined as described in Section E.1, and they are recorded in Table

E.1. The average values and their standard deviations are presented in Table 4.7.

257
Table E.1: Variation of peroxide, iodine, and refractive index values of rubber seed oil

with temperature and time

Temp Time Peroxide (mequiv O2/kg Iodine (I2/100 g oil) Refractive index @ 20 oC
(oC) (min) oil)
1st 2nd 1st 2nd 1st 2nd
Reading Reading Reading Reading Reading Reading
100 30 3.84 4.20 118.50 116.94 1.4555 1.4747
60 4.41 4.59 117.31 117.55 1.3982 1.4662
120 4.80 4.82 117.20 117.28 1.4018 1.4276
180 4.84 5.16 117.00 117.04 1.3820 1.3208
240 5.20 5.60 116.51 117.45 1.3001 1.3021
300 5.50 5.72 116.80 116.88 1.2580 1.2352
150 30 4.35 4.41 117.30 117.82 1.4601 1.4659
60 4.60 4.72 117.20 117.44 1.4281 1.4299
120 4.89 5.07 116.95 117.29 1.3991 1.4269
180 5.20 5.28 116.61 117.07 1.3811 1.3889
240 5.31 5.81 115.98 117.22 1.3700 1.3602
300 5.70 5.94 114.80 116.00 1.3420 1.3452
200 30 4.53 4.71 117.31 117.45 1.3628 1.3666
60 4.82 4.94 117.00 117.04 1.3225 1.3235
120 4.91 5.09 116.34 116.88 1.2970 1.3012
180 5.30 5.38 116.10 116.54 1.2900 1.2922
240 5.70 5.82 114.98 116.38 1.2815 1.2853
300 5.91 5.97 115.20 115.28 1.2431 1.2435
250 30 4.72 4.76 116.85 117.57 1.3001 1.3039
60 4.85 4.95 116.40 116.80 1.2745 1.2755
120 4.90 5.06 116.00 116.40 1.2326 1.2343
180 5.15 5.25 115.20 115.60 1.2319 1.2303
240 5.45 5.55 113.88 115.76 1.2254 1.2170
300 5.60 5.80 114.32 114.64 1.2085 1.1937

258
RAW RSS RSS CATALYST
o
Calcined @ 800 C

Figure E.1: Calcination of rubber seed shell (RSS) catalyst for biodiesel production

Appendix F

Section F.1

Esterification of rubber seed oil:

Prior to transesterification reaction, the acid value of the oil is reduced to ≤ 1% FFA to avoid

the consumption of the base catalyst during biodiesel production through saponification

reaction (soap formation). The initial value of the rubber seed oil was 18.02 mg KOH/g oil. For

each concentration of H2SO4 catalyst (1.5, 3, 4.5, and 6 % vol/vol) used, the experiment was

carried out in duplicates as described in Section 3.8.1. The acid value in each case was

determined as discussed in Section E.1.11 of Appendix E, and the values recorded in Table

F.1.

Table F.1: Esterification parameters of the rubber seed oil (Onoji et al., 2017b)

Std run H2SO4 (% vol/vol) Acid value (mg KOH/g oil) FFA (%)
0 0 18.20±0.141 9.10±0.07
1 1.5 12.50±0.125 6.25±0.0625
2 3 4.82±0.021 2.41±0.01
3 4.5 3.98±0.01 1.96±0.005
4 6 1.78±0.005 0.86±0.0025

259
Figure F.1: Esterified oil separation into two layers

Figure F.2: Gravity settling of biodiesel, catalyst and glycerol phases

260
Figure F.3: Processed rubber seed oil methyl ester (biodiesel) for analysis

Figure F.4: Regression line (−), upper (−), and lower (−) prediction intervals and

experimental data (∎) of oxidation stability determination for biodiesel without additive

(Adapted from Botella et al., 2014)

261
Appendix G

Table G.1: Infrared group absorption frequencies (Onoji et al., 2016b)

Type of bond Functional group Type of vibration Wavenumber (cm-1) Wavelength (μ) Intensity (I)

C–H Alkanes stretch 3000-2850 3.33-3.51 s

–CH3 bend 1450 and 1375 6.90-7.27 m

–CH2– bend 1465 6.83 m

Alkenes stretch 3100-3000 3.23-3.33 m

bend 1700-1000 5.88-10.0 s

Aromatics stretch 3150-3050 3.17-3.28 s

out-of-plane bend 1000-700 10.0-14.3 s

Alkynes stretch ca. 3300 ca. 3.03 s

Aldehydes 2900-2800 3.45-3.57 w

2800-2700 3.57-3.70 w

C–C Alkanes not usually useful

C=C Alkenes 1680-1600 5.95-6.25 m-w

262
Aromatics 1600-1400 6.25-7.14 m-w

C≡C Alkyne 2250-2100 4.44-4.76 m-w

C=O Aldehyde 1740-1720 5.75-5.81 s

Ketone 1725-1705 5.80-5.87 s

Carboxylic acids 1725-1700 5.80-5.88 s

Esters 1750-1730 5.71-5.78 s

Amide 1700-1640 5.88-6.10 s

Anhydride ca. 1810 ca. 5.52 s

ca. 1760 ca. 5.68 s

Acyl chloride 1800 5.55 s

C–O Alcohols, Ethers, Esters, Carboxylic acids 1300-1000 7.69-10.0 S

O–H Alcohols, Phenols:

Free 3650-3600 2.74-2.78 m

H-bonded 3400-3200 2.94-3.12 m

Carboxylic acids (2) 3300-2500 3.03-4.00 m

263
N–H Primary and secondary amines ca. 3500 ca. 2.86 M

C≡N Nitriles 2260-2240 4.42-4.46 M

N=O Nitro (R–NO2) 1600-1500 6.25-6.67 s

1400-1300 7.14-7.69 s

C-X Fluoride 1400-1000 7.14-10.0 s

Chloride 800-600 12.5-16.7 s

Bromide, Iodide <600 >16.7 s

(I): s = strong, m = medium and w = weak

264
Table G.2: Assignment of signals of 1H NMR spectra for vegetable oils (hazel nut and

walnut) (Guillèn & Ruiz, 2003)

Signal Chemical shift (ppm) Functional group

1 0.83–0.93 –CH3(saturated, oleic and linoleic acyl chains)

2 0.93–1.03 –CH3 (linolenic acyl chains)

3 1.22–1.42 – (CH2)n -(acyl chains)

4 1.52–1.70 –OCO–CH2–CH2– (acyl chains)

5 1.94–2.14 –CH2–CH=CH (acyl chains)

6 2.23–2.36 –OCO–CH2– (acyl chains)

7 2.70–2.84 =HC–CH2–CH= (acyl chains)

8 4.10–4.32 –CH2OCOR (glyceryl group)

9 5.20–5.26 >CHOCOR (glyceryl group)

10 5.26–5.40 –CH=CH– (acyl chains)

Figure G.1: 13C NMR spectra of rubber seed oil (12–28 ppm)

265
Figure G.2: 13C NMR spectra of rubber seed oil (28.4–34.8 ppm)

Figure G.3: 13C NMR spectra of rubber seed oil (40–84 ppm)

266
Figure G.4: 13C NMR spectra of rubber seed oil (126.4–132.0 ppm)

Figure G.5: 13C NMR spectra of rubber seed oil (165–184 ppm)

267
Figure G.6: Double bond peaks in the 13C NMR spectrum for pure oleic acid, linoleic acid,

and linolenic acid. Peaks for each characteristic double bond carbon are assigned the

letters A for the peak found in all acids, B for the peak found in both linoleic and linolenic

acid, and C and C’ for the peaks found only in linolenic acid (Adapted from Sadowska et

al., 2008)

268
Appendix H

Table H.1: XRF spectroscopic elemental analysis of rubber seed shells catalyst (Onoji et

al., 2017b)

Elements Raw (uncalcined) RSS Calcined RSS at 800 oC for 3 h


pH=7.0 pH=12.5
Intensity Concentration wt.% Intensity Concentration wt.%
(mg/mL) (mg/mL)
Magnesium, Mg 0.0000 0.0267 0.5801 0.0001 0.1349 1.9423
Aluminum, Al 0.0003 0.0817 1.7749 0.0007 0.2048 2.9487
Silicon, Si 0.0015 0.0533 1.1579 0.0043 0.3392 4.8838
Phosphorus, P 0.0026 0.1202 2.6113 0.0037 0.1746 2.5132
Sulfur, S 0.0055 0.3847 8.3576 0.0059 0.4225 6.0832
Potassium, K 0.0020 0.1566 3.4021 0.0081 0.6516 9.3817
Calcium, Ca 0.0071 0.1195 2.5961 0.0243 0.7247 10.434
Vanadium, V 0.0002 0.0082 0.1781 0.0002 0.0093 0.1339
Chromium, Cr 0.0003 0.0111 0.2411 0.0002 0.0066 0.0950
Manganese, Mn 0.0003 0.0160 0.3476 0.0005 0.0284 0.4089
Cobalt, Co 0.0002 0.0017 0.0369 0.0001 0.0013 0.0187
Iron, Fe 0.0021 0.2766 6.0091 0.0076 0.8000 11.518
Nickel, Ni 0.0018 0.1079 2.3441 0.0015 0.0931 1.3405
Copper, Cu 0.0034 0.0630 1.3687 0.0037 0.1334 1.9207
Zinc, Zn 0.0050 0.1707 3.7085 0.0056 0.1947 2.8033
Tungsten, W 0.0006 0.1971 4.2820 0.0007 0.2528 3.6398
Rubidium, Rb 0.0000 0.0000 0.0000 0.0005 0.0018 0.0259
Niobium, Nb 0.0058 0.0675 1.4664 0.0054 0.0633 0.9114
Molybdenum 0.0035 0.1890 4.1060 0.0045 0.1592 2.2922
Tin, Sn 0.0074 1.3830 30.0456 0.0077 1.4414 20.753
Antimony, Sb 0.0093 1.1685 25.3856 0.0089 1.1078 15.950

269
Figure H.1: Nitrogen adsorption/desorption log plot for calcined RSS (Adapted from

Onoji et al., 2017b)

Figure H.2: Nitrogen adsorption/desorption log plot for raw RSS (Adapted from Onoji et

al., 2017b)
270
Figure H.3: BJH adsorption pore-size distribution log plot for calcined RSS (Adapted

from Onoji et al., 2017b)

Table H.2: Biodiesel yields for fresh and reused catalyst loadings at optimum conditions

Catalyst Yield with fresh Yield with reused catalyst (%)

loading (g) catalyst (%)

1st cycle 2nd cycle 3rd cycle 4th cycle 5th cycle

2.0 75.15 ± 1.25 72.4 ± 1.31 69.31 ± 1.12 67.52 ± 1.3 64.3 ± 1.18 60.4 ± 1.11

2.2 85.03 ± 1.13 84.8 ± 1.09 84.4 ± 1.11 83.5 ± 1.16 80.8 ± 0.04 77.3 ± 0.12

2.5 82.08 ± 1.27 82.01 ± 1.17 81.8 ± 1.32 81.0 ± 1.4 78.1 ± 1.39 74.2 ± 0.94

3.0 81.36 ± 1.18 79.5 ± 1.09 77.6 ± 1.31 75.3 ± 2.04 72.3 ± 1.99 70.4 ± 1.16

3.5 79.01 ± 2.01 76.1 ± 1.23 73.2 ± 1.35 70.4 ± 1.88 67.8 ± 2.04 64.2 ± 1.67

value ± standard deviation of duplicate data.

271
Appendix I

Section I.1

The combustion efficiency of diesel engines has resulted to increase in its usage in the

transportation sector in the last few decades. Thus, the environmental concerns are growing

and the fuels used in road transportation are subject to increasing stringent regulations

(Aldhaidhawi et al., 2017). Biodiesel sourced from non-edible rubber seed oil is considered as

a solution to the existing emission problems arising from the use of fossil diesel.

The experimental data for the analyses of the engine performance and emissions arising from

the use of the rubber seed oil biodiesel-diesel blends used in the study are presented in this

section. This includes air/fuel ratio, brake mean effective pressure, brake thermal efficiency,

brake specific fuel consumption, engine brake power, engine torque, exhaust gas temperature,

and emissions of CO, CO2, NOx, total hydrocarbons, and smoke opacity.

272
Table I.1: Variation of engine speed with fuel and air parameters for B00 (diesel)

Engine Fuel Fuel drain Fuel density Fuel Ambient Air Ambient air Air Air Air/Fuel BMEP
speed volume time (s) (ρ), (gcm-3) flowrate, air temp., differential pressure, PA velocity flowrate, ratio, (bar)
(rpm) (cm3) 𝑚̇𝑓 (kgh-1) TA (oC) pressure, ∆P (Pa) (ms-1) 𝑚̇𝑎 (kgh-1) 𝑚̇𝑎
(Pa) 𝑚̇𝑓
1500 8 94 0.85 0.26 28.6 -136 100,900 14.6 10.32 39.6 3.48

2000 8 79 0.85 0.31 29.0 -140 100,900 14.8 10.46 33.7 4.06

2500 8 72 0.85 0.34 29.3 -153 100,800 15.5 10.93 32.1 4.75

3000 8 44 0.85 0.56 30.1 -184 100,800 17 11.97 21.4 3.89

3500 8 29 0.85 0.83 31.1 -225 100,800 18.75 13.21 15.9 3.14

Table I.2: Variation of engine speed with fuel and air parameters for B10

Engine Fuel Fuel drain Fuel density Fuel Ambient Air Ambient air Air Air Air/Fuel BMEP
speed volume time (s) (ρ), (gcm-3) flowrate, air temp., differential pressure, PA velocity flowrate, ratio, (bar)
(rpm) (cm3) 𝑚̇𝑓 (kgh-1) TA (oC) pressure, ∆P (Pa) (ms-1) 𝑚̇𝑎 (kgh-1) 𝑚̇𝑎
(Pa) 𝑚̇𝑓
1500 8 91 0.853 0.27 30.1 -138 100,500 14.7 10.35 38.3 1.79

2000 8 79 0.853 0.31 30.2 -140 100,500 14.8 10.42 33.6 2.39

2500 8 66 0.853 0.37 30.5 -156 100,500 15.6 10.99 29.7 2.71

3000 8 38 0.853 0.64 30.7 -171 100,400 16.3 11.5 18 2.22

3500 8 27 0.853 0.90 30.8 -188 100,400 17.1 12.06 13.4 2.11

273
Table I.3: Variation of engine speed with fuel and air parameters for B20

Engine Fuel Fuel drain Fuel density Fuel Ambient Air Ambient air Air Air Air/Fuel BMEP
speed volume time (s) (ρ), (gcm-3) flowrate, air temp., differential pressure, PA velocity flowrate, ratio, (bar)
(rpm) (cm3) 𝑚̇𝑓 (kgh-1) TA (oC) pressure, ∆P (Pa) (ms-1) 𝑚̇𝑎 (kgh-1) 𝑚̇𝑎
(Pa) 𝑚̇𝑓
1500 8 72 0.855 0.34 31.7 -178 100,400 16.7 11.72 34.5 2.76

2000 8 77 0.855 0.32 31.6 -182 100,400 16.9 11.85 37 2.81

2500 8 77 0.855 0.32 31.7 -220 100,400 18.5 13.03 40.7 3.46

3000 8 70 0.855 0.35 31.6 -240 100,500 19.4 13.6 38.9 2.71

3500 8 66 0.855 0.37 31.6 -247 100,500 19.6 13.8 37.3 2.6

Table I.4: Variation of engine speed with fuel and air parameters for B30

Engine Fuel Fuel drain Fuel density Fuel Ambient Air Ambient air Air Air Air/Fuel BMEP
speed volume time (s) (ρ), (gcm-3) flowrate, air temp., differential pressure, PA velocity flowrate, ratio, (bar)
(rpm) (cm3) 𝑚̇𝑓 (kgh-1) TA (oC) pressure, ∆P (Pa) (ms-1) 𝑚̇𝑎 (kgh-1) 𝑚̇𝑎
(Pa) 𝑚̇𝑓
1500 8 67 0.858 0.36 30.5 -230 101,000 19 13.39 37.2 1.89

2000 8 71 0.858 0.35 31.0 -244 101,000 19.5 13.78 39.4 2.71

2500 8 71 0.858 0.35 29.5 -253 100,900 19.9 14.06 40.2 3.14

3000 8 55 0.858 0.45 29.5 -259 100,900 20.1 14.22 31.6 2.76

3500 8 51 0.858 0.48 30.9 -272 100,900 20.6 14.54 30.3 2.27

274
Table I.5: Variation of engine speed with fuel and air parameters for B50

Engine Fuel Fuel drain Fuel density Fuel Ambient Air Ambient air Air Air Air/Fuel BMEP
speed volume time (s) (ρ), (gcm-3) flowrate, air temp., differential pressure, PA velocity flowrate, ratio, (bar)
(rpm) (cm3) 𝑚̇𝑓 (kgh-1) TA (oC) pressure, ∆P (Pa) (ms-1) 𝑚̇𝑎 (kgh-1) 𝑚̇𝑎
(Pa) 𝑚̇𝑓
1500 8 67 0.863 0.37 31.6 -186 100,700 17 12 32.4 2.89

2000 8 71 0.863 0.35 31.6 -243 100,700 19.5 13.71 39.2 3.47

2500 8 65 0.863 0.38 31.5 -282 100,700 21 14.78 38.9 3.88

3000 8 51 0.863 0.49 32.2 -282 100,700 21 14.76 30.1 3.39

3500 8 47 0.863 0.53 32.4 -298 100,600 21.6 15.16 28.6 3.13

Table I.6: Variation of engine speed with fuel and air parameters for B100 (biodiesel)

Engine Fuel Fuel drain Fuel density Fuel Ambient Air Ambient air Air Air Air/Fuel BMEP
speed volume time (s) (ρ), (gcm-3) flowrate, air temp., differential pressure, PA velocity flowrate, ratio, (bar)
(rpm) (cm3) 𝑚̇𝑓 (kgh-1) TA (oC) pressure, ∆P (Pa) (ms-1) 𝑚̇𝑎 (kgh-1) 𝑚̇𝑎
(Pa) 𝑚̇𝑓
1500 8 87 0.876 0.29 31.1 -125 100,400 14 9.83 33.9 1.89

2000 8 76 0.876 0.33 31.1 -150 100,400 15.3 10.76 32.6 2.26

2500 8 62 0.876 0.41 31.5 -154 100,400 15.5 10.90 26.6 2.60

3000 8 33 0.876 0.77 31.6 -160 100,400 15.8 11.11 14.4 2.38

3500 8 21 0.876 1.19 31.8 -171 100,400 16.3 11.48 9.6 2.17

275
Table I.7: Variation of brake thermal efficiency (BTE %) with engine speed

Engine speed (rpm) BTE (%)


B00 B10 B20 B30 B50 B100
1500 33 15.8 28.4 17 26.7 18.1
2000 42 25.2 30 28.4 36.8 25.1
2500 56 35.9 41.1 34.6 42.8 28.2
3000 33 17.1 22.1 20.4 24 16.7
3500 21 13.5 18 14.1 24 11.5

Table I.8: Variation of brake specific fuel consumption (BSFC) with engine speed

Engine speed (rpm) BSFC ((kg/kWh)


B00 B10 B20 B30 B50 B100
1500 0.26 0.53 0.42 0.65 0.44 0.52
2000 0.20 0.34 0.29 0.33 0.26 0.38
2500 0.15 0.28 0.19 0.23 0.20 0.33
3000 0.25 0.50 0.22 0.28 0.25 0.56
3500 0.39 0.63 0.21 0.31 0.25 0.81

Table I.9: Variation of engine brake power (BP) with engine speed

Engine speed (rpm) BP (kW)


B00 B10 B20 B30 B50 B100
1500 1.01 0.518 0.803 0.55 0.84 0.55
2000 1.57 0.921 1.089 1.047 1.34 0.88
2500 2.30 1.31 1.675 1.518 1.88 1.257
3000 2.26 1.29 1.57 1.602 1.97 1.382
3500 2.13 1.43 1.759 1.539 2.12 1.466

Table I.10: Variation of engine torque with engine speed

Engine speed (rpm) Torque (Nm)


B00 B10 B20 B30 B50 B100
1500 6.4 3.3 5.1 3.5 5.4 3.5
2000 7.5 4.4 5.2 5.0 6.4 4.2
2500 8.8 5.0 6.4 5.8 7.8 4.8
3000 7.2 4.1 5.0 5.1 6.0 4.4
3500 5.8 3.9 4.8 4.2 5.4 4.0

276
Table I.11: Variation of exhaust gas temperature (EGT) with engine speed

Engine speed (rpm) EGT (oC)


B00 B10 B20 B30 B50 B100
1500 175 180 187 192 215 233
2000 185 190 200 215 238 259
2500 217 220 230 240 250 274
3000 240 242 250 260 270 286
3500 300 305 315 320 328 340

Table I.12: Variation of carbon monoxide emissions with engine speed

Engine speed (rpm) CO (ppm)


B00 B10 B20 B30 B50 B100
1500 165 160 157 154 150 145
2000 140 138 136 132 128 120
2500 125 123 120 117 114 110
3000 135 132 130 128 125 119
3500 138 136 134 133 130 124

Table I.13: Variation of carbon dioxide emissions with engine speed

Engine speed (rpm) CO2 (ppm)


B00 B10 B20 B30 B50 B100
1500 130 145 170 152 160 175
2000 135 150 180 165 170 190
2500 150 168 220 180 200 228
3000 170 200 250 225 230 265
3500 215 235 292 260 270 300

Table I.14: Variation of oxides of nitrogen emissions with engine speed

Engine speed (rpm) NOx (ppm)


B00 B10 B20 B30 B50 B100
1500 5 8 10 13 15 15
2000 12 15 17 20 23 25
2500 20 22 25 27 30 33
3000 23 26 30 35 38 42
3500 30 35 44 49 52 55

277
Table I.15: Variation of total hydrocarbons emissions with engine speed

Engine speed (rpm) THCs (ppm)


B00 B10 B20 B30 B50 B100
1500 120 115 110 100 90 80
2000 115 110 100 94 88 75
2500 99 89 85 82 79 70
3000 85 80 76 74 70 65
3500 100 98 90 88 83 73

Table I.16: Variation of smoke opacity emissions with engine speed

Engine speed (rpm) Smoke (BSU)


B00 B10 B20 B30 B50 B100
1500 1.41 1.3 1.01 1.1 1.06 0.89
2000 2.48 2.21 1.12 1.6 1.18 0.92
2500 3.69 3.2 1.98 2.6 2.2 1.89
3000 4.91 4.5 3 3.8 3.3 2.88
3500 6.01 5.6 3.88 5.1 4.3 3.6
BSU: Bosch smoke unit

278

You might also like