You are on page 1of 500

Biodiesel Technology

and Applications
Scrivener Publishing
100 Cummings Center, Suite 541J
Beverly, MA 01915-6106

Publishers at Scrivener
Martin Scrivener (martin@scrivenerpublishing.com)
Phillip Carmical (pcarmical@scrivenerpublishing.com)
Biodiesel Technology
and Applications

Edited by
Inamuddin, Mohd Imran Ahamed,
Rajender Boddula
and Mashallah Rezakazemi
This edition first published 2021 by John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
and Scrivener Publishing LLC, 100 Cummings Center, Suite 541J, Beverly, MA 01915, USA
© 2021 Scrivener Publishing LLC
For more information about Scrivener publications please visit www.scrivenerpublishing.com.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or other-
wise, except as permitted by law. Advice on how to obtain permission to reuse material from this title
is available at http://www.wiley.com/go/permissions.

Wiley Global Headquarters


111 River Street, Hoboken, NJ 07030, USA

For details of our global editorial offices, customer services, and more information about Wiley prod-
ucts visit us at www.wiley.com.

Limit of Liability/Disclaimer of Warranty


While the publisher and authors have used their best efforts in preparing this work, they make no rep­
resentations or warranties with respect to the accuracy or completeness of the contents of this work and
specifically disclaim all warranties, including without limitation any implied warranties of merchant-­
ability or fitness for a particular purpose. No warranty may be created or extended by sales representa­
tives, written sales materials, or promotional statements for this work. The fact that an organization,
website, or product is referred to in this work as a citation and/or potential source of further informa­
tion does not mean that the publisher and authors endorse the information or services the organiza­
tion, website, or product may provide or recommendations it may make. This work is sold with the
understanding that the publisher is not engaged in rendering professional services. The advice and
strategies contained herein may not be suitable for your situation. You should consult with a specialist
where appropriate. Neither the publisher nor authors shall be liable for any loss of profit or any other
commercial damages, including but not limited to special, incidental, consequential, or other damages.
Further, readers should be aware that websites listed in this work may have changed or disappeared
between when this work was written and when it is read.

Library of Congress Cataloging-in-Publication Data

ISBN 9781119724643

Cover image: Pixabay.com


Cover design by Russell Richardson

Set in size of 11pt and Minion Pro by Manila Typesetting Company, Makati, Philippines

Printed in the USA

10 9 8 7 6 5 4 3 2 1
Contents

Preface xvii
1 Biocatalytic Processes for Biodiesel Production 1
Ubaid Mehmood, Faizan Muneer, Muhammad Riaz,
Saba Sarfraz and Habibullah Nadeem
1.1 Introduction and Background 2
1.2 Importance of Biodiesel Over Conventional Diesel Fuel 3
1.3 Substrates for Biodiesel Production 4
1.4 Methods in Biodiesel Production 6
1.5 Types of Catalysts Involved in Biodiesel Production 7
1.5.1 Chemical Homogenous Catalysts 7
1.5.2 Solid Heterogeneous Catalysts 8
1.5.3 Biocatalysts 8
1.6 Factors Affecting Enzymatic Transesterification Reaction 8
1.6.1 Effect of Water in Enzyme Catalyzed
Transesterification 9
1.6.2 Effect of Bioreactor 10
1.6.3 Effect of Acyl Acceptor on Enzymatic Production
of Biodiesel 10
1.6.4 Effect of Temperature on Enzymatic Biodiesel
Production 14
1.6.5 Effect of Glycerol on Enzymatic Biodiesel Production 14
1.6.6 Effect of Solvent on Biodiesel Production 16
1.7 Lipases as Biocatalysts for Biodiesel Production 17
1.7.1 Mechanisms of Lipase Action 19
1.7.2 Efficient Lipase Sources for Biodiesel Producing
Biocatalyst 19
1.8 Comparative Analysis of Intracellular and Extracellular
Lipases for Biodiesel Production 21
1.9 Recombinant Lipases for Cost-Effective Biodiesel Production 26
1.10 Immobilization of Lipases for Better Biodiesel Production 28

v
vi Contents

1.11 Recent Strategies to Improve Biodiesel Production 31


1.11.1 Combination of Lipases 31
1.11.2 Microwave and Ultrasonic-Assisted Reaction 33
1.12 Lipase Catalyzed Reaction Modeling and Statistical
Approaches for Reaction Optimization 35
1.13 Conclusion and Summary 38
References 38
2 Application of Low-Frequency Ultrasound for Intensified
Biodiesel Production Process 59
Mohd Razealy Anuar, Mohamed Hussein Abdurahman,
Nor Irwin Basir and Ahmad Zuhairi Abdullah
2.1 Current Fossil Fuel Scenario 60
2.2 Biodiesel 60
2.3 Transesterification 61
2.4 Challenges for Improved Biodiesel Production 62
2.5 Homogeneous Catalyst for Biodiesel Production 63
2.6 Heterogeneous Catalyst for Biodiesel Production 64
2.7 Immiscibility of the Reactants 65
2.8 Ultrasound-Assisted Biodiesel Production Process 66
2.8.1 Fundamental Aspects of the Process 66
2.8.2 Homogeneously Catalyzed Ultrasound-Assisted
System 69
2.8.3 Heterogeneously Catalyzed Ultrasound-Assisted
System 72
2.8.3.1 Heterogeneously Acid Catalyzed System 72
2.8.3.2 Heterogeneous Based Catalyzed
Ultrasound-Assisted System 74
2.8.3.3 Influence of Reaction Parameters 78
2.9 Conclusions 79
Acknowledgement 80
References 80
3 Application of Catalysts in Biodiesel Production 85
Anilkumar R. Gupta and Virendra K. Rathod
3.1 Introduction 85
3.2 Homogeneous Catalysis for the Biodiesel Production 89
3.2.1 Homogeneous Acid Catalyst 89
3.2.2 Homogeneous-Base Catalyst 93
3.3 Heterogeneous Catalyst 96
3.3.1 Heterogeneous Acid Catalyst 97
3.3.2 Heterogeneous-Base Catalyst 106
Contents vii

3.4 Biocatalysts 115


3.5 Conclusion 119
References 124
4 Hydrogenolysis as a Means of Valorization
of Biodiesel-Derived Glycerol: A Review 137
Manjoro T.T., Adeniyi A. and Mbaya R.K.K.
4.1 Introduction 138
4.2 Ways of Valorization of Biodiesel-Derived Glycerol 139
4.2.1 Catalytic Conversion of Glycerol Into Value-Added
Commodities 140
4.2.1.1 Catalytic Oxidation of Glycerol 140
4.2.1.2 Catalytic Dehydration of Glycerol 143
4.2.1.3 Pyrolysis of Bioglycerol 144
4.2.1.4 Glycerol Transesterification 145
4.2.1.5 Glycerol Direct Carboxylation 146
4.3 Hydrogenolysis of Glycerol 147
4.3.1 Definition of Hydrogenolysis 147
4.3.2 Catalytic Hydrogenolysis of Glycerol 148
4.3.3 Product Spectrum from Hydrogenolysis of Glycerol 148
4.3.4 Hydrogenolysis of Glycerol to 1,2-PDO (Propylene
Glycol): Reaction Systems Overview 149
4.3.5 Catalyst Selection 151
4.3.6 Reaction Conditions That Influence the
Hydrogenolysis of Glycerol to 1,2-PDO 153
4.3.6.1 Effect of Reaction Temperature 153
4.3.6.2 Effect of H2 Pressure 154
4.3.6.3 Effect of Initial Water Concentration 155
4.3.6.4 Effect of Reaction Time 156
4.3.6.5 Effect of Catalyst Weight 156
4.3.6.6 Proposed Reaction Mechanisms for Glycerol
Hydrogenolysis to Produce 1,2-PDO 157
4.4 Conclusion 159
References 159
5 Current Status, Synthesis, and Characterization of Biodiesel 167
Akshay Garg, Gaurav Dwivedi, Prashant Baredar
and Siddharth Jain
5.1 Introduction 167
5.2 Status of Biodiesel in India 169
5.3 Biodiesel Production in India 169
viii Contents

5.3.1 Feedstocks Popular in India 169


5.3.1.1 Jatropha (Jatropha curcas) Oil 171
5.3.1.2 Pongamia Oil 171
5.3.1.3 Mahua Oil 171
5.3.1.4 Neem Oil 171
5.3.1.5 Linseed Oil 171
5.3.1.6 Rubber Seed Oil 172
5.3.1.7 Tobacco Oil 172
5.3.1.8 Castor 172
5.3.1.9 Waste Cooking Oil 172
5.3.1.10 Algae Oil 172
5.3.2 Advantages of Non-Edible Oils 173
5.3.3 Modification Techniques 173
5.3.3.1 Blending 173
5.3.3.2 Micro-Emulsification 173
5.3.3.3 Cracking 174
5.3.3.4 Transesterification 174
5.3.4 Biodiesel Production Methodology 174
5.3.4.1 Catalytic Transesterification 174
5.3.4.2 Non-Catalytic Transesterification 178
5.3.5 Optimization Methodology for Biodiesel 179
5.3.5.1 Central Composite Design Technique 179
5.3.5.2 Box Behnken Technique 179
5.4 Properties of Biodiesel 180
5.5 Analytical Methods 181
5.5.1 Titration 181
5.5.2 Chromatic Methods 181
5.5.2.1 Gas Chromatography 183
5.5.2.2 High-Performance Liquid Chromatography 184
5.5.3 Spectroscopic Methods 184
5.5.3.1 Nuclear Magnetic Resonance Spectroscopy 184
5.5.3.2 Infrared Spectroscopy 185
5.5.4 Rancimat Method 185
5.5.5 Viscometry 186
5.6 Conclusion 186
References 187
6 Commercial Technologies for Biodiesel Production 195
Chikati Roick, Leonard Okonye, Nkazi Diankanua
and Gorimbo Joshua
Abbreviation 196
Contents ix

6.1 Introduction 196


6.2 Biodiesel Production 197
6.3 Technologies Used for Biodiesel Production 198
6.3.1 Chemical Reaction (Transesterification) 199
6.3.2 Thermochemical Conversion 199
6.3.3 Biomechanical Conversion 201
6.3.4 Direct Combustion 201
6.4 Other Technologies in Use for Biodiesel Production 201
6.5 Feedstock Requirement 203
6.6 Some Problems Facing Commercialization of Biodiesel
in Africa 203
6.7 Case Studies/Current Status and Future Potential 204
6.8 Conclusions 207
Acknowledgments 208
References 208
7 A Global Scenario of Sustainable Technologies and Progress
in a Biodiesel Production 215
M. B. Kumbhar, P. E. Lokhande,, U. S. Chavan
and V.G. Salunkhe
7.1 Introduction 216
7.2 Current Status of Feedstock for Biodiesel Production
Technology 218
7.3 Scenario of Biodiesel in Combustion Engine 222
7.4 Biodiesel Production Technologies 223
7.4.1 Direct Blending 223
7.4.2 Pyrolysis 224
7.4.3 Microemulsification 225
7.4.4 Transesterification 226
7.5 Microwave-Mediated Transesterification 227
7.6 Ultrasound-Mediated Transesterification 229
7.7 Catalysis in Biodiesel Production 230
7.7.1 Homogeneous Catalysts 230
7.7.2 Heterogeneous Catalysts 231
7.7.3 Heterogeneous Nanocatalysts 232
7.7.4 Supercritical Fluids 232
7.7.5 Biocatalysts 232
7.8 The Concept of Biorefinery 234
7.9 Summary and Outlook 236
7.10 Conclusion 237
References 237
x Contents

8 Biodiesel Production Technologies 241


Moina Athar and Sadaf Zaidi
8.1 Introduction 242
8.2 Biodiesel Feedstocks 242
8.2.1 Selection of Feedstocks 243
8.3 Biodiesel Production Technologies 248
8.3.1 Pyrolysis 248
8.3.2 Dilution 249
8.3.3 Micro-Emulsion 249
8.3.4 Transesterification 249
8.3.4.1 Homogeneously Catalyzed
Transesterification Processes 250
8.3.4.2 Heterogeneously Catalyzed
Transesterification Processes 252
8.3.4.3 Enzymatic Catalyzed Transesterification
Processes 252
8.4 Intensification Techniques for Biodiesel Production 253
8.4.1 Supercritical Alcohol Method 253
8.4.2 Microwave Heating 253
8.4.3 Ultrasonic Irradiation 255
8.4.4 Co-Solvent Method 256
8.5 Other Techniques of Biodiesel Production 256
References 257
9 Methods for Biodiesel Production 267
M.Gul, M.A. Mujtaba, H.H. Masjuki, M.A. Kalam
and N.W.M. Zulkifli
9.1 Selection of Feedstock for Biodiesel 267
9.1.1 First-Generation Feedstock 268
9.1.2 Second-Generation Feedstock 268
9.1.3 Third-Generation Feedstock 269
9.2 Methods for Biodiesel Production 269
9.2.1 Dilution With Hydrocarbons Blending 269
9.2.2 Micro-Emulsion 269
9.2.3 Pyrolysis (Thermal Cracking) 270
9.2.4 Transesterification (Alcoholysis) 271
9.2.4.1 In Situ Transesterification
(Reactive Extraction) 271
9.2.4.2 Conventional Transesterification 272
9.2.4.3 Microwave/Ultrasound-Assisted
Transesterification 278
Contents xi

9.2.4.4 Variables Affecting Transesterification


Reaction 278
References 282
10 Non-Edible Feedstock for Biodiesel Production 285
Chikati Roick, Kabir Opeyemi Otun, Nkazi Diankanua
and Gorimbo Joshua
List of Abbreviations 286
10.1 Introduction 286
10.2 Reports Relevant to Global Warming
and Renewable Energy 287
10.3 Biofuels as an Alternative Energy Source 288
10.3.1 First-Generation Biofuels 288
10.3.2 Second-Generation Biofuels 289
10.3.3 Third-Generation Biofuels 290
10.4 Benefits of Using Biodiesel 290
10.5 Technologies of Biodiesel Production From Non-Edible
Feedstock 291
10.6 Biodiesel Production by Transesterification 292
10.7 Non-Edible Feedstocks for Biodiesel Production 295
10.7.1 Non-Edible Vegetable Oils 296
10.7.2 Waste Cooking Oil 297
10.7.3 Algal Oil 298
10.7.4 Waste Animal Fat/Oil 299
10.8 Fuel Properties of Biodiesel Obtained From Non-Edible
Feedstock 299
10.9 Advantages of Non-Edible Feedstocks 302
10.10 Economic Importance of Biodiesel Production 302
10.11 Conclusions 303
Acknowledgments 303
References 304
11 Oleochemical Resources for Biodiesel Production 311
Gayathri R., Ranjitha J. and Vijayalakshmi Shankar
11.1 Introduction 311
11.2 Definition of Oleochemicals 312
11.3 Oleochemical Types 313
11.4 Production of Biodiesel 315
11.5 Types of Feedstocks 317
11.5.1 Non-Edible Feedstocks 317
11.5.2 Non-Edible Vegetable Oil 317
xii Contents

11.5.3 Tall Oil 318


11.5.4 Waste Cooking Oils 318
11.5.5 Animal Fats 318
11.5.6 Chicken Fat 319
11.5.7 Lard 319
11.5.8 Tallow 320
11.5.9 Leather Industry Solid Waste Fat 321
11.5.10 Fish Oil 322
11.6 Uses of Oleochemicals 322
11.6.1 Polymer Applications 322
11.6.2 Application of Plant Oil as a Substitute
for Petro-Diesel 323
11.6.3 Used as Surfactants 323
11.6.4 Oleochemicals Used in Pesticide 324
11.6.5 Oleochemicals Used in Spray Adjuvants
and Solvents 324
11.7 Methyl Ester or Biodiesel Production 324
11.7.1 Palm Oil 326
11.7.2 Sunflower Oil 326
11.7.3 ME From AFW 327
11.8 Parameters Affecting the Yield of Biodiesel 327
11.8.1 Reaction Conditions 327
11.8.2 Catalyst 327
11.8.2.1 Alkali Catalyst 327
11.8.2.2 Acid Catalyst 329
11.8.2.3 Biocatalyst 329
11.8.2.4 Heterogeneous Catalyst 329
11.8.2.5 ME Conversion by Supercritical
Method 329
11.8.3 Properties of Feedstock 330
11.8.3.1 Composition of FA 330
11.8.3.2 FFA 330
11.8.3.3 Heat 330
11.8.3.4 Presence of Unwanted Materials 330
11.8.3.5 Titer 332
11.8.4 Characteristic of Feedstock 332
11.9 Optimization of Reactions Conditions for High Yield
and Quality of Biodiesel 332
11.9.1 Pre-Treatment of Feedstock 332
11.9.1.1 Elimination of Water 332
11.9.1.2 Elimination of Insoluble Impurities 332
Contents xiii

11.9.1.3 Elimination of Unsaponifiables 333


11.9.2 Characterization and Selection of Feedstocks 333
11.9.3 Selection of Reaction Conditions 333
11.10 Oil Recovery 333
11.10.1 Alkaline Flooding Method 333
11.10.2 Additives 334
11.11 Quality Improvement of Biodiesel 334
11.11.1 Additives for Improving Combustion Ability 334
11.11.2 Additives for Enhancing the Octane Number 334
11.11.3 Additives for Improving the Stability 334
11.11.4 Additives to Enhance Cold Flow Property 334
11.11.5 Additives to Enhance Lubricity 335
11.11.6 Additives to Enhance Cetane Number 335
11.12 Conclusion 335
Abbreviations 335
References 336
12 Overview on Different Reactors for Biodiesel Production 341
V. C. Akubude, K.F. Jaiyeoba, T.F Oyewusi, E.C. Abbah,
J.A. Oyedokun and V.C. Okafor
12.1 Introduction 341
12.2 Biodiesel Production Reactors 342
12.2.1 Batch Reactor 343
12.2.2 Continuous Stirred Tank Reactor 344
12.2.3 Fixed Bed Reactor 346
12.2.4 Bubble Column Reactor 347
12.2.5 Reactive Distillation Column 349
12.2.6 Hybrid Catalytic Plasma Reactor 350
12.2.7 Microreactors Technology 350
12.2.8 Oscillatory Flow Reactors 353
12.2.9 Other Novel Reactors 353
12.3 Future Prospects 354
12.4 Conclusion 354
References 354
13 Patents on Biodiesel 361
Azira Abdul Razak, Mohamad Azuwa Mohamed
and Darfizzi Derawi
13.1 Introduction 361
13.2 Generation of Biodiesel 362
13.3 Development of Catalyst 363
xiv Contents

13.3.1 Homogeneous Catalyst 364


13.3.2 Heterogeneous Catalyst 364
13.4 Method Producing Biodiesel 365
13.4.1 Pre-Treatment Process 365
13.4.2 Direct Use and Blending of Oils 366
13.4.3 Esterification of FFA 366
13.4.4 Transesterification of TAG 367
13.4.5 Pyrolysis 368
13.5 Reactor’s Technology for Biodiesel Production 369
13.5.1 Continuous Stirred Tank Reactor 370
13.5.2 Fixed Bed Reactor 370
13.5.3 Micro-Mixer Reactor 371
13.6 Conclusion 372
References 372
14 Reactions of Carboxylic Acids With an Alcohol Over
Acid Materials 377
J.E. Castanheiro
14.1 Introduction 377
14.2 Zeolites 378
14.3 SO3H as Catalyst 379
14.4 Metal Oxides 380
14.5 Heteropolyacids 382
14.6 Other Materials 384
14.7 Conclusions 384
References 385
15 Biodiesel Production From Non-Edible and Waste
Lipid Sources 389
Opeoluwa O. Fasanya, Aishat A. Osigbesan
and Onoriode P. Avbenake
15.1 Introduction 390
15.2 Non-Edible Plant-Based Oils 394
15.2.1 Jatropha curcas 394
15.2.2 Calophyllum inophyllum 397
15.2.3 Mesua ferrea 397
15.2.4 Jojoba Oil 398
15.2.5 Azadirachta indica 398
15.2.6 Rubber Seed Oil 399
15.2.7 Ricinus communis as Feedstock (Castor Oil) 402
15.2.8 Other Non-Edible Oils 403
Contents xv

15.3 Waste Animal Fats 404


15.4 Expired and Waste Cooking Oils 405
15.5 Algae/Microalgae 406
15.6 Insects as Biodiesel Feedstock 411
15.7 Deacidification 414
15.8 Other Technologies 414
15.9 Conclusion 415
References 415
16 Microalgae for Biodiesel Production 429
Charles Oluwaseun Adetunji, Victoria Olaide Adenigba,
Devarajan Thangadura and Mohd Imran Ahamed
16.1 Introduction 430
16.2 Physicochemical Properties of Biodiesel From Microalgae 431
16.3 Genetic Engineering/Techniques Enhancing Biodiesel
Production 432
16.4 Nanotechnology in Microalgae Biodiesel Production 434
16.5 Specific Examples of Biodiesel Production From
Microalgae 434
16.6 Methodology Involved in the Extraction of Algae 438
16.6.1 Chemical Solvents Extraction 439
16.6.2 Extraction by Supercritical Carbon Dioxide 439
16.6.3 Extraction Using Biochemical Techniques 439
16.6.4 Extraction Involving Direct Transesterification 440
16.6.5 Extraction Using Transesterification Techniques 440
16.7 Conclusion and Future Recommendation to Knowledge 440
References 441
17 Biodiesel Production Methods and Feedstocks 447
Setareh Heidari and David A. Wood
17.1 Introduction 448
17.2 Biofuel Classification in Terms of Origin and
Technological Conversion of Raw Materials 449
17.3 Techniques Capable of Producing Biodiesel
on Commercial Scales 451
17.3.1 Direct and Blending Methods With the Aim
of Biodiesel Generation 452
17.3.2 Microemulsion Methods 452
17.3.3 Pyrolysis Methods 453
17.3.4 Transesterification Methods 453
17.4 Influential Parameters on Biodiesel Production 454
xvi Contents

17.4.1 The Choice of Transesterification Catalysts 454


17.4.2 Effects of Catalyst Characteristics on Biodiesel
Production Efficiency 454
17.5 Biodiesel Markets and Economic Considerations 455
17.6 Challenges Confronting Biodiesel Uptake 456
17.7 Corrosion and Quality Monitoring Issues for Biodiesel 457
17.8 Conclusions 457
References 458
18 Application of Nanoparticles for the Enhanced Production
of Biodiesel 465
Muhammad Hilman Mustapha, Akhsan Kamil Azizi,
Wan Nur Aini Wan Mokhtar and Mohamad Azuwa Mohamed
18.1 Introduction 465
18.2 Solid Nanoparticles 466
18.3 Nanobioparticles/Nanobiocatalyst 471
18.4 Magnetic Nanoparticles 473
18.5 How Nanoparticles Enhanced Biodiesel Production? 475
18.6 Conclusion 477
References 477
Index 481
Preface

Energy technologies have attracted great attention due to the fast develop-
ment of sustainable energy. Biodiesel technologies have been identified as
the sustainable route through which overdependence on fossil fuels can be
reduced. Biodiesel has played a key role in handling the growing challenge
of a global climate change policy. Biodiesel is defined as the monoalkyl
esters of vegetable oils or animal fats. Biodiesel is a cost-effective, renew-
able, and sustainable fuel that can be made from vegetable oils and ani-
mal fats. Compared to petroleum-based diesel, biodiesel would offer a
non-toxicity, biodegradability, improved air quality and positive impact
on the environment, energy security, safe-to-handle, store and transport,
and so on. Biodiesels have been used as a replacement of petroleum diesel
in transport vehicles, heavy-duty trucks, locomotives, heat oils, hydrogen
production, electricity generators, agriculture, mining, construction, and
forestry equipment.
This book describes a comprehensive overview, covering a broad range
of topics on biodiesel technologies and allied applications. Chapters cover
history, properties, resources, fabrication methods, parameters, formula-
tions, reactors, catalysis, transformations, analysis, in situ spectroscopies,
key issues and applications of biodiesel technology. It also includes bio-
diesel methods, extraction strategies, biowaste utilization, oleochemical
resources, non-edible feedstocks, heterogeneous catalysts, patents, and
case-studies. Progress, challenges, future directions, and state-of-the-art
biodiesel commercial technologies are discussed in detail. This book is
an invaluable resource guide for professionals, faculty, students, chemical
engineers, biotechnologists, and environmentalists in these research and
development areas. This book includes the eighteen chapters and the sum-
maries are given as follows.
Chapter 1 details the biocatalytic production of biodiesel. Microbial
enzymes such as lipases act as biocatalysts in the transesterification pro-
cess of biodiesel production. Suitable and cost-effective feedstocks or

xvii
xviii Preface

substrates for biodiesel production including their percentage yields are


discussed. Factors that affect the enzymatic transesterification reaction are
also explained.
Chapter 2 addresses ultrasonic energy which can increase the interface
area while creating a thermal effect in heterogeneous biodiesel production
process to result in higher biodiesel yield. Fundamental understanding of
the improved reactant-catalyst interaction, the nature of the thermal effect,
favorable process behaviors, reaction kinetic, as well as the effect on bio-
diesel quality is particularly addressed.
Chapter 3 is about the study of different types of catalysts used for
biodiesel production. The classification of catalysts, advantages, and lim-
itations, along with their mechanism, is explained. The heterogeneous
catalysts’ synthetic methods and immobilization of biocatalyst are also dis-
cussed in detail.
Chapter 4 discusses various methods used to produce value-added
chemicals from biodiesel-derived glycerol. The main focus being is given
to hydrogenolysis as a transformative process to selectively produce 1,2-
propanediol and the advancements in biodiesel technologies. Furthermore,
knowledge gaps are highlighted based on extensive literature research on
the subject.
Chapter 5 discusses various techniques of synthesizing biodiesel and
review of various existing analytical technologies for characterization of
biodiesel. The chapter focuses on the current status of biodiesel in India,
i.e., using non-edible sources and future feasibility of developing new
methods of characterization to reduce the cost of biodiesel production.
Chapter 6 examines various established technologies available for the
production of biodiesel, viz., chemical reaction, direct combustion, ther-
mochemical conversion, and biomechanical conversion. Each technology
is apportioned to a certain type of feedstock. Case studies, current sta-
tus, and future potential of commercialization of biodiesel production in
Africa are also discussed.
There is a huge demand for sustainable biofuel production in coming
decades. The key challenges for biodiesel production are high FFA with the
desired level of yield, stability, optimized and flexible production, commer-
cialization of feedstock and environmentally friendly cycle. The collective
effort and commitment of research survey regard feedstocks and commer-
cialization of technology around the globe towards sustainable energy are
expressed in terms of accelerating the biofuel economy in Chapter 7.
Chapter 8 provides an overview of the available feedstocks, pro-
duction methods, and the benefits and constraints of using homoge-
neous, heterogeneous, and enzymatic catalysts for biodiesel. Some latest
Preface xix

intensification techniques to manage mass transfer restrictions of oil and


alcohol phases along with some production cost reduction measures are
also highlighted.
Chapter 9 discusses different types of feedstocks used for synthesizing
biodiesel and feedstock selection criteria. Moreover, all biodiesel produc-
tion methods (i.e., dilution with hydrocarbons blending, micro-emulsion,
pyrolysis, and transesterification) are also described in detail with their
advantages and disadvantages. The major focus is given to the various
transesterification methods. Production methods also include experimen-
tal setup layouts, all process parameters, reaction conditions, the latest
advancement in reaction processes, and their effects on biodiesel yield.
Chapter 10 reviews the potential use of non-edible feedstocks in the
production of biodiesel. Special attention is given to the types of feedstocks
available and their production pathways to biodiesel. The state-of-the-art
technology, the properties of the fuel produced, and the environmental
concerns of biofuels are also discussed.
Chapter 11 discusses the various types of oleochemicals and their usage.
Optimization and production of biodiesel derived from oleochemicals
and their properties are also discussed. The primary focus is given for the
advantage of oleochemicals to be used as a potential feedstock for biodiesel
production from the available literature.
Chapter 12 provides details about the different configurations of reac-
tors used in biodiesel production. There are two types, namely, batch and
continuous reactors. Recently, other improved configurations like micro-
reactors have emerged. This chapter also discusses the merits and demerits
of these reactors.
Chapter 13 highlights and discusses the international patents on bio-
diesel applications. This chapter reviews the recent patents on the generation
of biodiesel which depends on the feedstock used, catalysts development,
the latest method for biodiesel production, and reactor technology for the
biodiesel production.
Chapter 14 overviews different reactions between a carboxylic acid
(fatty acids) and alcohol (methanol and ethanol) over heterogeneous cata-
lysts, an important step in biodiesel production. The nature of solid materi-
als, like zeolites, heteropolyacids, materials with sulfonic groups, inorganic
mixed oxides, and clays towards biodiesel production is discussed.
Chapter 15 sheds light on inedible feedstock that could be utilized for
biodiesel production. Plant-based and non-plant feedstock are discussed.
The waste lipid sources which are unfit for consumption are also high-
lighted. The chemical composition, economic viability, and sustainability
of some of these feedstocks are equally explored.
xx Preface

Chapter 16 provides detailed information on the fabrication of biodiesel


from microalgae. Specific information on the physical properties, amount
of biodiesel production, and level of transesterification of biodiesel are dis-
cussed. The application of photobioreactors for the production of biodiesel
with the special consideration of several factors such as flow rate, tempera-
ture, light intensity, CO2 concentration, and time is highlighted. Several
techniques for the extraction of biodiesel such as supercritical CO2, physi-
cochemical, direct transesterification, chemical solvents, and biochemical
respectively are highlighted.
Chapter 17 discusses the biofuel classification in terms of origin and
technological conversion of raw materials. Techniques capable of produc-
ing biodiesel on commercial scales are also presented. Furthermore, influ-
ential parameters and their roles in biodiesel production are elaborately
covered. Finally, challenges and limitations confronting biodiesel uptake
are presented.
Chapter 18 mainly explicates the application of nanoparticle catalysis
for the high production of biodiesel. In particular, various types of catalyst
nanoparticles with different synthesis strategy and their roles in enhancing
the biodiesel production are discussed.

Inamuddin, Mohd Imran Ahamed, Rajender Boddula


and Mashallah Rezakazemi
1
Biocatalytic Processes for
Biodiesel Production
Ubaid Mehmood1, Faizan Muneer2, Muhammad Riaz3, Saba Sarfraz4
and Habibullah Nadeem2*

College of Chemistry, Chemical Engineering and Biotechnology,


1

Donghua University, China


2
Department of Bioinformatics and Biotechnology, Government College University
Faisalabad, Pakistan
3
Department of Food Sciences, University College of Agriculture, Bahauddin
Zakariya University, Multan, Pakistan
4
Department of Chemistry, Government College Women University Faisalabad,
Faisalabad, Pakistan

Abstract
Enzymes such as microbial lipases can be effectively used as biocatalysts for bio-
diesel production in a sustainable manner. Biocatalytic processes to produce bio-
diesel or biofuel is the need of time to reduce the emission of greenhouse gases
produced from conventional diesel or fossil fuels. Lipases with excellent biochem-
ical and physiological properties are most commonly used to catalyze the trans-
esterification process for biodiesel production. Lipases obtained from microbes
such as bacteria and fungi produce 70%–95% ethanol and methanol. Biodiesel is
usually composed of fatty acid alkyl esters which are mono-alkyl esters of either
fatty acid methyl esters or fatty acid ethyl esters depending upon the alcohol (acyl
acceptor) being used in the reaction. Factors such as bioreactor type, acyl accep-
tor, temperature, and glycerol can affect the enzymatic transesterification reaction.
Recombinant enzymes such as recombinant lipases can be employed to obtain
higher percentage of biodiesel due to their high specificity and biocatalytic activity
for different substrates used for biodiesel production.
Keywords: Lipases, biodiesel, biocatalysis, biofuels, Novozyme, free fatty acids,
ethyl acceptors

*Corresponding author: habibullah@gcuf.edu.pk

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (1–58) © 2021 Scrivener Publishing LLC

1
2 Biodiesel Technology and Applications

1.1 Introduction and Background


Biofuels are crucial for the conservation of our natural environment and
the climate. Biofuel such as bioethanol can be used for energy generation
purposes which are currently being produced by fossil fuels such as petrol,
diesel, and kerosene oil [1]. Being non-renewable energy sources, fossil fuels
will not only deplete from the planet earth but will also leave a long-term
impact on the globe both in terms of economy and climate change. Apart
from being limited natural fuel reserves, there are countless reasons available
that justify the need of natural and eco-friendly energy sources such as bio-
fuels. Transportation, power generation, and house hold appliances use fuels
directly or indirectly and for that purpose we are almost dependent on fossil
fuels [2]. If efficient and robust methods and technologies are not worked
out, we might come to a permanent stand still condition in the future when
all our natural fossil fuel reserves will be vanished. The use of fossil fuel pro-
duces gases such as carbon dioxide (CO2), carbon monoxide (CO), sulfur
oxides (SOx), and nitrogen oxides (NOx) which are unhealthy for human
beings causing health issues such as asthma, skin diseases, and even cancers
[3]. These by-products of fuel consumption affect not only human but also
animals and plants on a broader view. Plant production and growth rates are
highly effected by the changing environmental and climatic conditions due
to heavy use of fossil fuels and their derivatives such as plastics [4].
Vehicular CO2 emission in the past decade was 20%, and it is estimated
that by 2030, it will reach up to 80%. Liquide biofuels got prominence with
the automobile industry. Peanut oil was used to make biofuel, i.e., biodiesel by
Rudolph Diesel in 1898. Henri Ford who was the founder of Ford Company
an automobile industry was also convinced by the idea of using biofuels in
his automobile. During World War-II, Germany used biomass-based fuels for
their machines which is the evidence of its use back in 1940s. The utilization
of biofuels was presented, but after two major oil crises, first was in 1973 and
second in 1978, and brought back its importance to public again. Biofuels that
are produced using a large number of biomass sources are a sustainable solu-
tion for the environment and biosphere conservation. Being renewable energy
resources and eco-friendly to the environment and life on earth, these are
highly desirable products produced from renewable biomass substrates [5].
Currently, biofuels from various agricultural sources such as soybean oil,
rapeseed oil, recycled waste oils, and waste plant residues are being studied.
Depending on the feedstock type, processing technology and their devel-
opmental level, biofuels can be classified into first-, second-, and third-­
generation biofuels. Biofuels produced directly from edible feedstock such
Biocatalytic Processes 3

as crops, sugars, and edible oil using conventional techniques are considered
as first-generation biofuels [6]. Non-edible feedstock such as waste crop res-
idues like lignocelluloses and waste vegetable oils are required to produce
second-generation biofuels which are comparatively economical and more
sustainable as there is no food versus fuel competition. Highly advanced
methods are used to produce second-generation biofuels which has certainly
less flaws and ultimately improved to get greater yield [7]. We are currently in
the phase of second-generation biofuels. Most of the processing techniques
for second-generation biofuel production are not available at commercial
level. One must think that the land dedicated for edible feedstock/crops will
be compromised if we start cultivating non-edible crops in that land. Marginal
lands can be used for the cultivation of grasses and other plants that are not
a food for human or nor a fodder for animals on a larger scale. These plants
or marginal grasses can be used for the production of second-generation bio-
fuels. There have been a lot of research investigations to produce biodiesel
using non-edible plant oils such as keranja oil, Jatropha curcas oil, tobacco oil,
Calophyllum inophyllum oil, and castor oil [8]. Jatropha is an effective source
of biodiesel production because of 30%–50% oil contents in its seeds [9]. The
actual precursors of most of the second-generation biodiesel production are
waste oils either in the form of waste cooking or industrial oils or animal fats.
The utilization of these waste materials as feed stock helps in managing and
disposing of waste material, which is one of the biggest problem for earth, for
the benefit of environment [10]. In order to comprehend different biofuels,
we can categorize them into four types which include biodiesel, bioalcohol
(biomethanol, bioethanol, biobutanol), biogas, and biohydrogen. The most
widely used biofuels are liquid biofuels such as biodiesel and bioethanol.
Biofuels can be blended with other petro-based fuels in order to manage and
enhance quality and quantity of fuel. Biofuel production includes chemical,
thermal, and enzymatic methods. Among all methods, the most effective way
to produce biofuels is through enzymes or biocatalysts [11]. Enzymes are
becoming the focus of research to produce biofuels because of their advan-
tages over other biofuel production techniques [12]. In this chapter, we dis-
cuss biodiesel production using biocatalytic processes and methods where
different microbial enzymes (obtained from microorganisms) are used.

1.2 Importance of Biodiesel Over Conventional


Diesel Fuel
Chemically, biodiesel is composed of fatty acid alkyl esters (FAAEs) which
are mono-alkyl esters of either fatty acid methyl esters (FAME) or fatty
4 Biodiesel Technology and Applications

acid ethyl esters (FAEE) depending upon the alcohol (acyl acceptor) being
used in the reaction [10]. Rudolf Diesel, the inventor of diesel engine, first
used biodiesel in 1900 but that was highly viscous so that engine could
not run effectively for a longer time [13]. Biodiesel is very suitable alter-
native to diesel fuel because of its remarkable properties and advantages,
i.e., biodiesel carries 4.5 times greater energy than fossil fuel [14] and sim-
ilar in chemical structure and energy content to conventional diesel [15].
It reduces approximately 85% carcinogenic compounds emission that is
why it is very less toxic than conventional diesel fuel, free of sulfur, free of
polycyclic aromatic hydrocarbons and metals, biodegradable, high cetane
number (CN), and flash point [16]. It has the potential to reduce pollutants
and emission of greenhouse gases [17] and is 66% more efficient lubri-
cating agent than petro-diesel, which enhances life and performance of
engine [18]. Blending of biodiesel with petro-diesel fuel that can affect
important properties of fuel such as flash point, CN, kinematic viscosity,
and lubricity is enhanced. It also decreases exhaust emissions and heat of
combustion [19]. The largest biodiesel producer is EU (European Union)
and biodiesel accounts 80% of the overall transport fuel in EU [20–22].
Biodiesel produced from different resources will have different composi-
tion and properties, but it must fulfill the standards and requirements of
international standards of American society for testing materials and EU
standards for biodiesel. Biodiesel has lots of applications such as it can be
used as a fuel for aviation purposes [20], for electricity production using
generators [21, 22] and in diesel fueled marine engines, because of its non-
toxic and biodegradable properties environmental impacts on engines can
be reduced. Alcohol type, quality of substrate that is to be converted, cata-
lyst used, temperature of the reaction, and alcohol-to-oil molar ratio deter-
mine the performance of biodiesel production [23–25].

1.3 Substrates for Biodiesel Production


Biodiesel feedstock accounts for 60%–80% of the total cost; therefore,
appropriate feedstock is required for economically valuable production
of biodiesel [26]. In order to obtain economically beneficial and sustain-
able biodiesel, feedstock must be easily available, cheap, and sustainable.
Feedstock is selected on the basis of biodiesel production that must be
compatible to chemical composition and properties of feedstock to be
used, percentage per dry biomass, agricultural potential, yield per hect-
are, and geographical region of that feedstock [27]. For example, soybean
oil, palm oil, coconut oil, and rapeseed oil are mainly used as feedstock in
Biocatalytic Processes 5

US, tropical countries like Indonesia, coastal areas, and European coun-
tries, respectively. Cultivation and climate conditions of the feedstock pro-
duction area are also considered for its selection [28]. Depending on the
nature, there are two types of feedstock for biodiesel production. First is the
lipid raw material and second includes alcohol feedstock. Lipid sources can
be divided into three categories, i.e., oils derived from plant sources (edible
and non-edible oils), waste oils (waste cooking oils, industrial wastewater,
lard, yellow grease, and animal fats), and oils from oleaginous microorgan-
isms such as bacteria, fungi, and microalgae [29]. Properties of biodiesel
like cold filter plugging point and oxidation stability are determined from
the feedstock used for production. Feedstock properties like moisture con-
tent, impurities, content, and composition of free fatty acids (FFAs) affect
the performance of engine [27, 28]. Composition of fats and oils including
monoglycerides, diglycerides, and triglycerides are used for biodiesel pro-
duction. Utilization of edible plant oils as feedstock is an expensive way for
biodiesel production that leads to imbalance in food market and indus-
try. It is also associated with some environmental problems like disruption
of vital soil resources and deforestation due to mass propagation [29]. In
order to solve problems linked with edible plant oils, the best alternate is
the production of second-generation biodiesel which is produced by using
non-edible (inedible) feedstock which are more favorable than edible oils
due to reduction in cost and waste pollution, lower aromatic, sulfur con-
tents, and high calorific value [8]. Inedible oils involve inedible plant oils,
industrial waste, cooking oils, animal fats, and microalgal oils. Inedible oil
producing plants have certain remarkable features that make them favor-
able to use, for example, they can be managed to grow in arid and semi-arid
conditions and they do not require fertilizers and moisture for growth [24].
Repeated use of fried vegetable oils at high temperature leads to the
production of waste cooking oils. Moreover, chemical composition of
waste cooking oil is totally dependent on the oil from which it is derived.
Hydrogenation, oxidation, and polymerization are the main chemical
reactions that lead to production of very toxic and detrimental compounds
for consumption. Fatty acid content of these oils lies in the range of 0.5%
to 15% which is very much higher than refined oil having fatty acid con-
tent less than 0.5%. The waste cooking oil is known as yellow grease if the
fatty acid content is less than 15% and it is called low value brown grease
if the fatty acid content is higher than 15% [28]. Animal waste products
like lard, tallow, animal fat, poultry fat, fish oil, and pork fat are also very
effective feedstock for biodiesel production [30, 31]. Animal-based bio-
diesel is a good lubricating agent and has high percentage of saturated
fats which decreases sedimentation risk and low temperature fluidity.
6 Biodiesel Technology and Applications

Moreover, it increases oxidative stability and cold filter plugging point of


biodiesel which are the characteristics of good quality biodiesel. Utilizing
these waste materials is an effective solution to encounter waste disposal.
Apart from all these mentioned advantages of non-edible or waste oils,
there are also some shortcomings or disadvantages, for example, low oil
yield, higher carbon residue, unsaturated fatty acid content, and low vola-
tility [29]. In some cases, large plantation land for inedible oils is required
compared to edible ones, e.g., Pongamia pinnata and Jatropha has 2–50
folds less oil yield per hectare than palm oil so that is why they require
much area to meet the demand [31]. Because of the drawbacks associated
with second-generation biodiesel, scientists are looking for more efficient
methods for biodiesel production. Biodiesel production using oleaginous
microorganisms like bacteria, algae, microalgae, and fungi are considered
as the future of biodiesel production that can meet global biodiesel demand
for transportation fuels and other energy consuming applications [32].
Microbial oils are better than other plant oils because of their short life
cycle and rapid growth, less requirement of space, labor, and easier scaling
[33–35]. Microalgae as a feedstock is very effective because of its enormous
advantages like they have high oil yield, can grow in salty and waste waters,
use of non-arable land, and growth in 24 hours so multiple harvesting in a
year is possible. If we give land area for microalgal growth then according
to an estimate, only 2% of the US cropping land is enough for meeting
1/3 demand of US transportation fuels and less than 5% land is required
to completely replace all transportation fuels [36–38]. Moreover, dry algal
biomass can accumulate more than 80% oil without water and they have
a capacity to produce oil yield 250 times greater than soybean water free
oil [35]. Some examples of microalgae used for biodiesel production are
Botryococcus sp., Cylindrotheca sp., Schizochytrium sp., Chlorella sp., and
Nitzschia sp.

1.4 Methods in Biodiesel Production


There can be many ways for biodiesel production but esterification and
transesterification are the two most widely used methods. Esterification is
the reaction of FFAs and alcohol to make FAAEs and water is released, while
transesterification is the reaction of triglycerides or triacylglycerols (TAGs)
with alcohol to make FAAE and glycerol is produced as by-product [9].
Transesterification is slower than esterification process because of its multi-
ple steps or reactions. It is a three-step process to convert TAGs into FAAE.
In the first step, TAG reacts with one molecule of alcohol to produce one
Biocatalytic Processes 7

molecule of FAAE and diacylglycerol (DAG). In second step, DAG further


reacts again with one molecule of alcohol to produce one molecule of FAAE
and monoacylglycerol and in the last step monoacylglycerol is converted
into one molecule of glycerol and FAAE after reacting with an alcohol mole-
cule. In each of these three steps, FAAEs are produced and in total one mol-
ecule of TAG and three molecules of alcohol are consumed to produce three
molecules of FAAE and one molecule of glycerol [6–10]. Transesterification
is a reversible reaction, and in order to make the reaction go forward to
produce more biodiesel, we have to supply alcohol in large excess so that the
reaction equilibrium shifts toward the product [36, 37].

1.5 Types of Catalysts Involved in Biodiesel


Production
Biodiesel production process is carried by either catalytic or non-catalytic
methods. Non-catalytic methods include use of alcohols or supercriti-
cal fluids or ionic liquids in the reaction system to produce biodiesel but
mostly catalytic methods have been used for last 2 or 3 decades because of
their advantages over non-catalytic methods [38]. Catalytic methods can
be categorized into chemical homogenous catalysts, solid heterogenous
catalysts, and biocatalysts.

1.5.1 Chemical Homogenous Catalysts


Chemical homogenous catalysts include combination of base and acid
catalysts. NaOH, KOH, and methoxides are the base catalysts while HCl
and H2SO4 are the acid catalysts [39]. Acid catalysts are mostly used to
overcome the problem of FFAs in the reaction system but the rate of trans
esterification by acid catalyst is slower than alkaline or base catalysts [8].
Chemical catalytic processes either alkaline or acid catalysis both have sev-
eral disadvantages. Alkaline catalysis provides high conversion of triacyl
glycerol into the alkyl esters in a very short time but it has many draw-
backs. Alkaline catalysis is very prone to FFA concentration (>2.5%) in the
reaction system because high FFA concentration results in saponification
reaction producing soaps and leads to loss in enzymatic activity and makes
difficult to separate transesterification by-product, i.e., glycerol from bio-
diesel. Hence, biodiesel yield decreases. Moreover, it needs high energy
requirement [40]. To counter FFA problem, acid catalysts are used, e.g.,
sulfuric acid but it also causes some technical problems regarding separa-
tion and purification of glycerol. Moreover, acid catalysis is a slow process
8 Biodiesel Technology and Applications

compared to alkaline process. Reactors, pipelines, and other equipment are


badly affected by acid catalysts because of their corrosive nature that can
increase the cost of biodiesel production [41].

1.5.2 Solid Heterogeneous Catalysts


Solid heterogeneous catalysts include acid heterogenous catalysts and base
heterogenous catalysts. Solid acid heterogenous catalysts include hetero-
polyacid catalysts (HPAs), mineral salts, acids, and cationic exchange res-
ins. Among these, titanium oxide, sulfonic ion exchange resin, tin oxide,
sulfonated carbon-based catalysts, zirconium oxide, zeolites, and sulfonic
modified mesostructured silica are the main acid heterogeneous catalysts.
Solid base heterogeneous catalysts have been categorized as mixed metal
oxides, supported alkaline earth metals, single metal oxides, and nano-­
oxides. Among these, the most studied are magnesium oxide, calcium
oxide, and strontium oxide [44, 45].

1.5.3 Biocatalysts
Biocatalysts include enzymes especially lipases which are very popular in bio-
diesel production [43]. Enzymatic biodiesel production method diminishes
problems associated with alkali and acid catalyzed methods. Use of enzyme
catalysts has several economic and environmental advantages over chemi-
cal biodiesel production processes. Advantages of enzyme catalysis include
production of pure and high market value glycerol, minor, or no waste water
generation that is why treatment of waste water is not required, mild reaction
conditions are required, no soap formation because enzymes can esterify low
quality feedstock having high concentration of FFA that is why this method
is insensitive to feedstock concentration. Enzymatic biodiesel production is
simple so energy consumption is very low, enzymes can be reused because of
their easy separation from the reaction mixture, and overall chance of con-
tamination is lower than other transesterification methods [13].

1.6 Factors Affecting Enzymatic Transesterification


Reaction
There are a lot of factors effecting enzymatic transesterification reaction
such as source of enzyme, its type, preparation method, applying tech-
nique, its dosage, activity, and life time. Apart from these enzymes related
factors, there are also some other factors which affect transesterification
Biocatalytic Processes 9

reaction, e.g., feedstock type and its quality, type of alcohol as acyl accep-
tor, reaction pH, presence or absence of solvent, type of solvent, reaction
temperature, alcohol-to-oil molar ratio [46].

1.6.1 Effect of Water in Enzyme Catalyzed Transesterification


Presence of water is not only required for chemically catalyzed biodiesel
production but also very much required for enzymatic biodiesel produc-
tion. It helps in maintaining enzyme structural confirmation and stabil-
ity so it directly affects activity of enzyme. Oil-water interface is required
for enzyme-substrate complex to proceed and water helps to increase this
interfacial area [44]. So, without water, transesterification is not possible
and absence of water can lead to permanent or temporal changes in protein
(enzyme) structure. If water content is minimal, then increase in water con-
centration moves the reaction equilibrium toward more hydrolysis. Thus, it
enhances reaction rate by providing greater stability to enzyme [45]. Excess
of water content also has some negative effects on the reaction as well as on
enzyme. Excess water content can be accumulated in the reaction medium
and within enzyme active site, that leads to decrease the reaction rate as
well as its alkyl ester yield [46]. So, concentration of water should be opti-
mally perfect in order to gain maximum benefit from it. Every enzyme has
its specific water content requirement, i.e., optimal water requirement, at
which that particular enzyme performs its best [47, 48]. Optimal water con-
tent not only provides great support, flexibility, and stability to the enzyme
but also maximizes transesterification yield by diluting methanol that has
an inhibitory effect on enzyme. Factors that determine optimal water con-
tent include feedstock and type of solvent used, enzyme, and its immo-
bilization technique used [48]. Chaudhary et al. [49] studied the effect of
water content in lipase catalyzed transesterification. At low water activity
(aw = 0.33), synthetic activity of enzyme was increased and at high water
activity (aw = 0.96) enzyme became more hydrolytically active. They tested
various enzymes/lipases at different water activity to check transesterifica-
tion rates. The lipase from Aspergillus niger was found more prominent to
give maximum transesterification rate of 0.341 mmolmin−1 mg−1 at aw = 0.75.
Measuring water content as weight percentage is a better choice and more
convenient to use than water activity (aw), measured by Karl-Fischer method
[50]. Maximum methyl ester yield was at water concentration of 10-15%
while increasing water content from 0% to 40% to study the effect of water
in conversion of salad oil into methyl ester. But after much increased water
concentration, methyl ester yield became very low. So, for maximum trans-
esterification yield, optimum water concentration is required.
10 Biodiesel Technology and Applications

1.6.2 Effect of Bioreactor


In order to maximize production and benefit of product we need to per-
form optimized laboratory experimental procedure at a large industrial
scale, so, bioreactors are used in this regard. But results should be equiva-
lent to laboratory procedure [45]. There are some complications like pro-
duction should be cost effective and in good quality. Carefully planned
methodologies and objectives should be designed for effectively large-scale
production. This also includes bioreactor parameters like fluid flow per-
formance and unexpected environmental variation. In case of industrial
transesterification process, the main hurdle is multiphasic nature of lipase
catalyzed synthesis and hydrolysis because this does not allow the bioreac-
tor equivalent to laboratory experiment. Many types of bioreactors such as
fluid beds, recirculation membrane reactors, expanding beds, static mixers,
batch stirred tank reactors (STRs), and packed bed reactors (PBRs) have
been used for enzymatic biodiesel production [51, 52]. One of the leading
differences between STRs and PBRs is presence of enzyme at specific loca-
tion in reactor, e.g., in STRs it is dispersed in the reaction mixture but in
PBRS it is fixed in a column. STRs are the simplest type of bioreactors con-
taining just reactor and propeller that stirs reaction mixture mechanically.
Batch operated STR need to be empty, clean, and again add reactants for
the reaction in order to start new batch process and this is main reason of
batch process to produce less yield of the product. Solution of this problem
is to use STRs with continuous mode. This does not require to remove
enzyme and ingredients to start another cycle. There is a filter attached at
the reactor outlet that preserves enzyme in the reactor [52]. PBRs can also
be used in both batch and continuous mode but later is more advantageous
because of its low labor cost, stable and automated controlled operating
conditions, high efficiency, protects enzyme from shearing stress, continu-
ous glycerol removal, and ease of maintenance [53–55]. Currently, most of
the bioreactors are used in batch mode with STRs but a lot of research has
been done on PBRs usage and its optimization for enzymatic biodiesel pro-
duction to find this PBR method is better than batch mode STR [56–59].

1.6.3 Effect of Acyl Acceptor on Enzymatic Production


of Biodiesel
Alcohols are mostly used as an acyl acceptor for biodiesel production. To
get maximum economical profit at industrial scale, acyl acceptor (alcohol)
should be cheap and readily available and that is why ethanol and especially
methanol are widely used for this purpose. Usually, three moles of alcohol
Biocatalytic Processes 11

are required for each mole of oil and in order to keep the reaction moving
forward [56]. By increasing alcohol concentration, yield also increases but
up to a certain limit [57]. Methanol as an acyl acceptor is frequently used for
biodiesel production [58] because it is less expensive, has low chain length,
more volatile, and more reactive, and gives high yield than other alcohols
[55]. A lot of research has been done utilizing methanol as acyl acceptor to
convert various types of oils such as soybean oil, jatropha oil, and canola
oil, in the presence of free or immobilized lipase 96.4% yield of FAME was
obtained from microalgal oil using methanol as acyl acceptor in the pres-
ence of Candida rugosa lipase immobilized on bio-silica polymer [59, 60].
Different alcohols with different substrates may result in different yield so
alcohol-substrate combination should be kept in mind for maximum output.
Excess of methanol causes inhibitory effect in the reaction because it changes
the stability and configuration of biocatalyst/lipase that can leads to partial or
complete inactivation of lipase [60–62]. Moreover, it also causes hindrance
in separation of glycerol [61, 62]. Methanol inhibition was observed with
Novozym® 435 lipase in transesterification of waste oils [63], microalgae
oils, and various vegetable oils [64–67]. Inhibitory effect was also observed
with some other lipases such as lipases obtained from Rhizopus oryzae and
Burkholderia glumae [65]. Addition of alcohol in each step should be done
after considering type of substrate and enzyme to determine alcohol sub-
strate molar ratio [66]. This method of sequential addition of alcohol in reac-
tion system was first performed by [67]. 98% biodiesel yield was obtained
utilizing T. lanuginosus lipase to convert soybean oil using stepwise addition
of methanol [68]. Inhibition of lipases such as C. rugosa lipase, P. cepacia
lipase, R. oryzae lipase, and P. fluorescens lipase was prevented using step-
wise addition of methanol and 90% yield was also obtained by converting
waste cooking oil into biodiesel [69]. Methanolysis of olive oil increases by
34% using stepwise addition of methanol compared to batch methanolysis
[70]. Transesterification of waste cooking oil using Novozym 435 was also
reported to yield 93% and 96% conversion for continuous and batch process
and lipase did not lose its activity even after 20 cycles [50]. Three-step addi-
tion of methanol resulted in 97% conversion of plant oil with 0.25- to 0.4-h
intervals. But this method of stepwise addition requires low level mainte-
nance of methanol concentration so it cannot be effectively used for indus-
trial scale. Alcohols as an acyl acceptor other than methanol include high
chain primary alcohols, secondary, branched, and linear chain alcohols such
as ethanol, isopropanol, t-butanol, and octanol [71].
Choice and selection of appropriate alcohol is important as it can
influence some biodiesel properties like lubricity and cold flow proper-
ties [75, 76]. Moreover, high chain alcohols cause less lipase inhibition
12 Biodiesel Technology and Applications

and produce high yield than methanol because lipase show more affin-
ity toward higher chain alcohols than lower chain [12]. The most widely
used alcohol as an acyl acceptor after methanol is ethanol as it is less
inhibitory, less toxic, and derived from renewable resources [73] unlike
methanol which is derived from coal and natural gas. There is minor
difference between characteristics of fuels obtained after methanol and
ethanol, i.e., FAME and FAEE, respectively. As FAEE has large viscosity
and lower pour and cloud points [74, 75]. Hernandez-Martin and Otero
[76] showed that, Novozym 435 catalyzed transesterification of sun-
flower oil, that was performed using methanol and ethanol separately
to check which acyl acceptor would perform better. Methanol-mediated
transesterification showed more lipase inhibition than ethanol contain-
ing reaction. Moreover, ethanol transesterification reaction was faster
than methanol reaction. Acyl acceptors other than alcohols can also be
used as an alternative such as methyl acetate, ethyl acetate, and dimethyl
carbonate (DMC). Methyl acetate was utilized as an acyl acceptor for
transesterification of soybean oil catalyzed by Novozym 435. In addition,
92% methyl ester yield was obtained [77]. Similarly, >90% ethyl ester
yield was obtained when utilizing ethyl acetate as an acyl acceptor for
transesterification catalyzed by Novozym 435 [78].
Use of DMC resulted in over 90% yield even after 10 times reuse of
Novozym 435 lipase to convert Chorella sp. KR-1–derived triglyceride
[79]. But use of methyl acetate and ethyl acetate is cost expensive and
also make the product difficult to separate. Another strategy can be used
to reduce methanol inhibition problem, i.e., use of solvents in the reac-
tion mixture [80]. Use of solvents is beneficial for various reasons such
as it increases solubility of alcohol and glycerol that results in preven-
tion of lipase denaturation [81]. It increases the rate of reaction because
it improves mass transfer rate. Use of solvents do not allow to form new
separate phase that hinders enzyme activity because it dissolves most part
of alcohol that makes a separate phase if remained undissolved. Moreover,
it reduces viscosity and stabilizes lipase [45, 55]. Enzyme stabilization is
associated with the presence of water molecules and their activity sur-
rounding the lipase structure. So, use of polar, less hydrophobic solvents
is not a good idea because that can lead to distortion of enzyme con-
firmation [82]. A higher yield of FAME was obtained from microalgae
lipids catalyzed by intracellular lipase when non-polar n-hexane solvent
was used as compared to polar tert-butanol solvent [83]. Organic solvents
such as hexane, petroleum ether, tert-butanol, n-heptane, and ionic liq-
uids are widely used for lipase catalyzed transesterification purpose [88].
Sometimes, it also happens that use of solvents becomes necessary in
Biocatalytic Processes 13

transesterification reaction if short chain alcohols are being used as an


acyl acceptor in order to completely dissolve alcohol and produce max-
imum output but for the same reaction conditions solvent-free reaction
system can be used if higher chain alcohols are used as an acyl acceptor.
Iso et al. [85], immobilized P. fluorescence lipase catalyzed transesterifica-
tion was performed using methanol and ethanol as an acyl acceptor. They
also provided 1,4-dioxane solvent to the reaction to carry out effective
transesterification reaction. But when they used propanol and butanol as
an acyl acceptor, they did not provide any solvent to the reaction because
addition of solvents was not necessary required for transesterification.
Without solvent reaction worked and appropriate result was obtained.
Similar type of findings was also observed in another experiment that
in which hexane as solvent was used for methanolysis of various oils or
substrates like rapeseed oil, soybean oil, recycled restaurant grease, and
tallow. That reaction was catalyzed by C. antartica lipase (SP 435) and
M. miehei lipase (lipozyme IM 60) [53]. Solvents stabilized the lipase activity
shown by Li et al. [86], where lipase AK did not loss its activity. 1,4-dioxane
was used as solvent and gave higher yields. Presence of t-butanol as sol-
vent also resulted in the improvement of methyl ester yield [84].
Effects of various solvents like benzene, tetrahydrofuran, chloroform,
and 1-4 dioxane were investigated using different enzymes such as P. cepa-
cia (Lipase PS), C. rugosa (Lipase AY), M. javanicus (Lipase M), P. fluo-
rescens (Lipase AK), and R. niveus (Newlase F) by [85]. Use of solvents
definitely increases the reaction rate and solubility of alcohol which is ben-
eficial but it is not economical and environment friendly because to sepa-
rate organic solvent from reaction mixture a solvent recovery unit is also
required that increases the cost of recovery. Its flammability and toxicity
is another concerning factor [87]. Choice of lipase according to alcohol
is also important as some lipases show more resistance toward different
alcohols. Yang et al. [88] found that Photobaterium lipolyticum lipase was
more tolerant to methanol inhibition than C. antartica lipase B (Novozym
435) when transesterification was performed using one step methanol
addition. Similarly, Pseudomonas lipases were found to be more alcohol
tolerant compared to lipases from R. miehei and T. lanuginosus. That is why
pseudomonas lipases have higher methanol-to-oil molar ratio. Out of nine
lipases, only Pseudomonas cepacia lipase showed high ester yield from soy-
bean oil with 8.2:1 methanol-to-oil molar ratio [93, 94]. A recent and novel
approach of countering methanol inhibition is addition of silica gel in the
reaction system. Silica gel absorbs methanol and keeps its concentration
level below to prevent lipase inhibition. But presence of silica gel makes
separation of products difficult.
14 Biodiesel Technology and Applications

1.6.4 Effect of Temperature on Enzymatic Biodiesel Production


Use of enzyme in any chemical reaction makes the reaction less energy
intensive, but, like every other chemical reaction, increase in temperature
enhances reaction speed and rate. Similarly, in case of enzymatic biodiesel
production, increase in temperature increases enzyme activity, reaction
speed, its rate, and production yield [45]. When Lipozyme TL IM lipase
was used to transesterify crude palm oil using methanol then the resulting
FAME yield was 96.15% and 85.86% at 40°C and 30°C, respectively [95,
96]. But this effect is limited to certain extent because beyond enzyme opti-
mum temperature, enzyme structure becomes unstable and that leads to
enzyme denaturation and reaction becomes slower and yield also decreases.
Novozym 435 catalyzed biodiesel production from microalgal lipids, there
was 19% decrease in product yield when temperature went from 45°C to
55°C, i.e., higher than optimum temperature [97, 98]. Enzyme tempera-
ture should remain below the boiling point of alcohols being used in the
reaction system to avoid evaporation of alcohol. In case of methanol and
ethanol-mediated transesterification, reaction temperature is 65°C and
78°C, respectively [8]. Free bacterial lipases are considered thermally sta-
ble but if they get immobilized thermal stability increases [45]. Optimum
enzyme temperature is influenced by lipase thermal stability, type of sol-
vent, ­alcohol-to-oil molar ratio, and lipase immobilization. Every enzyme
has different optimum temperature depending on the source and type
of enzyme. Normally, lipases have optimum temperature range that is
20°C–70°C. Optimum temperature for C.antartica lipase is 40°C [99, 100].

1.6.5 Effect of Glycerol on Enzymatic Biodiesel Production


Glycerol is another product obtained along with TAG in transesterification
reaction for biodiesel production when alcohols are used as an acyl acceptor.
Like methanol inhibition effect, glycerol also hinders enzymatic transester-
ification. Production of glycerol cause reversal of reaction equilibrium and
thus opposes biodiesel production. Glycerol insolubility in the reaction sys-
tem makes it to accumulate in the system and increases the viscosity of the
reaction mixture. Not only it just increases viscosity but because of its hydro-
philic nature, it surrounds the lipase and hence prevents substrate to interact
with enzyme and binding at enzyme catalytic site. This effect is more promi-
nent when enzyme is in immobilized state. These things all together impacts
negatively on the transesterification reaction [93]. According to a research
study, rapeseed oil was transesterified using ethanol as an acyl acceptor with
different immobilized enzymes such as Novozym 435, Lipozyme TL HC
Biocatalytic Processes 15

(immobilized on polymeric resin) and Lipozyme TL IM (immobilized on


silica). Among all reactions, glycerol became more accumulated and hin-
dered reaction when silica was used for immobilization purpose because it
had large number of micropores that helped in accumulation [94].
A lot of methods have been devised to overcome and tolerate glycerol
inhibition problem such as continuous removal of glycerol, use of solvents,
and use of acyl acceptors other than alcohols. Biodiesel production in PBR
is effective for continuous production and also to tolerate glycerol inhibi-
tion because it allows continuous removal of glycerol from it [95]. Bélafi-
Bakó [96] showed that 97% conversion yield was obtained by methanolysis
due to continuous removal of glycerol. Use of solvents is another strategy
to resolve glycerol inhibition problem. Solvents, for example, tert-butanol
and ionic liquids dissolve glycerol and thus reduce the glycerol inhibition
problem. Moreover, lipases also perform better in the presence of solvents
[97]. Azócar et al. [4] inferred that inhibition effect was eliminated when
tert-butanol was used as solvent to convert soybean oil into biodiesel pro-
duction in a continuous way because it is an excellent solvent for methanol
and glycerol to dissolve in it and, thus, reduces inhibitory effects of both
methanol and glycerol. These methods have shown promising work but
these are not as good for industrial scale production.
There is another novel method in which instead of alcohols other
compounds like methyl acetate, ethyl acetate, and DMC are in use. Use
of methyl acetate or ethyl acetate does not produce glycerol as a product
along with the main product instead it leads to produce triacetylglyc-
erol that does not inhibit any enzymatic or reaction activity and further
downstream processes are not halted [98]. According to Zhang et al. [99],
transesterification of palm oil was done using Novozyme 435 as catalyst
and DMC as acyl acceptor with reaction conditions were 10:1 DMC to oil
ratio, 55°C reaction temperature and 20% lipase in a solvent-free system.
In addition, 90.5% conversion yield, i.e., FAME was obtained, and after
eight reaction cycles, no reduction in enzyme activity and loss of yield was
observed. It was just because glycerol was not produced instead glycerol
dicarbonate was formed because of DMC as acyl acceptor for the reaction.
After discussing all negativity about glycerol, it seems glycerol does not
have any positive effect but it’s not true. Glycerol if purified absolutely from
the transesterification, as mostly huge biodiesel producing companies do,
has vast number of uses in diverse industrial fields. Moreover, 99.7% puri-
fied glycerol can be used as a raw material for various types of fields such as
paints, toiletries, animal feed, emulsifiers, pharmaceuticals, textiles, drugs,
tobacco, cosmetics, toothpaste, leather, plasticizers, paper, food, and for
different chemicals production [100, 101].
16 Biodiesel Technology and Applications

1.6.6 Effect of Solvent on Biodiesel Production


One of the major problems in enzymatic biodiesel production is enzyme
inhibition by short chain alcohols such as methanol and ethanol that are
used in the reaction. These alcohols’ insolubility in the reaction system
denatures the enzyme and hence reduces yield of biodiesel. So, solvent
application plays its role in this regard. Organic solvents are used to solubi-
lize these excessive alcohols so that enzyme denaturation can be prevented.
Hence, it stabilizes the enzyme. Solubility of oils and alcohols become
increased due to presence of organic solvents, this provides the required
environment for substrate to interact with enzyme at its active site. Organic
solvents also reduce viscosity of the reaction mixture and enhances mass
transfer toward the enzyme that leads to improved reaction rate [101].
Organic solvents also eliminate the need of stepwise addition of alcohol.
All these things in combination increase production of biodiesel. Organic sol-
vents that are commonly used include tert-butanol, petroleum ether, hexane,
and n-heptane [102]. Some other organic solvents that are used are 2-butanol,
cyclohexane, isooctane, acetone, 1,4-dioxane, and chloroform. While consid-
ering nature of organic solvents, hydrophobic organic solvents are majorly
used. Hydrophobicity of the organic solvents helps in accumulating water
molecules around enzyme which is important for enzyme structural stability
[103]. Polar or hydrophilic solvents work opposite to hydrophobic organic
solvents by playing role in distortion of enzymatic structure. But solvent with
little polarity can be beneficial to dissolve oil and alcohol. For example, hydro-
philic 1,4-dioxane and tert-butanol have produced some good results by pro-
ducing high enzymatic transesterification yield [104]. Tert-butanol, having
moderate polarity, eliminates glycerol and methanol inhibition problem for
enzyme because it can dissolve both in itself. This makes the enzyme more
stable and active and then ultimately produce better reaction yield [105]. Tert-
butanol is the most common solvent that proved its effectiveness in various
cases. According to Royon et al. [84], cottonseed oil was transesterified in the
presence of Candida antartica lipase. Methanol was found to be the cause of
enzyme inhibition in the reaction but when tert-butanol was used as solvent,
reaction yield goes up to 97% with minimal enzyme inhibition. Similarly,
in another research experiment, tert-butanol was tested for its effectiveness
when rapeseed oil was used as substrate for biodiesel production. In sol-
vent-free system, methyl ester yield was 10% but after utilizing tert-butanol
yield was 75%. But under optimum conditions having Lipozyme TL IM and
Novozyme 435 both in the reaction system, biodiesel yield reached 95% and
the reaction was so stable that enzymes did not lose their activity even after
200 cycles. Reaction was favored and well supported by tert-butanol [86].
Biocatalytic Processes 17

Use of solvents provide many benefits but they also come with some
disadvantages such as organic solvents do not completely dissolve glycerol,
by-product of the reaction, that causes the enzyme to lose its activity and
become unstable. Use of solvents also make the process very costly because
there is a need of extra purification step to separate out solvent and product
from the reaction mixture. Organic solvents are mostly toxic and highly
flammable so there are also environmental and health concerns while
using them [11]. In order to tackle problems of conventional organic sol-
vents, researchers have suggested some alternatives. Diesel oil was found
to be an interesting alternative but the most recent, beneficial, and popular
alternatives are super critical carbon dioxide (SC-CO2) and ionic liquids
(ILs). Researchers have also confirmed the positive effect of using SC-CO2
and ILs in the enzymatic transesterification [106, 107].

1.7 Lipases as Biocatalysts for Biodiesel Production


Transesterification of oils for biodiesel production is done using either
chemical or enzymatic catalyst [108]. An enzymatic catalyst is used at first
place due to their normal reaction conditions, reusability, easy products
separation, and production of high-quality product. There is less energy
consumption in enzyme catalysis as it occurs at a low temperature as
compared to chemical catalysis requiring high energy consumption [109,
110]. Further, enzymatic catalysis is environment-friendly as there is no
wastewater production and produces pure biodiesel as compared to chem-
ical catalysis [107]. Among enzymatic catalysts, lipase with excellent bio-
chemical and physiological properties is most commonly used to catalyze
the transesterification process. Lipases play their role in several industrial
processes like alcoholysis, acidolysis, amynolysis, and hydrolysis reactions
but their leading role in biodiesel production is considered very important
[108–111]. The use of lipase in biodiesel production is proved to be bene-
ficial due to its characteristics like high efficiency, convert FFAs completely
into methyl/ethyl esters, reaction specificity, require low temperature, min-
imum energy consumption, and fewer side products [109]. Lipases belong
to class “hydrolases” as they carry out hydrolyses of triglycerides produc-
ing glycerol and fatty acids from it in an oil-water interface [110]. A general
reaction for biodiesel production using lipase is as follows:

lipase
Triglycerides + 3 Alcohol 3 Fatty acid methyl ester + Glycerol
(biodiesel) (by-product)
18 Biodiesel Technology and Applications

Lipases work on specific substrates and carry out catalysis of hetero-


geneous reactions in water-soluble as well as insoluble systems. Further,
lipases have the properties like chemo-specificity, region-specificity, and
stereo-specificity [111]. When classification is made based on region-­
specificity, there come three classes of lipases: 1) non-specific lipases, 2) 1,3-
specific lipases, and 3) fatty acid-specific lipases. Non-specific lipases have
ability to attach with all the possible positions of triglycerides to give FFAs
and glycerol. The intermediates of the reaction, diglycerides, and mono-
glycerides do not accumulate in the reaction as they are instantly hydro-
lysed into fatty acids and glycerol [112]. 1,3-specific lipases are specific for
the 1 and 3 positions of triglycerides and remove fatty acids from these
positions. 1,3-specific lipases carry out the conversion of triglyceride to
diglycerides much faster than diglyceride to monoglyceride [113]. Fatty
acid-specific lipases carry out hydrolysis of a specific type of esters which
have double bonded long chains of fatty acids in cis position between C-9
and C-1. Hydrolysis of esters with unsaturated fatty acids occur slowly and
such class of lipases is not much common [114]. All the hydrolytic enzymes
including lipases have common folding pattern involve in a hydrolytic
activity called α/β hydrolase fold which is made up of a β sheet of eight
strands (one of which is antiparallel while remaining seven strands are
parallel) connected by α helices. Histidine residue, catalytic acid residue
and Nucleophilic residue are present in α/β hydrolase fold. Pentapeptide
sequence (Gly-X-Ser-X-Gly) which is a highly conserved in most of the
lipases involved in the construction of ‘nucleophilic elbow’ which is a typ-
ical β-turn-α motif having active nucleophilic serine residue between a β
strand and an α-helix. Catalytic triad made up of amino acids like his-
tidine, serine, and aspartic acid or glutamic acid build the active site of
lipases. The same catalytic triad is seen in serine proteases predicting com-
mon catalytic mechanism in them. Amphiphilic α helix peptide sequence
forms a lid or flap which covers the active site of lipase and has a struc-
tural variability depending upon the lipase source organism. Changes in
the structure of the lid are responsible for the activation/inactivation of
lipases [114]. Changes in the conformation of lipase structure as well as the
quality and quantity of interface being used in the reaction are responsible
for the activation of lipase. When the lipase enzyme meets the oil/water
interface there occur some changes in lipase structure that results in its
activation. For the activation of lipase first, the lid opens to uncover the
active site of lipase upon its contact with the ordered interface [115]. Due
to this restructuring of lipase, electrophilic region is created around serine
residue present in active site, lid hydrophilic side which was exposed in
native form now partly buried inside the polar cavity and hydrophobic side
Biocatalytic Processes 19

of lid completely exposed, thus creating a non-polar surface around the


active site for efficient attachment of lipid interface with it [115].

1.7.1 Mechanisms of Lipase Action


Lipases interact with ester bonds of their substrate like acylglycerols to
catalyze the reactions of hydrolysis, synthesis, and transesterification.
Triglycerides, which are insoluble and long chained fatty acids, are precisely
catalyzed by lipases [113]. Lipase carries out triglyceride oil transesterifica-
tion with methanol in three reversible steps with the first step for conversion
of triglycerides to diglycerides followed by the second step of diglycerides to
monoglycerides conversion, and finally, monoglycerides convert into glyc-
erol molecules. Here, each conversion step produces one FAME molecule;
hence, a total of three FAME molecule are produced from one triglycer-
ide [116]. Two models are mainly under discussion to describe the kinet-
ics mechanism for esterification reactions, Michaelis-Menten kinetics and
Ping Pong Bi Bi model. Lipase catalyzed esterification mainly elaborated by
Ping Pong Bi Bi mechanism which is a bi-substrate reaction that releases
two products. It involves following steps: 1) acyl-donor donate their acyl
group to the enzyme resulting in the formation of acyl-enzyme complex,
2) release of the water molecule as a product, 3) binding of acyl acceptor
with the enzyme complex, and 4) release of ester [117, 118]. Many research-
ers made some modifications in this model depending upon inhibiting
factors [118]. The catalytic activity begins with the transient tetrahedral
intermediate formation with a negatively charged carbonyl oxygen atom.
The reaction between the hydroxyl group oxygen present in nucleophilic
serine residue of lipase enzyme and activated carbonyl carbon of the sub-
strate involved building this transient tetrahedral intermediate. The inter-
mediate thus formed is stabilized by its interaction with two peptide NH
groups. After that nucleophilic hydroxyl group of water react with the car-
bonyl carbon of acyl-enzyme complex resulting in the formation of acyl
product and enzyme is released for further catalysis [119].

1.7.2 Efficient Lipase Sources for Biodiesel Producing


Biocatalyst
Lipases can be obtained from plants, animals, and microorganisms and
based on that lipases can be classified on it as plant, animal, and micro-
bial lipases depending on their origin respectively. Lipases from different
sources with different structure has different properties and catalytic activ-
ities. We can use this to counter lipase problems in biodiesel production
20 Biodiesel Technology and Applications

like lipase cost and methanol inhibition. We need to optimize reaction


conditions according to chosen substrate and lipase from specific source
[107]. Microbial lipases are widely used at industrial and commercial
level as biocatalysts for biodiesel production because microbial lipases are
more stable and can be produced in bulk amount from microorganism
[23]. Microbial lipases can be manipulated genetically with ease, seasonal
changes have nothing to do with lipase production, and rapid growth of
microbes makes them the ideal candidates as lipase source [120, 121]. Use
of microbial lipase is increasing day after day and currently 5% of the world
enzyme market is being shared with microbial lipases [121]. Microbial
lipase weighs around 30–50 kDa and their optimum pH to work at is 7.5.
Based on temperature tolerance, microbial lipases can be categorized as
mesophilic and thermophilic lipases. Mesophilic lipases normally work at
35°C–50°C and they become denature above 70°C while thermophilic or
thermostable lipases normally work at 60°C–80°C, but some also work at
100°C under specific conditions [122, 123]. According to Hotta et al. [123],
lipases from Pyrobaculum calidifonti (hyperthermophilic archae) showed
its activity at 90°C. Similarly, thermostable lipases from Caldanaerobacter
subterraneus and Thermoanaerobacter thermohydrosulfuricus (highly ther-
mophilic bacteria) performed well in the range of 40°C–90°C. They not
only performed well at high temperatures but also were resistant toward
organics solvents [124]. Lipase producing microorganisms can be iso-
lated from soil, waste water, marine water, and industrial wastes. Isolates
of Mucor, Sclerotina, Candida, and Aspergillus strains have been reported,
which were isolated from soil. Similarly, various strains of microbes such
as P. alcaligenes, Bacillus acidophilus, Enterobacter intermedium, P. flu-
orescens, and Geotrichum asteroids are also reported as lipase producing
strains when isolated from vegetable oil processing plants [141]. Screening
of microbes is done by checking their lipolytic. Generally, lipases are
screened using batch cultures having agar as substrate but this is time con-
suming. So, the two mostly used methods for lipase production are solid
state fermentation and using submerged culture [142]. Purified form of
lipases is used in biochemical reaction to get maximum benefit from it, so
purification is required. Purification of lipase requires several techniques
including ammonium sulfate precipitation or ultrafiltration and after that
more sensitive and advanced techniques are utilized like gel filtration, ion
exchange chromatography, and affinity chromatography [143]. Moreover,
some other novel techniques can also be applied to purify lipases such as
immunopurification, column chromatography, hydrophobic interaction
chromatography, and membrane process. Generally, the strategy used for
purification lipases starts with the removal of lipase producing cells from
Biocatalytic Processes 21

their growth culture to get extracellular lipases after fermentation. Then,


the extract without cells is concentrated by organic solvents, ultrafiltration,
and precipitation using ammonium sulfate. Ammonium sulfate precipi-
tation is used in the first stages of purification and it crudely separate out
things from mixture. After that advanced techniques of chromatography
are used to finely purify lipases [144]. According to Javed et al. [112],
diverse data from various research work suggested that purification had
been done from 2.4 to 500 folds with an increase in yield from 10.3% to
36%. Effective production and purification strategies of lipases are being
designed to get maximum yield at a very small expense. Among microbial
lipases, the most commonly used lipase sources are bacteria, fungi, yeast,
and algae; see Table 1.1 for some bacterial lipases and Table 1.2. for fungal
lipases that are used for biodiesel production [46]. Some of the commer-
cially available enzymes for biodiesel production are enlisted in Table 1.3.

1.8 Comparative Analysis of Intracellular and


Extracellular Lipases for Biodiesel Production
Transesterification reaction for biodiesel production is done with both
extracellular and intracellular lipases. Preference for their use is depen-
dent upon either we want simple upstream processes as in case of intra-
cellular lipase or high enzymatic conversion as in case of extracellular
lipase. But either we use intracellular lipase or extracellular lipase there is
no need for some downstream processes including separation and recy-
cling. Further, both immobilized lipase (extracellular) and immobilized
whole-cell lipase (intracellular) are proved to have highly efficient when
compared with free lipase used for transesterification [165]. Some exper-
imental studies for both intracellular and extracellular lipases are given
in Table 1.5. Intracellular lipases are the enzymes present inside the cells
or linked to the walls of cells producing it known as whole-cell biocat-
alysts. They are not purified or separated from their cells and used as a
whole-cell for transesterification (whole-cell biocatalyst) or immobilized
(whole-cell immobilization) [55]. Rhizopus and Aspergillus which are fila-
mentous fungi are most widely used as whole-cell biocatalyst for transes-
terification process [166]. As the main issue related to biodiesel production
at large scale is cost and the use of intracellular lipase for transesterifica-
tion resolves this problem like the use of intracellular lipase is considered
cost-effective because of the elimination of costly processes of lipase isola-
tion and purification before immobilization which are required in case of
extracellular lipase [167]. Intracellular lipase producing cells or whole-cell
22 Biodiesel Technology and Applications

Table 1.1 Some of the commonly used bacterial lipases for biodiesel production.
Acyl
Enzyme Immobilized on Substrate acceptor Yield Reference

Burkholderia Hydrophobic Jatropha oil Methanol 95% [125]


cepacia silica monolith
Lipase

Hybrid matrix of Jatropha Ethanol 100% [126]


alginate and curcas L.
κ-carrageenan oil

κ-carrageenan Palm oil Methanol 100% [127]

Modified Jatropha oil Methanol 94% [128]


attapulgite

SiO2-PVA Babassu oil Ethanol 100 % [129, 130]

SiO2-PVA Babassu oil Ethanol 100% [131]

Nb2O5 Babassu oil Ethanol 74.1% [132]

Pseudomonas Epoxy-acrylic Waste Ethanol 46–47% [133]


cepacia resin vegetable
Lipase oil

Phyllosilicate sol- restaurant Methanol 98% [134]


gel matrix grease and
Ethanol

Fe3O4 nanoparticle Soybean oil Methanol >88% [135]


biocomposite

Accurel Madhuca Ethanol 96% [136]


indica

Protein-coated Soybean oil Ethanol 98.93% [137]


microcrystals

Celite Jatropha oil Ethanol 98% [89]

Pseudomonas Octyl-silica resin Babassu oil Ethanol 97.5% [85]


fluorescence
Hydrophobic Soybean oil Methanol 65% [138]
sol-gel

Porous kaolinite Triolein Methanol 90% [139]


particles and
Ethanol

Asymmetric Triolein Methanol 80% [140]


membrane
Biocatalytic Processes 23

Table 1.2 Some of the commonly used fungal lipases for biodiesel production.
Acyl
Enzyme Immobilized on Substrate acceptor Yield Reference
Candida Activated textile Waste cooking Methanol 91.08% [157]
antartica cloth oil
Lipase
Polyurethane foam Soybean oil Ethanol 81% [158]
Acrylic resin Sunflower oil Ethyl acetate 92.7% [78]
Soybean oil Methanol 83.31% [95]
Candida Microporous bio Scenedesmus Methanol 96.4% [59]
rugosa silica-polymer quadricauda
Lipase microalgal
oil
Poly(styrene- Soybean oil Methanol 86% [159]
methacrylic
acid)
microsphere
within an activated Palm oil Methanol 70% [160]
carbon as
support
Thermomyces Olive pomace Pomace oil Methanol 93% [161]
lanuginosus
Phyllosilicate sol-gel Grease Ethanol 80-90% [162]
Lipase
matrix
Mesoporous poly- Oleic acid Methanol 90% [163]
hydroxybutyrate and
particles (PHB) Ethanol
Toyopearl Babassu oil Ethanol 86.6% [164]
AF-amino-650M
resin

biocatalysts are directly employed for immobilization without separation


and purification steps for lipase enzymes [55]. Porous biomass support
particles (BSPs) are mostly used for whole-cell immobilization. BSPs was
developed by Atkinson et al. [168] and used by many scientists and each
scientist provide an efficient way of immobilization on it giving out a high
yield of biodiesel. A study to check the lipolytic activities of Bacillus species
using intracellular as well as extracellular lipase showed higher intracellu-
lar lipase activity than extracellular lipase activity [169]. Reported whole-
cell biocatalysts are Aspergillus oryzae, Burkholderia cepacia, filamentous
fungus Rhizopus chinensis, R. oryzae, and Enterococcus faecium [170–174].
Aspergillus oryzae used as whole-cell biocatalyst exhibited 98.1% relative
24 Biodiesel Technology and Applications

Table 1.3 Some examples commercial lipases commonly used for biodiesel
production.
Reaction
Enzyme Substrate Acyl acceptor yield Reference
Novozyme Chlorella sp. Dimethyl carbonate 90% [145]
435 KR-1 and methanol
mixture
Sunflower oil Methanol, Absolute >90% [146]
ethanol,
1-propanol
Oleic acid Ethanol, n-propanol, >90% [147]
and n-butanol
Crude Methanol 94% [148]
soybean oil
Soybean oil Ethyl acetate 63.3 % [149]
Soybean oil Ethanol [150]
Lipozyme Soybean oil Methanol >90% [151]
TL IM
Crude palm Methanol 96.15% [90]
oil
Waste Methanol 92.8% [152]
cooking oil
Palm oil Oleyl alcohol 79.54% [153]
Corn oil Methanol 92% [154]
Lipozyme Crude Monoacylglycerol 90% [155]
RM IM rapeseed
oil
Sunflower oil Methanol >80% [82]
Castor oil Ethanol 98% [3]
Soybean oil Ethanol >88% [156]
deodorizer
distillate
Biocatalytic Processes 25

stability after the fourth batch and produced more than 97% FAME in 32
hours. Extracellular lipases are the purified form mainly fungal and bacte-
rial cells for their use in transesterification process.
Extracellular lipases are separated from broth containing lipase produc-
ing cells and after purification used as a catalyst in biodiesel production
processes [55]. The way to purify extracellular lipases depends upon its
structure and source organism [80]. Mostly extracellular lipases are used
in the immobilized form for transesterification than as free lipases because
of the low conversion rate and costly process [11]. Literature is full of
different methods as well as materials used for immobilization of extra-
cellular lipases. Main methods for immobilization involve cross linking,
carrier binding and entrapment while the most commonly used materi-
als for immobilization include silica, magnetic particles, and nanofibers
or nanoparticles for carrier binding, alginate beads, gels, and silicon poly-
mers for entrapment and glutaraldehyde for cross-linking [47]. The use
of a suitable solvent in case of extracellular lipase is a key factor for high
yield in transesterification as the use of unrelated solvent or absence of
solvent results in very low yield [102, 165]. The use of extracellular lipase
is also adapted because the use of intracellular lipase results in difficul-
ties of extraction and purification of the final product [171]. Extracellular
lipases are obtained from Candida guilliermondii, Burkholderia glumae,
Pseudomonas aeruginosa, and Yarrowia lipolytica [176–179]. Table 1.4
indicated a comparison between intracellular and extracellular lipases.

Table 1.4 Comparison between intracellular and extracellular lipase.


Intracellular lipase Extracellular lipase
Present inside the cell or linked to its Separated from cells producing it
walls (cell bound lipase)
No need of isolation and purification Complex isolation and purification
steps are required before using it for
biodiesel production
Low conversion rate High conversion rate
Not analyzed by direct sampling Analyzed by direct sampling
Direct immobilization of lipase Purification is required before
producing cells (whole-cell immobilization
immobilization
Biodiesel production is cost-effective Biodiesel production is costly
26 Biodiesel Technology and Applications

1.9 Recombinant Lipases for Cost-Effective Biodiesel


Production
In order to produce economically feasible and cost-effective enzymatic
biodiesel production, various methods have been used. For example,
improvement in purification procedures, use of better bioreactors, var-
ious immobilization techniques, use of easy to handle solvents and solu-
tion, finding novel organisms that can produce more stable and effective
lipases. These genetic manipulations enable recombinant organism
to overexpress active lipases. A lot of work has been done to produce
recombinant lipases from recombinant organisms by deriving desired
gene from another specie or organism to meet our demands. According
to Huang et al. [173], a good conversion of microalgal oil into FAME and
FAEE was obtained (<90%) when recombinant lipase from Rhizomucor
miehei was expressed in Pichia pastoris, in the presence of methanol and
ethanol as acyl acceptor, respectively. That recombinant lipase enzyme
was called Lipase GH2 and it was used as free enzyme for catalysis.
Aspergillus oryzea has been used majorly for lipase expression as manip-
ulating organism. Adachi et al. [174] created thermostable and solvent
tolerant whole-cell biocatalysts that worked as robust biocatalyst for bio-
diesel production. Geobacilus thermocatenulatus was used as a source
organism from which thermostable lipase (BTL2) gene was obtained and
then introduced into Aspergillus oryzea (r-BTL). Immobilized whole-
cell r-BTL was then used to convert palm oil into FAME production and
the resultant yield obtained was 100%. Whole-cell biocatalyst tolerated
40-50°C temperature and 30% (v/v) solvents such as methanol, DMC, and
acetone. These results proved the thermostability and solvent tolerance
of Aspergillus oryzea (r-BTL). Lipase from B68 strain of Pseudomonas
fluorescence (lipB68) was transformed into E. coli BL21 which resulted in
92% yield upon transesterification. The amazing thing about this exper-
iment was use of that novel psychrophilic lipase (lipB68) that catalyzed
reaction at 20°C and energy consumption became reduced in the reaction
system [175]. So, use of psychrophilic lipase for biodiesel production can
become more useful in the future. In many research experiments, recom-
binant lipases have performed better than commercially used lipases. For
example, according to Yan et al. [176], recombinant Pichia pastoris strain
was developed that used to overexpress Thermomyces anuginosus lipase.
When this strain was used as a whole-cell biocatalyst, it efficiently con-
verted waste cooking oil into FAME with 82% yield within 84 h. After
comparing this recombinant strain performance with commercial lipase
Biocatalytic Processes 27

Lipozyme TLIM, it was found that the earlier one outperformed the
later one. In another research, lipase LIP2 coding gene was cloned from
Candida rugosa and expressed in Pichia pestoris after changing some
codons of that gene by side directed mutagenesis to make it more useful
and active. So, then that recombinant Candida rugose gave better results
when results of commercial C.rugosa lipase and recombinant LIP4 were
compared because recombinant LIP2 lipase was tolerating temperature
30°C–50°C and work even at 70°C [177]. A lot of research has been done
in utilizing recombinant lipases and found better results. Tables 1.5 and
1.6 showed some examples of recombinant whole-cell lipases with and
without immobilization.

Table 1.5 Some research investigations using recombinant whole-cell biocatalyst


after immobilization.
Expressing lipase Acyl
source Immobilized on Substrate acceptor Yield Reference
Fusarium Biomass support Aqueous Methanol 96.1% [178]
heterosporum particles plant oil
emulsion

Soybean oil Methanol 98% [179]

Thermomyces Protein-coated Palm olein Ethanol 89.9% [180]


lanuginosus microcrystals
Crude palm 82.1%
oil

Palm fatty 75.5%


acid
distillate

Geobacillus Reticulated Palm oil Methanol ~100% [174]


thermocatenulatus polyurethane
(BTL2) foam BSPs

Proteus mirabilis hydrophobic Canola oil Methanol 76% [181]


Immobead
350 oxirane
functionalized
beads

Staphylococcus poly Olive oil Methanol 90% [182]


haemolyticus L62 (methacrylate-
co-divinyl
benzene) resin
28 Biodiesel Technology and Applications

1.10 Immobilization of Lipases for Better Biodiesel


Production
Immobilization is the physical binding of enzymes to a solid support in
such a way that when the substrate passes over the enzyme support it
converted into their product [189]. The vast use of Enzyme immobiliza-
tion technique at industrial scale is because it results in enzyme stabiliza-
tion and prolonged enzyme–substrate contact which results in reduction
of purification steps. Further, it lowers the reaction cost by making the
reaction contamination free and enabling recycling of the enzyme [190].
As the cost of industrial production of lipase enzyme is one of the main
obstacle in its industrial application, so it can be eliminated through
lipase immobilization [191]. Industrial processes/reactions can be made
continuous with little protein using immobilized lipases because of their
easy separation from reaction system containing products, by-products,
residual substrates, and medium [192]. In transesterifications is the

Table 1.6 Some other experiments using recombinant whole catalyst without
immobilization.
Expressing lipase
source Substrate Acyl acceptor Yield Reference
Serratia marcescens Waste grease Methanol 97% [183]
YXJ-1002
Rhizomucor miehei Soybean oil Methanol >95% [184]
and Penicillium
cyclopium
Fusarium Soybean oil Methanol 95% [185]
heterosporum
Rhizomucor miehei Soybean oil Methanol 83.14% [186]
(RML)
Microalgal Methanol >90% [187]
oil
Thermomyces Waste Methanol 82% [176]
lanuginosus (Tll) cooking
oil
Rhizopus oryzae Soybean oil Methanol 71% [188]
IFO4697
Biocatalytic Processes 29

main process involves in biodiesel production using lipase enzyme, the


use of immobilized enzyme (extracellular) and immobilized whole-cell
(intracellular enzyme) both reported highly efficient as compared to the
use of free enzymes. Bayramoglu et al. [59] set two reaction setups, one
with free enzyme and other using immobilized enzyme, and then reac-
tion comparison was measured using gas chromatography-mass spec-
trophotometry (GC-MS) by noticing FAME components. They Candida
rugosa lipase was immobilized on microporous bio-silica polymer. Both
enzymes then further used for transesterification of Scenedesmus quadri-
cauda microalgal oil using methanol in the presence of n-hexane solvent.
Reaction setup having immobilized lipase produced 96.4% yield as com-
pared to free lipase that produced 85.7% yield. Moreover, immobilized
lipase was very much stable as it just lost 17% of its activity after six
cycles. Kalantari et al. [193] used Pseudomonas cepacia lipase, immobi-
lized on mixture of mesoporous and nanoporous silica coated composite
particles, for transesterification of soybean oil into FAME using metha-
nol. Immobilization effect retains 55% enzymatic stability after five con-
secutive cycles compared to free lipase reaction. Shah et al. [194] showed
that three enzymes were selected initially and then Chromobacterium
viscosum was further chosen because of its better results. Celite 545 was
used as immobilization material applied to enzyme. Transesterification of
Jatropha oil using ethanol was conducted with and without immobilized
lipase. After immobilization, reaction yield increased from 62% to 71%
and then after further addition of water and immobilization, it reached
to 92%. Moreover, in another study [85], triolein transesterification was
conducted using methanol, ethanol, and immobilized Pseudomonas flu-
orescens in the presence of 1,4-dioxane solvent. Reaction yield for both
free and immobilized lipase reaction setups was same, i.e., 90%. But the
stability provided by immobilization, loss of lipase activity was mini-
mal in immobilized reaction setup than other. Reaction time was 10h
for immobilized as compared to 25h reaction time in free lipase system.
Further, when we use water free media, free enzyme molecules tend to
clump together and decreasing the surface area. Enzyme immobilization
inhibit clump formation and increase the surface area of the biocatalyst
[195]. Kumari et al. [97] carried out transesterification of jatropha oil
with Enterobacter aerogenes lipase immobilized on surface and reported
high yield (94%) of biodiesel production and noticed that lipase activity
reduced in little extent even after repeated use. Enzyme properties like
resistance to proteolytic digestion and denaturants, temperature pro-
file, pH-dependence, thermostability, and kinetics are mainly affected
by enzyme immobilization. The chief issues for enzyme immobilization
30 Biodiesel Technology and Applications

are preference for the selection of immobilization techniques and sup-


port matrices that permit both rapid enzyme activity and enzyme sta-
bility under the constraints imposed by the substrate medium [89, 196].
Some examples of different investigations regarding biodiesel produc-
tion methods using whole-cell immobilized biocatalysts are given in
Table 1.7.

Table 1.7 Investigation of immobilized whole-cell biocatalysts to produce


biodiesel.
Lipase cell Immobilized Acyl
factory on Substrate acceptor Yield Reference
Rhizopus Biomass Soybean oil Methanol 70%– [197]
oryzae cells support 83%
particles
Mucor Poly-Urethane Sardine Methanol NA [198]
circinelloides Foam (Sardinella
lemuru) oil
Mucor polyurethane Babassu oil Ethanol 98.1% [199]
circinelloides support
URM 4182
Rhizopus biomass Oleic acid Methanol 90% [200]
oryzae support
IFO4697 particles Soybean oil Methanol 72%

Rhizopus biomass Jatropha Methanol 90% [201]


oryzae support curcas oil
(ROL) particles
Soybean oil Methanol 90% [202]
Rhizopus Calcium waste Methanol 84% [203]
oryzae 262 alginate cooking oil
beads (sunflower
oil)
Rhizopus biomass Cottonseed Methanol 27.9% [204]
oryzae support oil
ATCC particles
34612
Pseudomonas Sodium Jatropha oil Methanol 72% [205]
fluorescens alginate
MTCC 103
Aspergillus Biomass Palm oil Methanol >90% [104]
niger support
particles
Biocatalytic Processes 31

1.11 Recent Strategies to Improve Biodiesel


Production
1.11.1 Combination of Lipases
Lipases from different sources or organisms possess different properties.
Just like no ideal gas really exists, similarly, there is no ideal lipase that
will be considered as absolutely perfect. Just one lipase cannot have all
properties that are ideally required for optimal biodiesel production. But
researchers are trying to find the best among the lot. Novel sources pro-
vide different features of lipase that can be good for biodiesel production.
Some lipases are good at tolerating high temperatures, some may work at
low temperature, some provide a good temperature range, some may have
good catalytic activity, and some may have good hydrolytic activity. Raw
material or substrate that is used for biodiesel production is very diverse in
nature. Feedstock is mainly comprised of FFA, triglycerides, diglycerides,
or monoglycerides, so it is difficult for one lipase to tackle all these things
with best possibility. Likewise, every lipase has different alcohol inhibition
capacity to tolerate it. So, all in all, these things make it very difficult for one
lipase to be perfectly suitable to give best results. So, researchers have come
with an idea to use combination of enzymes/lipases. One may be best at
providing specific feature and other may be good at giving another specific
feature and so on. This way combination of enzymes will provide many
features that will be good for biodiesel production. A lot of combination of
enzymes have been studied recently for biodiesel production. According
to Li et al. [206], combined use of Lipozyme TL IM and Novozym 435
lipase resulted in 96.38% conversion of stilingia oil into FAME. Reaction
conditions were methanol-to-oil molar ratio (6.1:4), temperature 20°C
along with the use of co-solvent consisting of acetonitrile and t-butanol.
At these conditions even after 30 cycles, there was no loss in lipase activity.
Similarly, many investigations have been done and some of them are given
in Table 1.8.
Another way of using compound lipases is cloning and expression of
two or more different lipase coding genes from different organism in a host
organism. But cloning and effective expression of lipases from host organism
is a difficult and cumbersome process, but it has also provided better results.
For example, according to Guan et al. [184], Pichia pestoris was selected as
a host organism to express two lipase coding genes cloned from two dif-
ferent organisms, one of them was R. miehei (source of 1,3-­specific lipase)
and the other was Penicillium cyclopium (source of non-specific lipase).
32 Biodiesel Technology and Applications

Table 1.8 Some examples to produce biodiesel using combination of lipases.


Combination of Acyl
lipases Immobilized on Substrate acceptor Yield Reference
Thermomyces Lewatit VP OC 1600 Soybean oil Ethanol 90% [207]
lanuginosus
lipase and TLL immobilized on Palm oil Ethanol 80% [208]
Rhizomucor acrylic resin, RML
miehei lipase. immobilized on
anion-exchange
resin
Pseudomonas Accurel PE-100 Used palm oil Ethanol >67% [209]
fluorescens microporous
and Candida polypropylene
rugosa powder
Immobilized macroporous Soybean oil Methanol 98% [210]
Pseudomonas polypropylene
fluorescens
lipase
followed by
immobilized
Pseudomonas
cepacia lipase
Combined use Lipozyme TL IM waste cooking Methanol 99% [211]
of Lipozyme immobilized on sunflower
TL IM and acrylic resin and oil
Novozym 435 Novozym 435
immobilized on
microporous resin
Novozym 435 RML immobilized on Soybean oil Ethanol 80% [212]
(CALB) and an anion-exchange
Canola oil Methanol >95% [213]
Lipozyme resin, and CALB
RM-IM immobilized on a
(RML) macroporous resin
Combination microporous Triolein Ethanol 96% [214]
of Rhizopus polypropylene
arrhizus lipase Accurel MP1000
and Candida
antarctica
lipase B
Candida rugosa Macroporous acrylic Acid oil, Methanol 91% [215]
lipase resin product of
followed by vegetable
Novozym 435 oil
Biocatalytic Processes 33

Extract containing these two lipases when applied for biodiesel produc-
tion to transesterify soybean oil at 30°C with 4:1 alcohol-to-oil molar ratio
resulted in 99.7% yield after 24 h. In another research, a special recom-
binant Aspergillus oryzae whole-cell biocatalyst was created that used to
co-express two different lipases genes, one of these two lipases was derived
from Fusarium heterosporum (FHL) and the other one was mono and dia-
cyl glycerol lipase B. Use of that whole-cell recombinant Aspergillus ory-
zae biocatalyst resulted in 98% methyl ester yield [179]. According to Yan
et al. [216], recombinant Pichia pestoris was developed that displayed two
lipases, i.e., T. lanuginosus lipase (TLL) and C. antarctica lipase B (CALB)
from different sources on its surface. This whole-cell biocatalyst co-express-
ing both lipases produced 95.4% biodiesel yield under optimum conditions.
Apart from ILs advantages, there is a problem with the biodiesel recovery
because during continuous biodiesel removal reaction moves backward
and affects the resultant yield. So, to avoid this problem SC-CO2 has been
suggested along with ILs. SC-CO2 is very effective in recovering biodiesel
because ester molecules have good solubility in it. IL-SC-CO2 combina-
tion not only provides easy recovery but also prevent glycerol inhibition
effect. SC-CO2 saturated with (substrate) oil is introduced into the reaction
system and this creates two phases because of immiscibility in each other.
SC-CO2 can diffuse through IL (ionic liquid) phase bringing substrate with
it, reaches the enzyme active site and makes the reaction easily possible.
After enzyme activity when biodiesel is formed, biodiesel esters become
soluble in SC-CO2 phase, So, in this way, biodiesel becomes separate from
ILs. Glycerol (by-product of the reaction) does not dissolve into SC-CO2 so
it makes another separate layer and then glycerol can be easily taken out in
pure form. SC-CO2 containing biodiesel is then processed to recover bio-
diesel from it by depressurization [217, 218]. In this way, two phase system
due to IL-SC-CO2 combination enable to recover biodiesel in good qual-
ity. IL-SC-CO2 combination system was first used to extract naphthalene
using [bmim] [PF6] as IL [219]. Enzymatic biodiesel was produced with
IL-SC-CO2 system by transesterification of triolein using Novozyme 435 as
biocatalyst and methanol as acyl acceptor. In IL-SC-CO2 system, 12 differ-
ent ILs were used and there were two different temperature conditions, i.e.,
60°C and 85°C. After 6 h, the resultant biodiesel yield obtained was more
than 98% that made this reaction a successful one [220].

1.11.2 Microwave and Ultrasonic-Assisted Reaction


Microwave employment in the chemical or biochemical reaction is very
clean, economic, fast and energy efficient method. Use of microwave is
34 Biodiesel Technology and Applications

very helpful in substrate preparation, extraction, and biodiesel production.


Microwave irradiation assists the reaction by providing direct energy to
the reactants, and hence, the loss of energy becomes minimal. This fea-
ture makes it economic than using other conventional methods of heating
to make the reaction possible [221]. Easy energy transfers and use makes
the enzyme to perform better and its easy separation leads to cost reduc-
tion. So, this environment friendly method provides energy efficient, faster
and cost-effective reaction. There have some research investigations been
done to analyze how effective it would be to use microwave. Macauba oil
transesterification was performed using Novozyme 435 and Lipozyme IM
in the presence of ethanol as acyl acceptor. Activities of these enzymes
were checked and compared in the presence and absence of microwave.
Before microwave assistance enzymatic activity values of Novozyme 435
and Lipozyme IM were 0.09 and 0.08, respectively. But after microwave
presence activity value were 0.6 and 0.4, respectively. Results were clear
that microwave assistance enhanced enzymatic activity about one order of
magnitude [222, 223]. In another experiment, soybean oil and sunflower
oil was transesterified using Novozyme 435 and ethanol to oil molar ratio
was 3:1. Resultant ester yield was more than 99% within just 3 h when reac-
tion system was accompanied with microwave while conventional heating
produced yield about 57% with the same reactions conditions as provided
in case of microwave reaction system for biodiesel production. So, micro-
wave-assisted reaction produced better and faster biodiesel yield compared
to conventional heating reaction system [224, 225].
According to Zhang et al. [224], microwave and deep eutectic solvents
assisted transesterification of yellow horn seed oil using Novozyme 435
resulted in 95% conversion yield. Ultrasonication is another advancement
in effective biodiesel production. Just like microwave, ultrasonic assistance
increases the reaction rate and ultimately makes the transesterification
cost effective [225]. Increase in reaction rate is linked with enhancement
in mass transfer and good energy input. Ultrasonication also increases the
surface area for the reactant to work efficiently. Moreover, it also boosts the
enzymatic catalytic activity and efficiency which provides optimum results
even when catalyst is in low quantity [226, 227]. Sometimes, ultrasoni-
cation when combined with other method it produces better results and
effective rapid transesterification. According to Yu et al. [227], transester-
ification of soybean oil in the presence of methanol and Novozyme 435 as
biocatalyst produced good results when ultrasonic irradiation applied on
the reactants. When ultrasonic irradiation with vibration (UIV) applied
then conversion of soybean oil into FAME produced 96% yield in just
4 h and Novozyme 435 did not lose any activity after five cycles. But in
Biocatalytic Processes 35

the absence of UIV, comparable results were obtained after 12 h. Results


proved that presence of UIV in the enzymatic transesterification reaction
gave faster and better yields with no loss in enzyme activity. In another
research, Novozyme 435 was utilized to catalyze transesterification of
waste cooking oil in an ultrasound-assisted environment. The resultant
biodiesel yield was 75.19% in just 2 h. It was also found that ultrasound
along with little agitation/vibration made the reaction more efficient [228].
Just like microwave, ultrasonication is also taking attention of researchers
to produce better and cost-efficient biodiesel.
Researchers have produced some good results in the past couple of years
using ultrasonication method. Ultrasound-assisted transesterification of
soybean oil using ethanol and Novozyme 435 resulted in ~78 % yield in
just 1 h. this method of assistance has the potential to become alternative to
the alkali catalyzed or conventional enzymatic biodiesel production [229].
In another experiment, sunflower oil and methanol were used for biodiesel
production in the presence of Lipozyme TL-IM under ultrasound-assisted
system. Results indicated that ultrasound assistance suppresses the much
needed requirement of using methanol, making the reaction fast and
clean [230]. Conventional transesterification using bath process technique
requires a lot of time because phase separation and recovery of glycerol
and biodiesel is a time-consuming process. Use of ultrasound and micro-
wave is very helpful in this regard and it has been proved that ultrasound
and microwave assistance make the reaction rapid and cost efficient.

1.12 Lipase Catalyzed Reaction Modeling


and Statistical Approaches for Reaction
Optimization
A lot of methods have been tried for the modelling of lipase catalyzed
reactions to fully understand reaction mechanism. Need of better catalyst
with great substrate specificity, immobilization, whole-cell catalyst, and
manipulation of lipase, all these methods have been applied with various
techniques to get better results [231]. But all these methods have been
used by hit and trial method to check feasibility. Apart from dealing with
lipases, there are many more parameters to set optimally to achieve max-
imum output. These parameters include temperature, pH, water content,
alcohol-to-oil molar ratio, and bioreactor choice. All reaction conditions
using various combination of parameters with diverse conditions have
been used but it is not actually we need because hit and trial method
leads to expensive experimentation. In this era of science and technology,
36 Biodiesel Technology and Applications

in silico experimentation is in great demand and it can apply to each and


every field. Optimization of biochemical processes is one of key things to
produce cost effective results. In order to optimize a reaction, there are
various parameters that are dealt with but normally optimization accounts
for changing just one parameter at a time by keeping others constant and
similarly one by one all parameters are optimized. This technique is called
one variable at a time technique. The major limitation of this technique
is that it does not show interactive results of parameters. So, it does not
produce results based on all parameters effects at the same time on the
process.
Process modeling, simulation, and statistical approaches are the best
way to tackle this problem. Modeling and simulation of the processes are
the base of in silico experimentation that provides similar environmental
conditions of the processes just like real system without so much invest-
ment. This not only provides process mechanism information but also
gives support for process development. Reaction optimization is difficult
to achieve and it requires a lot of efforts and financial support. While opti-
mization, high efficiency of one of many parameters sometimes reduces
the demand of some other parameter and reaction product may not alter
so much. For example, according to Tufvesson et al. [232], if enzyme has
excellent catalytic activity then this minimizes the need of enzyme specific
activity and similarly, if enzyme catalytic activity is fair enough to product
healthy amount of product then ultimately need of efficient downstream
processes will be reduced. So, all in all optimization of chemical processes
is difficult and tedious to achieve. That is why in silico experimentation is
helpful in process of optimization and handling.
Modeling and simulation of any process gives idea to the researchers
what they actually need regarding system conditions, type of catalyst with
respect to their mechanism kinetics and energetics from ping pong bi bi
mechanism, type of substrate, and various parameters setup. This greatly
prevents hit and trial approach and makes the process cost effective. There
are two basic approaches in process modeling: (a) empirical approach deal
with design of experiments and (b) statistical approach that gives relation
between process outcome and parameters [232]. Parameter estimation is
a key thing to avoid false outcome in process modeling and more chances
of error in estimation leads to wrong direction. Yu et al. [231] observed
that, optimal experimental design was developed for the improvement of
process estimation in biodiesel production process catalyzed by Novozym.
Simulation demonstrated the reduction in producing errors while estimat-
ing parameters through optimal experimental design. Response surface
methodology (RSM) is a collection of various mathematical and statistical
Biocatalytic Processes 37

techniques that is used when there are multiple variables influencing the
reaction response at the same time. RSM is used for improvement, devel-
opment, and especially optimization of chemical and biochemical reac-
tions. RSM creates a mathematical model that can be used to predict what
can be the response of the process under certain variable conditions. That
is why it is called RSM.
A lot of research has been done using RSM to optimized chemical
reactions. For example, protease production in a bioreactor taken from
Bacillus mojavensis and statistically optimization of media [233] and
lipase catalyzed esterification reactions [234]. RSM is widely used for
optimization but it can also be used to determine kinetic constants and
enzyme stability. Kinetic constants for protease derived from Bacillus
mojavensis was determined using RSM [235]. According to Rana et al.,
kinetic and stability of b-1,3-glucanase from Trichoderma harzianum
and alcohol dehydrogenase was investigated using RSM [236]. RSM for
optimization is carried out in 3 stages, (a) independent variable and their
level is determined, (b) experimental design is selected then model equa-
tion is predicted and verified, and (c) different response plots are obtained
and optimum points are determined [237]. The most common and suc-
cessfully implemented design used in RSM is central composite rotatable
design (CCRD) for the optimization of reactions [238, 239]. Like every
other thing, RSM surely has advantages as well as disadvantages (Table
1.8). Like other chemical processes, optimization of enzymatic catalyzed
biodiesel production process has been widely investigated using RSM
statistical approach [241].
According to Sheih et al. [241], conversion of soybean oil into biodiesel
was catalyzed using lipase from Rhizomucor miehei (Lipozyme IM-77) in
the presence of methanol which was being investigated in this experiment.
RSM was implemented with having 5-level-5-factor CCRD. Through this
design and RSM, effect of time, methanol-to-oil molar ratio, temperature,
water content, etc., were measured and evaluated. Under optimal condi-
tions that were temperature of 36.5°C, reaction time of 6 h, water content
of 5.8%, and substrate molar ratio of 3.4:1, resulted in 92.2% conversion
yield when analysis was done using ridge max analysis. In another experi-
ment, optimization was done in pretreatment step to lower the concentra-
tion of FFA of Mahua oil that would halt the downstream processes and
ultimately biodiesel production. CCRD was again used to check the effects
of acid and methanol concentration and reaction time in reducing the pro-
duction of FFA. After pretreatment, Mahua oil was catalyzed through alka-
line catalyst, and then, the resultant biodiesel produced was according to
American and European standards [242].
38 Biodiesel Technology and Applications

1.13 Conclusion and Summary


Currently, the use of fossil fuels is on its verge because of the dependency
of our transportation and main industries on it. If its usage goes with even
at the same pace, there are great chances of its shortage due to depletion in
reservoirs in very near future. So, there is a strict need to divert toward alter-
nate sources like biofuels which not only overcome the shortage of fuel but
also environment-friendly. Biodiesel is the most widely used biofuel due to
its vast use from direct fuel to a lubricating agent. The production and use
of biodiesel are restricted to limited countries but there is a need to broaden
its use worldwide. The barriers which constrain its worldwide use are mainly
related to biodiesel production cost which can be reduced with optimization
of the production process by selection of preferable feedstock, reaction con-
ditions, type of catalyst used, enzymes for catalysis, and many other factors.
Researchers are making their full effort to construct a cost-effective bio-
diesel production process. Oils extracted from various sources are used as a
substrate or feedstock for biodiesel production, but microbial oil extracted
from microalgae is used at first place due to low cost, easy, and quick pro-
duction when compared with plant and animal oils. However, biodiesel
produced using animal oil act as a good lubricating agent along with good
oxidative stability. Biocatalysis involving enzymatic biodiesel production
via lipase enzyme is considered a low-price process as compared to other
ways. Lipase enzyme is preferably used as biocatalysis because of its easy
separation and reusability.
There are many factors like acyl receptors, water, and temperature that
have a serious impact on enzymatic transesterification and great research
have been done for to optimization of these factors to make the process
easy and affordable. Lipases from a broad range of sources immobilized
by different techniques on a variety of supports have their own resulting
impact in biodiesel production process. However, the use of whole-cell
immobilization or combination of lipase with optimized reaction condi-
tions like temperature, time of reaction, and pH makes the process indus-
trially favorable. Still, there exists a space for improvement to make the
biodiesel production process affordable and easy to conduct.

References
1. M. Naik, L.C. Meher, S.N. Naik, L.M. Das, Production of biodiesel from high
free fatty acid Karanja (Pongamia pinnata) oil, Biomass and Bioenergy. 32,
354–357, 2008. https://doi.org/10.1016/j.biombioe.2007.10.006.
Biocatalytic Processes 39

2. W. Parawira, Biodiesel production from Jatropha curcas: A review, Sci. Res.


Essays. 5, 1796–1808, 2010. https://doi.org/10.3109/07388551.2010.487185.
3. Oliveira, M. Di Luccio, C. Faccio, C.D. Rosa, J.P. Bender, N. Lipke, S. Menoncin,
C. Amroginski, J.V. de Oliveira, Optimization of enzymatic production of bio-
diesel from castor oil in organic solvent medium., Appl. Biochem. Biotechnol.
113–116, 771–780, 2004. https://doi.org/10.1385/ABAB:115:1-3:0771.
4. L. Azócar, R. Navia, L. Beroiz, D. Jeison, G. Ciudad, Enzymatic biodiesel pro-
duction kinetics using co-solvent and an anhydrous medium: A strategy to
improve lipase performance in a semi-continuous reactor, N. Biotechnol. 31,
422–429, 2014. https://doi.org/10.1016/j.nbt.2014.04.006.
5. L. Azócar, G. Ciudad, H.J. Heipieper, R. Muñoz, R. Navia, Improving fatty
acid methyl ester production yield in a lipase-catalyzed process using waste
frying oils as feedstock, J. Biosci. Bioeng. 109, 609–614, 2010. https://doi.
org/10.1016/j.jbiosc.2009.12.001.
6. K.R. Jegannathan, L. Jun-Yee, E.S. Chan, P. Ravindra, Production of biodiesel
from palm oil using liquid core lipase encapsulated in κ-carrageenan, Fuel.
89, 2272–2277, 2010. https://doi.org/10.1016/j.fuel.2010.03.016.
7. X. Meng, G. Chen, Y. Wang, Biodiesel production from waste cooking oil via
alkali catalyst and its engine test, Fuel Process. Technol. 89, 851–857, 2008.
https://doi.org/10.1016/j.fuproc.2008.02.006.
8. L.P. Christopher, Hemanathan Kumar, V.P. Zambare, Enzymatic biodiesel:
Challenges and opportunities, Appl. Energy. 119, 497–520 2014. https://doi.
org/10.1016/j.apenergy.2014.01.017.
9. B.R. Moser, Biodiesel production, properties, and feedstocks, Vitr. Cell. Dev.
Biol. - Plant. 45, 229–266 2009.
10. Z. Amini, Z. Ilham, H.C. Ong, H. Mazaheri, W.H. Chen, State of the art
and prospective of lipase-catalyzed transesterification reaction for bio-
diesel production, Energy Convers. Manag. 141, 339–353, 2016. https://doi.
org/10.1016/j.enconman.2016.09.049.
11. A. Guldhe, B. Singh, T. Mutanda, K. Permaul, F. Bux, Advances in synthesis
of biodiesel via enzyme catalysis: Novel and sustainable approaches, Renew.
Sustain. Energy Rev. 2015. https://doi.org/10.1016/j.rser.2014.09.035.
12. H. Fukuda, A. Kondo, H. Noda, Biodiesel fuel production by transesterifi-
cation of oils, J. Biosci. Bioeng. 92, 405–416, 2001. https://doi.org/10.1016/
S1389-1723(01)80288-7.
13. C.C. Akoh, S. Chang, G. Lee, J. Shaw, Enzymatic approach to biodiesel pro-
duction, J. Agric. Food Chem. 55, 8995–9005, 2007. https://doi.org/10.1021/
jf071724y.
14. A.P.D.S.S.A.M.W.Y.M. Haas, J.A.D.H. Shapouri, Energy Life-Cycle
Assessment of Soybean Biodiesel, 2009.
15. A. Demirbas, Progress and recent trends in biodiesel fuels, Energy Convers.
Manag. 50, 14–34, 2009.
16. S. Al-Zuhair, Production of biodiesel: possibilities and challenges, Biofuels,
Bioprod. Biorefining. 1, 57–66, 2007. https://doi.org/10.1002/bbb.2.
40 Biodiesel Technology and Applications

17. L.J. Konwar, J. Boro, D. Deka, Review on latest developments in biodiesel


production using carbon-based catalysts, Renew. Sustain. Energy Rev. 29,
546–564, 2014. https://doi.org/10.1016/j.rser.2013.09.003.
18. A. Demirbas, Biodegradable plastics from renewable resources, Energy
Sources, Part A Recover. Util. Environ. Eff. 29, 419–424, 2007. https://doi.
org/10.1080/009083190965820.
19. M. Lapuerta, J.M. Herreros, L.L. Lyons, R. García-Contreras, Y. Briceño,
Effect of the alcohol type used in the production of waste cooking oil bio-
diesel on diesel performance and emissions, Fuel. 87, 3161–3169, 2008.
https://doi.org/10.1016/j.fuel.2008.05.013.
20. R.O. Dunn, Alternative jet fuels from vegetable oils, Trans. ASAE. 44, 1751–
1757, 2001. https://doi.org/10.13031/2013.6988.
21. N. Hashimoto, Y. Ozawa, N. Mori, I. Yuri, T. Hisamatsu, Fundamental
combustion characteristics of palm methyl ester (PME) as alternative
fuel for gas turbines, Fuel. 87, 3373–3378, 2008. https://doi.org/10.1016/j.
fuel.2008.06.005.
22. Y.C. Lin, C.H. Tsai, C.R. Yang, C.H.J. Wu, T.Y. Wu, G.P. Chang-Chien,
Effects on aerosol size distribution of polycyclic aromatic hydrocarbons
from the heavy-duty diesel generator fueled with feedstock palm-biodiesel
blends, Atmos. Environ. 42, 6679–6688, 2008. https://doi.org/10.1016/j.
atmosenv.2008.04.018.
23. M. Aarthy, P. Saravanan, M.K. Gowthaman, C. Rose, N.R. Kamini, Enzymatic
transesterification for production of biodiesel using yeast lipases: An over-
view, Chem. Eng. Res. Des. 92, 1591–1601, 2014. https://doi.org/10.1016/j.
cherd.2014.04.008.
24. J.C. Juan, D.A. Kartika, T.Y. Wu, T.Y.Y. Hin, Biodiesel production from jatro-
pha oil by catalytic and non-catalytic approaches: An overview, Bioresour.
Technol. 102, 452–460, 2011. https://doi.org/10.1016/j.biortech.2010.09.093.
25. Z. Yaakob, M. Mohammad, M. Alherbawi, Z. Alam, K. Sopian, Overview of
the production of biodiesel from waste cooking oil, Renew. Sustain. Energy
Rev. 18, 184–193, 2013. https://doi.org/10.1016/j.rser.2012.10.016.
26. P. Nautiyal, K.A. Subramanian, M.G. Dastidar, Production and characteriza-
tion of biodiesel from algae, Fuel Process. Technol. 120, 79–88, 2014. https://
doi.org/10.1016/j.fuproc.2013.12.003.
27. A. Karmakar, S. Karmakar, S. Mukherjee, Properties of various plants and
animals feedstocks for biodiesel production, Bioresour. Technol. 101, 7201–
7210, 2010. https://doi.org/10.1016/j.biortech.2010.04.079.
28. G. Knothe, Dependence of biodiesel fuel properties on the structure of fatty
acid alkyl esters, Fuel Process. Technol. 86, 1059–1070, 2005. https://doi.
org/10.1016/j.fuproc.2004.11.002.
29. D.Y.C. Leung, X. Wu, M.K.H. Leung, A review on biodiesel production using
catalyzed transesterification, Appl. Energy. 87, 1083–1095, 2010. https://doi.
org/10.1016/j.apenergy.2009.10.006.
Biocatalytic Processes 41

30. M. Canakci, H. Sanli, Biodiesel production from various feedstocks and their
effects on the fuel properties, J. Ind. Microbiol. Biotechnol. 35, 431–441, 2008.
https://doi.org/10.1007/s10295-008-0337-6.
31. M.M. Gui, K.T. Lee, S. Bhatia, Feasibility of edible oil vs. non-edible oil vs.
waste edible oil as biodiesel feedstock, Energy. 33, 1646–1653, 2008. https://
doi.org/10.1016/j.energy.2008.06.002.
32. K. Jawaharraj, R. Karpagam, B. Ashokkumar, S. Kathiresan, P. Varalakshmi,
Green renewable energy production from Myxosarcina sp.: Media optimiza-
tion and assessment of biodiesel fuel properties, RSC Adv. 5, 51149–51157,
2015. https://doi.org/10.1039/C5RA09372D.
33. Q. Li, W. Du, D. Liu, Perspectives of microbial oils for biodiesel produc-
tion, Appl. Microbiol. Biotechnol. 80, 749–756, 2008. https://doi.org/10.1007/
s00253-008-1625-9.
34. C. Yusuf, Biodiesel from microalgae, Biotechnol. Adv. 25, 294–306, 2007.
35. J. Janaun, N. Ellis, Perspectives on biodiesel as a sustainable fuel, Renew.
Sustain. Energy Rev. 14, 1312–1320, 2010. https://doi.org/10.1016/j.rser.
2009.12.011.
36. P. Suarez, S. Meneghetti, Transformation of triglycerides into fulels, polymers
and chemicals: some applications of catalysis in oleochemistry, Quim. Nova.
30, 667–676, 2007. https://doi.org/10.1590/S0100-40422007000300028.
37. M. Leca, L. Tcacenco, M. Micutz, T. Staicu, Optimization of biodiesel pro-
duction by transesterification of vegetable oils using lipases, 2010.
38. G. Kafuku, K.T. Lee, M. Mbarawa, Non-Catalytic and catalytic transesterifi-
cation: A reaction kinetics comparison study, Int. J. Green Energy. 12, 551–
558, 2015. https://doi.org/10.1080/15435075.2013.834820.
39. A.F. Lee, J.A. Bennett, J.C. Manayil, K. Wilson, Heterogeneous catalysis for
sustainable biodiesel production via esterification and transesterification,
Chem. Soc. Rev. 43, 7887–7916, 2014. https://doi.org/10.1039/C4CS00189C.
40. M. Mouaimine, J.D. Andriano, P.T. Fidele, V. Pierre, M.S. Mohamed, Plant
latex lipase as biocatalysts for biodiesel production, African J. Biotechnol. 15,
1487–1502, 2016. https://doi.org/10.5897/AJB2015.14966.
41. S.A.-Z. Hanifa Taher, The use of alternative solvents in enzymatic biodiesel
production: a review, Biofuels, Bioprod. Bioref. 11, 168–194, 2016. https://doi.
org/10.1002/bbb.1727.
42. F.G. and Z. Fang, Biodiesel production with solid catalyst, Biodiesel–
Feedstocks and Processing Technologies, 2011.
43. G. Jin, T.J. Bierma, C.G. Hamaker, R. Rhykerd, L.A. Loftus, Producing bio-
diesel using whole-cell biocatalysts in separate hydrolysis and methanolysis
reactions., J. Environ. Sci. Health. A. Tox. Hazard. Subst. Environ. Eng. 43,
589–95, 2008. https://doi.org/10.1080/10934520801893576.
44. T. Tan, K. Nie, F. Wang, Production of biodiesel by immobilized Candida sp.
lipase at high water content, Appl. Biochem. Biotechnol. 128, 109–116, 2006.
https://doi.org/10.1385/ABAB:128:2:109.
42 Biodiesel Technology and Applications

45. L. Fjerbaek, K. V. Christensen, B. Norddahl, A review of the current state of


biodiesel production using enzymatic transesterification, Biotechnol. Bioeng.
2009. https://doi.org/10.1002/bit.22256.
46. P.Y. Stergiou, A. Foukis, M. Filippou, M. Koukouritaki, M. Parapouli, L.G.
Theodorou, E. Hatziloukas, A. Afendra, A. Pandey, E.M. Papamichael,
Advances in lipase-catalyzed esterification reactions, Biotechnol. Adv. 31,
1846–1859, 2013. https://doi.org/10.1016/j.biotechadv.2013.08.006.
47. K.R. Jegannathan, S. Abang, D. Poncelet, E.S. Chan, P. Ravindra, Production
of biodiesel using immobilized lipase--a critical review., Crit. Rev. Biotechnol.
28, 253–264, 2008. https://doi.org/10.1080/07388550802428392.
48. J. Lu, Y. Chen, F. Wang, T. Tan, Effect of water on methanolysis of glycerol tri-
oleate catalyzed by immobilized lipase Candida sp. 99-125 in organic solvent
system, J. Mol. Catal. B Enzym. 56, 122–125, 2009. https://doi.org/10.1016/j.
molcatb.2008.05.004.
49. G.. Chowdary, S.. Prapulla, The influence of water activity on the lipase cat-
alyzed synthesis of butyl butyrate by transesterification, Process Biochem. 38,
393–397, 2002. https://doi.org/10.1016/S0032-9592(02)00096-1.
50. K. Nie, F. Xie, F. Wang, T. Tan, Lipase catalyzed methanolysis to produce
biodiesel: Optimization of the biodiesel production, J. Mol. Catal. B Enzym.
43, 142–147, 2006. https://doi.org/10.1016/j.molcatb.2006.07.016.
51. A. Canet, Kí. Bonet-Ragel, M.D. Benaiges, F. Valero, Biodiesel synthe-
sis in a solvent-free system by recombinant Rhizopus oryzae: comparative
study between a stirred tank and a packed-bed batch reactor, Biocatal.
Biotransformation. 35, 35–40, 2017. https://doi.org/10.1080/10242422.2016.
1278211.
52. S. Nigam, S. Mehrotra, B. Vani, R. Mehrotra, Lipase immobilization tech-
niques for biodiesel production: An Overview, Int. J. Renew. Energy Biofuels.
1–16, 2014. https://doi.org/10.5171/2014.664708.
53. P.M. Nielsen, J. Brask, L. Fjerbaek, Enzymatic biodiesel production: Technical
and economical considerations, Eur. J. Lipid Sci. Technol. 110, 692–700, 2008.
https://doi.org/10.1002/ejlt.200800064.
54. G.-Q. Chen, Biofunctionalization of polymers and their applications
Guo-Qiang, Adv. Biochem. Eng. Biotechnol. 125, 29–45, 2011. https://doi.
org/10.1007/10_2010_89.
55. A. Gog, M. Roman, M. Toşa, C. Paizs, F.D. Irimie, Biodiesel production using
enzymatic transesterification-Current state and perspectives, Renew. Energy.
2012. https://doi.org/10.1016/j.renene.2011.08.007.
56. A.P. Vyas, J.L. Verma, N. Subrahmanyam, Effects of molar ratio, alkali cata-
lyst concentration and temperature on transesterification of Jatropha oil with
methanol under ultrasonic irradiation, Adv. Chem. Eng. Sci. 01, 45–50, 2011.
https://doi.org/10.4236/aces.2011.12008.
57. M. Mathiyazhagan, a Ganapathi, Factors affecting biodiesel production, Res.
Plant Biol. 1, 1–5, 2011.
Biocatalytic Processes 43

58. T. Zhao, D.S. No, Y. Kim, Y.S. Kim, I.H. Kim, Novel strategy for lipase-catalyzed
synthesis of biodiesel using blended alcohol as an acyl acceptor, J. Mol. Catal.
B Enzym. 107, 17–22, 2014. https://doi.org/10.1016/j.molcatb.2014.05.002.
59. G. Bayramoglu, A. Akbulut, V.C. Ozalp, M.Y. Arica, Immobilized lipase on
micro-porous biosilica for enzymatic transesterification of algal oil, Chem.
Eng. Res. Des. 95, 12–21, 2015. https://doi.org/10.1016/j.cherd.2014.12.011.
60. Y. Wang, J. Liu, H. Gerken, C. Zhang, Q. Hu, Y. Li, Highly-efficient enzy-
matic conversion of crude algal oils into biodiesel, Bioresour. Technol. 172,
143–149, 2014. https://doi.org/10.1016/j.biortech.2014.09.003.
61. E. Bardone, A. Brucato, T. Keshavarz, T.M. Mata, I.R.B.G. Sousa, N.S.
Caetano, Transgenic corn oil for biodiesel production via enzymatic catalysis
with ethanol, Chem. Eng. Trans. 27, 19–24, 2012.
62. J.M. Encinar, J.F. González, J.J. Rodríguez, A. Tejedor, Biodiesel fuels from
vegetable oils: Transesterification of Cynara cardunculus L. Oils with ethanol,
Energy and Fuels. 16, 443–450, 2002. https://doi.org/10.1021/ef010174h.
63. J. Yan, Y. Yan, S. Liu, J. Hu, G. Wang, Preparation of cross-linked lipase-
coated micro-crystals for biodiesel production from waste cooking oil,
Bioresour. Technol. 102, 4755–4758, 2011. https://doi.org/10.1016/j.
biortech.2011.01.006.
64. K.P. Zhang, J.Q. Lai, Z.L. Huang, Z. Yang, Penicillium expansum lipase-­
catalyzed production of biodiesel in ionic liquids, Bioresour. Technol. 102
2767–2772, 2011. https://doi.org/10.1016/j.biortech.2010.11.057.
65. C. Santambrogio, F. Sasso, A. Natalello, S. Brocca, R. Grandori, S.M. Doglia,
M. Lotti, Effects of methanol on a methanol-tolerant bacterial lipase,
Appl. Microbiol. Biotechnol. 97, 8609–8618, 2013. https://doi.org/10.1007/
s00253-013-4712-5.
66. J.B. Rédéo Wilfried Moussavou Mounguenguia, Christel Brunschwiga,
Bruno Baréab, Pierre Villeneuveb, Are plant lipases a promising alterna-
tive to catalyze transesterification for biodiesel production?, Prog. Energy
Combust. Sci. 39, 441–456, 2013.
67. Y. Shimada, Y. Watanabe, T. Samukawa, A. Sugihara, H. Noda, H. Fukuda,
Y. Tominaga, Conversion of vegetable oil to biodiesel using immobilized
Candida antarctica lipase, J. Am. Oil Chem. Soc. 76, 789–793, 1999. https://
doi.org/10.1007/s11746-999-0067-6.
68. Y. Xu, Wei Du, Jing Zeng and Dehua Liu, Conversion of soybean oil to
biodiesel fuel using lipozyme TL IM in a solvent-free medium, Biocatal.
Biotransformation. 22, 45–48, 2004. https://doi.org/10.1080/1024242041000
1661222.
69. T.Y. Shimada Y, Watanabe Y, Sugihara A, Enzymatic alcoholysis for biodiesel
fuel production and application of the reaction to oil processing, J. Mol.
Catal. B Enzym. 17, 133–142, 2002.
70. S.H. Lee, D.R. Park, H. Kim, J. Lee, J.C. Jung, S.Y. Woo, W.S. Song, M.S. Kwon,
I.K. Song, Direct preparation of dichloropropanol (DCP) from glycerol using
44 Biodiesel Technology and Applications

heteropolyacid (HPA) catalysts: A catalyst screen study, Catal. Commun. 9


1920–1923, 2008. https://doi.org/10.1016/J.CATCOM.2008.03.020.
71. J.M. Marchetti, V.U. Miguel, A.F. Errazu, Techno-economic study of differ-
ent alternatives for biodiesel production, Fuel Process. Technol. 89, 740–748,
2008. https://doi.org/10.1016/j.fuproc.2008.01.007.
72. P.S. Wang, M.E. Tat, J. Van Gerpen, The production of fatty acid isopropyl
esters and their use as a diesel engine fuel, J. Am. Oil Chem. Soc. 82, 845–849,
2005. https://doi.org/10.1007/s11746-005-1153-7.
73. Q. Li, J. Xu, W. Du, Y. Li, D. Liu, Ethanol as the acyl acceptor for biodiesel
production, Renew. Sustain. Energy Rev. 25, 742–748, 2013. https://doi.
org/10.1016/J.RSER.2013.05.043.
74. K. Bozbas, Biodiesel as an alternative motor fuel: Production and policies in
the European Union, Renew. Sustain. Energy Rev. 12, 542–552, 2008. https://
doi.org/10.1016/J.RSER.2005.06.001.
75. P. Verma, V.M. Singh, Assessment of diesel engine performance using cotton
seed Biodiesel, Integr. Res. Adv. 1, 1–4, 2014.
76. E. Hernández-Martín, C. Otero, Different enzyme requirements for the syn-
thesis of biodiesel: Novozym® 435 and Lipozyme® TL IM, Bioresour. Technol.
99, 277–286, 2008. https://doi.org/10.1016/j.biortech.2006.12.024.
77. W. Du, Y. Xu, D. Liu, J. Zeng, Comparative study on lipase-catalyzed trans-
formation of soybean oil for biodiesel production with different acyl accep-
tors, J. Mol. Catal. B Enzym. 30, 125–129, 2004. https://doi.org/10.1016/j.
molcatb.2004.04.004.
78. M.K. Modi, J.R.C. Reddy, B.V.S.K. Rao, R.B.N. Prasad, Lipase-mediated
conversion of vegetable oils into biodiesel using ethyl acetate as acyl accep-
tor, Bioresour. Technol. 98, 1260–1264, 2007. https://doi.org/10.1016/j.
biortech.2006.05.006.
79. O.K. Lee, Y.H. Kim, J.G. Na, Y.K. Oh, E.Y. Lee, Highly efficient extraction and
lipase-catalyzed transesterification of triglycerides from Chlorella sp. KR-1
for production of biodiesel, Bioresour. Technol. 147, 240–245, 2013. https://
doi.org/10.1016/j.biortech.2013.08.037.
80. A.E. Ghaly, D. Dave, M.S. Brooks, S. Budge, Production of biodiesel by enzy-
matic transesterification: Review, Am. J. Biochem. Biotechnol. 2010. https://
doi.org/10.3844/ajbbsp.2010.54.76.
81. Y. Liu, H. Xin, Y. Yan, Physicochemical properties of stillingia oil: Feasibility
for biodiesel production by enzyme transesterification, Ind. Crops Prod. 30,
431–436, 2009. https://doi.org/10.1016/J.INDCROP.2009.08.004.
82. M.M. Soumanou, U.T. Bornscheuer, Improvement in lipase-catalyzed syn-
thesis of fatty acid methyl esters from sunflower oil, Enzyme Microb. Technol.
33, 97–103, 2003. https://doi.org/10.1016/S0141-0229(03)00090-5.
83. M. Xiao, R. Intan, J.P. Obbard, Biodiesel production from microalgae oil-
lipid feedstock via immobilized whole-cell biocatalysis, in: Third Int. Symp.
Energy from Biomass Waste Venice, Italy; 8-11 Novemb. 2010, 2010.
Biocatalytic Processes 45

84. D. Royon, M. Daz, G. Ellenrieder, S. Locatelli, Enzymatic production of bio-


diesel from cotton seed oil using t-butanol as a solvent, Bioresour. Technol.
98, 648–653, 2007. https://doi.org/10.1016/j.biortech.2006.02.021.
85. M. Iso, B. Chen, M. Eguchi, T. Kudo, S. Shrestha, Production of biodiesel fuel
from triglycerides and alcohol using immobilized lipase, J. Mol. Catal. - B
Enzym. 16, 53–58, 2001. https://doi.org/10.1016/S1381-1177(01)00045-5.
86. L. Li, W. Du, D. Liu, L. Wang, Z. Li, Lipase-catalyzed transesterification of
rapeseed oils for biodiesel production with a novel organic solvent as the
reaction medium, J. Mol. Catal. B Enzym. 43, 58–62, 2006. https://doi.
org/10.1016/j.molcatb.2006.06.012.
87. A. Robles-Medina, P.A. González-Moreno, L. Esteban-Cerdán, E. Molina-
Grima, Biocatalysis: Towards ever greener biodiesel production, Biotechnol.
Adv. 27, 398–408, 2009. https://doi.org/10.1016/j.biotechadv.2008.10.008.
88. K.S. Yang, J.H. Sohn, H.K. Kim, Catalytic properties of a lipase from
Photobacterium lipolyticum for biodiesel production containing a high
methanol concentration, J. Biosci. Bioeng. 107, 599–604, 2009. https://doi.
org/10.1016/j.jbiosc.2009.01.009.
89. H. Noureddini, X. Gao, R.S. Philkana, Immobilized Pseudomonas cepacia
lipase for biodiesel fuel production from soybean oil, Bioresour. Technol. 96,
769–777, 2005. https://doi.org/10.1016/j.biortech.2004.05.029.
90. J.H. Sim, A.H. Kamaruddin, S. Bhatia, Biodiesel (FAME) productivity, cat-
alytic efficiency and thermal stability of lipozyme TL im for crude palm oil
transesterification with methanol, JAOCS, J. Am. Oil Chem. Soc. 87, 1027–
1034, 2010. https://doi.org/10.1007/s11746-010-1593-y.
91. H. Taher, S. Al-Zuhair, A.H. Al-Marzouqi, Y. Haik, M. Farid, Enzymatic bio-
diesel production of microalgae lipids under supercritical carbon dioxide:
Process optimization and integration, Biochem. Eng. J. 90, 103–113, 2014.
https://doi.org/10.1016/j.bej.2014.05.019.
92. G.T. Jeong, D.H. Park, Lipase-catalyzed transesterification of rapeseed oil
for biodiesel production with ter-butanol, Appl. Biochem. Biotechnol. 148,
131–139, 2008. https://doi.org/10.1007/s12010-007-8050-x.
93. M. Szczesna Antczak, A. Kubiak, T. Antczak, S. Bielecki, Enzymatic biodiesel
synthesis - Key factors affecting efficiency of the process, Renew. Energy. 34,
1185–1194, 2009. https://doi.org/10.1016/j.renene.2008.11.013.
94. Y. Xu, M. Nordblad, P.M. Nielsen, J. Brask, J.M. Woodley, In situ visualization
and effect of glycerol in lipase-catalyzed ethanolysis of rapeseed oil, J. Mol. Catal.
B Enzym. 72, 213–219, 2011. https://doi.org/10.1016/J.MOLCATB.2011.06.008.
95. H.C. Chen, H.Y. Ju, T.T. Wu, Y.C. Liu, C.C. Lee, C. Chang, Y.L. Chung, C.J.
Shieh, Continuous production of lipase-catalyzed biodiesel in a packed-bed
reactor: Optimization and enzyme reuse study, J. Biomed. Biotechnol. 2011,
5–10, 2011. https://doi.org/10.1155/2011/950725.
96. K. Bélafi-Bakó, F. Kovács, L. Gubicza, J. Hancsók, Enzymatic biodiesel pro-
duction from sunflower oil by Candida antarctica lipase in a solvent-free
46 Biodiesel Technology and Applications

system, Biocatal. Biotransformation. 20, 437–439, 2002. https://doi.org/10.10


80/1024242021000040855.
97. A. Kumari, P. Mahapatra, V.K. Garlapati, R. Banerjee, Enzymatic transester-
ification of Jatropha oil, Biotechnol Biofuels. 2, 1, 2009. https://doi.org/1754-
6834-2-1 [pii]\r10.1186/1754-6834-2-1.
98. Y. Xu, W. Du, D. Liu, J. Zeng, A novel enzymatic route for biodiesel pro-
duction from renewable oils in a solvent-free medium, Biotechnol. Lett. 25,
1239–1241, 2003. https://doi.org/10.1023/A:1025065209983.
99. L. Zhang, S. Sun, Z. Xin, B. Sheng, Q. Liu, Synthesis and component con-
firmation of biodiesel from palm oil and dimethyl carbonate catalyzed by
immobilized-lipase in solvent-free system, Fuel. 89, 3960–3965, 2010. https://
doi.org/10.1016/j.fuel.2010.06.030.
100. Z. Wang, J. Zhuge, H. Fang, B.A. Prior, Glycerol production by microbial
fermentation: A review, Biotechnol. Adv. 19, 201–223, 2001. https://doi.
org/10.1016/S0734-9750(01)00060-X.
101. Y. Zheng, J. Quan, X. Ning, L.M. Zhu, B. Jiang, Z.Y. He, Lipase-catalyzed
transesterification of soybean oil for biodiesel production in tert-amyl alco-
hol, World J. Microbiol. Biotechnol. 25, 41–46, 2009. https://doi.org/10.1007/
s11274-008-9858-4.
102. P.S. Bisen, B.S. Sanodiya, G.S. Thakur, R.K. Baghel, G.B.K.S. Prasad,
Biodiesel production with special emphasis on lipase-catalyzed transes-
terification, Biotechnol. Lett. 2010. https://doi.org/10.1007/s10529-010-
0275-z.
103. H. Ghamgui, M. Karra-Chaâbouni, Y. Gargouri, 1-Butyl oleate synthesis
by immobilized lipase from Rhizopus oryzae: a comparative study between
n-hexane and solvent-free system, Enzyme Microb. Technol. 35, 355–363,
2004. https://doi.org/10.1016/J.ENZMICTEC.2004.06.002.
104. M. Xiao, C. Qi, J.P. Obbard, Biodiesel production using Aspergillus niger as a
whole-cell biocatalyst in a packed-bed reactor, GCB Bioenergy. 3, 293–298,
2011. https://doi.org/10.1111/j.1757-1707.2010.01087.x.
105. Ö. Aybastıer, C. Demir, Immobilization of Candida antarctica lipase A on
chitosan beads for the production of fatty acid methyl ester from waste fry-
ing oil, energy sources, Part A Recover. Util. Environ. Eff. 36, 2313–2319,
2014. https://doi.org/10.1080/15567036.2011.567233.
106. S. Kojima, D. Du, M. Sato, E.Y. Park, Efficient production of fatty acid methyl
ester from waste activated bleaching earth using diesel oil as organic solvent,
J. Biosci. Bioeng. 98, 420–424, 2004. https://doi.org/http://dx.doi.org/10.1016/
S1389-1723(05)00306-3.
107. A. Bajaj, P. Lohan, P.N. Jha, R. Mehrotra, Biodiesel production through lipase
catalyzed transesterification: An overview, J. Mol. Catal. B Enzym. 62, 9–14,
2010. https://doi.org/10.1016/j.molcatb.2009.09.018.
108. S. Ycel, P. Terziolu, D. zime, Lipase Applications in Biodiesel Production,
in: Biodiesel - Feed. Prod. Appl., InTech, 2012. https://doi.org/10.5772/
52662.
Biocatalytic Processes 47

109. M.Y. Noraini, H.C. Ong, M.J. Badrul, W.T. Chong, A review on potential
enzymatic reaction for biofuel production from algae, Renew. Sustain. Energy
Rev., 2014. https://doi.org/10.1016/j.rser.2014.07.089.
110. P. Priji, K.N. Unni, S. Sajith, P. Binod, S. Benjamin, Production, optimiza-
tion, and partial purification of lipase from Pseudomonas sp. strain BUP6, a
novel rumen bacterium characterized from Malabari goat, Biotechnol. Appl.
Biochem., 2015. https://doi.org/10.1002/bab.1237.
111. A. Kumar, S.S. Parihar, N. Batra, Enrichment, isolation and optimization of
lipase-producing Staphylococcus sp. from oil mill waste (Oil cake), J. Exp.
Sci., 2012.
112. S. Javed, F. Azeem, S. Hussain, I. Rasul, M.H. Siddique, M. Riaz, M. Afzal, A.
Kouser, H. Nadeem, Bacterial lipases: A review on purification and character-
ization, Prog. Biophys. Mol. Biol. 132, 23–34, 2018. https://doi.org/10.1016/j.
pbiomolbio.2017.07.014.
113. B.D. Ribeiro, A.M. de Castro, M.A.Z. Coelho, D.M.G. Freire, Production and
use of lipases in bioenergy: A review from the feedstocks to biodiesel produc-
tion, Enzyme Res. 2011, 1–16, 2011. https://doi.org/10.4061/2011/615803.
114. M. Kapoor, M.N. Gupta, Lipase promiscuity and its biochemical applica-
tions, Process Biochem., 2012. https://doi.org/10.1016/j.procbio.2012.01.011.
115. A.L. Paiva, V.M. Balcão, F.X. Malcata, Kinetics and mechanisms of reactions
catalyzed by immobilized lipases, Enzyme Microb. Technol., 2000. https://doi.
org/10.1016/S0141-0229(00)00206-4.
116. A. Türkan, Ş. Kalay, Monitoring lipase-catalyzed methanolysis of sunflower
oil by reversed-phase high-performance liquid chromatography: Elucidation
of the mechanisms of lipases, J. Chromatogr. A. 1127, 34–44, 2006. https://
doi.org/10.1016/j.chroma.2006.05.065.
117. D. Bezbradica, M. Stojanović, D. Veličković, A. Dimitrijević, M. Carević, M.
Mihailović, N. Milosavić, Kinetic model of lipase-catalyzed conversion of
ascorbic acid and oleic acid to liposoluble vitamin C ester, Biochem. Eng. J.,
2013. https://doi.org/10.1016/j.bej.2012.12.001.
118. A.E.M. Janssen, B.J. Sjursnes, A. V. Vakurov, P.J. Halling, Kinetics of lipase-­
catalyzed esterification in organic media: Correct model and solvent effects
on parameters, Enzyme Microb. Technol., 1999. https://doi.org/10.1016/
S0141-0229(98)00134-3.
119. K.-E. Jaeger, B.W. Dijkstra, M.T. Reetz, Bacterial biocatalysts: Molecular
biology, three-dimensional structures, and biotechnological applications of
lipases, Annu. Rev. Microbiol. 53, 315–351, 1999. https://doi.org/10.1146/
annurev.micro.53.1.315.
120. F. Hasan, A.A. Shah, A. Hameed, Industrial applications of microbial lipases,
Enzyme Microb. Technol. 39, 235–251, 2006. https://doi.org/10.1016/j.
enzmictec.2005.10.016.
121. H. Treichel, D. de Oliveira, M.A. Mazutti, M. Di Luccio, J.V. Oliveira, A
Review on Microbial Lipases Production, Food Bioprocess Technol. 3, 182–
196, 2010. https://doi.org/10.1007/s11947-009-0202-2.
48 Biodiesel Technology and Applications

122. C. Song, L. Sheng, X. Zhang, Immobilization and Characterization of


a Thermostable Lipase, Mar. Biotechnol. 15, 659–667, 2013. https://doi.
org/10.1007/s10126-013-9515-2.
123. Y. Hotta, S. Ezaki, H. Atomi, T. Imanaka, Extremely stable and versatile car-
boxylesterase from a Hyperthermophilic archaeon., Appl. Environ. Microbiol.
68, 3925–31, 2002. https://doi.org/10.1128/AEM.68.8.3925-3931.2002.
124. M. Royter, M. Schmidt, C. Elend, H. Höbenreich, T. Schäfer, U.T. Bornscheuer,
G. Antranikian, Thermostable lipases from the extreme thermophilic
anaerobic bacteria Thermoanaerobacter thermohydrosulfuricus SOL1 and
Caldanaerobacter subterraneus subsp. tengcongensis, Extremophiles. 13,
769–783, 2009. https://doi.org/10.1007/s00792-009-0265-z.
125. K. Kawakami, Y. Oda, R. Takahashi, Application of a Burkholderia cepacia
lipase-immobilized silica monolith to batch and continuous biodiesel pro-
duction with a stoichiometric mixture of methanol and crude Jatropha oil,
Biotechnol. Biofuels. 4, 42, 2011. https://doi.org/10.1186/1754-6834-4-42.
126. R. Abdulla, P. Ravindra, Immobilized Burkholderia cepacia lipase for bio-
diesel production from crude Jatropha curcas L. oil, Biomass and Bioenergy.,
2013. https://doi.org/10.1016/j.biombioe.2013.04.010.
127. K.R. Jegannathan, L. Jun-Yee, E.-S. Chan, P. Ravindra, Design an immobi-
lized lipase enzyme for biodiesel production, J. Renew. Sustain. Energy. 1,
063101, 2009. https://doi.org/10.1063/1.3256191.
128. Q. You, X. Yin, Y. Zhao, Y. Zhang, Biodiesel production from jatropha oil
catalyzed by immobilized Burkholderia cepacia lipase on modified atta-
pulgite, Bioresour. Technol. 148, 202–207, 2013. https://doi.org/10.1016/j.
biortech.2013.08.143.
129. P.C.M. Da Rós, W.C. e Silva, D. Grabauskas, V.H. Perez, H.F. de Castro,
Biodiesel from babassu oil: Characterization of the product obtained by
enzymatic route accelerated by microwave irradiation, Ind. Crops Prod. 52,
313–320, 2014. https://doi.org/10.1016/j.indcrop.2013.11.013.
130. P.C.M. Da Rós, G.A.M. Silva, A.A. Mendes, J.C. Santos, H.F. de Castro,
Evaluation of the catalytic properties of Burkholderia cepacia lipase immo-
bilized on non-commercial matrices to be used in biodiesel synthesis from
different feedstocks, Bioresour. Technol. 101, 5508–5516, 2010. https://doi.
org/10.1016/j.biortech.2010.02.061.
131. C.G. Lopresto, S. Naccarato, L. Albo, M.G. De Paola, S. Chakraborty, S.
Curcio, V. Calabro, Enzymatic transesterification of waste vegetable oil to
produce biodiesel, Ecotoxicol. Environ. Saf. 121, 229–235, 2015. https://doi.
org/10.1016/j.ecoenv.2015.03.028.
132. A.-F. Hsu, K. Jones, T.A. Foglia, W.N. Marmer, Immobilized lipase-catalysed
production of alkyl esters of restaurant grease as biodiesel, Biotechnol. Appl.
Biochem. 36, 181–186, 2002. https://doi.org/10.1042/.
133. X. Wang, X. Liu, C. Zhao, Y. Ding, P. Xu, Biodiesel production in packed-
bed reactors using lipase-nanoparticle biocomposite, Bioresour. Technol. 102,
6352–6355, 2011. https://doi.org/10.1016/j.biortech.2011.03.003.
Biocatalytic Processes 49

134. V. Kumari, S. Shah, M.N. Gupta, Preparation of biodiesel by lipase-catalyzed


transesterification of high free fatty acid containing oil from Madhuca indica,
Energy and Fuels. 21, 368–372, 2007. https://doi.org/10.1021/ef0602168.
135. J. Zheng, L. Xu, Y. Liu, X. Zhang, Y. Yan, Lipase-coated K2SO4 micro-crystals:
Preparation, characterization, and application in biodiesel production using
various oil feedstocks, Bioresour. Technol. 110, 224–231, 2012. https://doi.
org/10.1016/J.BIORTECH.2012.01.088.
136. S. Shah, M.N. Gupta, Lipase catalyzed preparation of biodiesel from Jatropha
oil in a solvent free system, Process Biochem. 42, 409–414, 2007. https://doi.
org/10.1016/j.procbio.2006.09.024.
137. L.N. Lima, G.C. Oliveira, M.J. Rojas, H.F. Castro, P.C.M. Da Rós, A.A.
Mendes, R.L.C. Giordano, P.W. Tardioli, Immobilization of Pseudomonas
fluorescens lipase on hydrophobic supports and application in biodiesel syn-
thesis by transesterification of vegetable oils in solvent-free systems, J. Ind.
Microbiol. Biotechnol. 42, 523–535, 2015. https://doi.org/10.1007/s10295-
015-1586-9.
138. A.L. Machsun, M. Gozan, M. Nasikin, S. Setyahadi, Y.J. Yoo, Membrane
microreactor in biocatalytic transesterification of triolein for biodiesel
production, Biotechnol. Bioprocess Eng. 15, 911–916, 2010. https://doi.
org/10.1007/s12257-010-0151-7.
139. G. Dors, L. Freitas, A.A. Mendes, A. Furigo, H.F. De Castro, Transesterification
of palm oil catalyzed by Pseudomonas fluorescens lipase in a packed-bed
reactor, Energy and Fuels. 26, 5977–5982, 2012. https://doi.org/10.1021/
ef300931y.
140. A.B.R. Moreira, V.H. Perez, G.M. Zanin, H.F. de Castro, Biodiesel synthesis
by enzymatic transesterification of palm oil with ethanol using lipases from
several sources immobilized on silica-PVA composite, Energy and Fuels. 21,
3689–3694, 2007. https://doi.org/10.1021/ef700399b.
141. L. Wang, Z. Chi, X. Wang, Z. Liu, J. Li, Diversity of lipase-producing yeasts
from marine environments and oil hydrolysis by their crude enzymes, Ann.
Microbiol. 57, 495–501, 2007. https://doi.org/10.1007/BF03175345.
142. N. Li, M.-H. Zong, Lipases from the genus Penicillium: Production, purifi-
cation, characterization and applications, J. Mol. Catal. B Enzym. 66, 43–54,
2010. https://doi.org/10.1016/j.molcatb.2010.05.004.
143. R. Gupta, N. Gupta, P. Rathi, Bacterial lipases: An overview of production,
purification and biochemical properties, Appl. Microbiol. Biotechnol. 64,
763–781, 2004. https://doi.org/10.1007/s00253-004-1568-8.
144. R.. Saxena, A. Sheoran, B. Giri, W.S. Davidson, Purification strategies
for microbial lipases, J. Microbiol. Methods. 52, 1–18, 2003. https://doi.
org/10.1016/S0167-7012(02)00161-6.
145. J.H. Lee, S.B. Kim, H.Y. Yoo, Y.J. Suh, G.B. Kang, W.I. Jang, J. Kang, C. Park,
S.W. Kim, Biodiesel production by enzymatic process using Jatropha oil and
waste soybean oil, Biotechnol. Bioprocess Eng. 18, 703–708, 2013. https://doi.
org/10.1007/s12257-012-0805-8.
50 Biodiesel Technology and Applications

146. L. Deng, X. Xu, G.G. Haraldsson, T. Tan, F. Wang, Enzymatic production of


alkyl esters through alcoholysis: A critical evaluation of lipases and alcohols,
JAOCS, J. Am. Oil Chem. Soc. 82, 341–347, 2005. https://doi.org/10.1007/
s11746-005-1076-3.
147. I.G. Rosset, M.C.H.T. Cavalheiro, E.M. Assaf, A.L.M. Porto, Enzymatic
esterification of oleic acid with aliphatic alcohols for the biodiesel production
by Candida antarctica lipase, Catal. Letters. 143, 863–872, 2013. https://doi.
org/10.1007/s10562-013-1044-0.
148. W. Du, Y. Xu, J. Zeng, D. Liu, Novozym 435-catalysed transesterification
of crude soya bean oils for biodiesel production in a solvent-free medium.,
Biotechnol. Appl. Biochem. 40, 187–90, 2004. https://doi.org/10.1042/
BA20030142.
149. S.J. Kim, S.M. Jung, Y.C. Park, K. Park, Lipase catalyzed transesterification
of soybean oil using ethyl acetate, an alternative acyl acceptor, Biotechnol.
Bioprocess Eng. 12, 441–445, 2007. https://doi.org/10.1007/BF02931068.
150. C. Dalla Rosa, M.B. Morandim, J.L. Ninow, D. Oliveira, H. Treichel, J.V.
Oliveira, Continuous lipase-catalyzed production of fatty acid ethyl esters
from soybean oil in compressed fluids, Bioresour. Technol. 100, 5818–5826,
2009. https://doi.org/10.1016/j.biortech.2009.06.081.
151. W. Xie, N. Ma, Enzymatic transesterification of soybean oil by using immo-
bilized lipase on magnetic nano-particles, Biomass and Bioenergy. 34, 890–
896, 2010. https://doi.org/10.1016/j.biombioe.2010.01.034.
152. F. Yagiz, D. Kazan, A.N. Akin, Biodiesel production from waste oils by using
lipase immobilized on hydrotalcite and zeolites, Chem. Eng. J. 134, 262–267,
2007. https://doi.org/10.1016/j.cej.2007.03.041.
153. M. Basri, M.A. Kassim, R. Mohamad, A.B. Ariff, Optimization and kinetic
study on the synthesis of palm oil ester using Lipozyme TL im, J. Mol.
Catal. B Enzym. 85–86, 214–219, 2013. https://doi.org/10.1016/j.molcatb.
2012.09.013.
154. Y. Wang, H. Wu, M.H. Zong, Improvement of biodiesel production by
lipozyme TL IM-catalyzed methanolysis using response surface methodol-
ogy and acyl migration enhancer, Bioresour. Technol. 99, 7232–7237, 2008.
https://doi.org/10.1016/j.biortech.2007.12.062.
155. D. von der Haar, A. Stabler, R. Wichmann, U. Schweiggert-Weisz, Enzymatic
esterification of free fatty acids in vegetable oils utilizing different immo-
bilized lipases, Biotechnol. Lett. 37, 169–174, 2015. https://doi.org/10.1007/
s10529-014-1668-1.
156. N.L. Facioli, D. Barrera-Arellano, Optimisation of enzymatic esterification
of soybean oil deodoriser distillate, J. Sci. Food Agric. 81, 1193–1198, 2001.
https://doi.org/10.1002/jsfa.928.
157. Y. Chen, B. Xiao, J. Chang, Y. Fu, P. Lv, X. Wang, Synthesis of biodiesel
from waste cooking oil using immobilized lipase in fixed bed reactor,
Energy Convers. Manag. 50, 668–673, 2009. https://doi.org/10.1016/j.
enconman.2008.10.011.
Biocatalytic Processes 51

158. C.M.T. Santin, R.P. Scherer, N.L.D. Nyari, C.D. Rosa, R.M. Dallago, D. de
Oliveira, J.V. Oliveira, Batch esterification of fatty acids charges under ultra-
sound irradiation using Candida antarctica B immobilized in polyurethane
foam, Biocatal. Agric. Biotechnol. 3, 90–94, 2014. https://doi.org/10.1016/j.
bcab.2014.02.005.
159. W. Xie, J. Wang, Enzymatic production of biodiesel from soybean oil by
using immobilized lipase on Fe3O4/Poly(styrene-methacrylic acid) magnetic
microsphere as a biocatalyst, Energy and Fuels. 28, 2624–2631, 2014. https://
doi.org/10.1021/ef500131s.
160. J.C. Moreno-Pirajàn, L. Giraldo, Study of immobilized candida rugosa lipase
for biodiesel fuel production from palm oil by flow microcalorimetry, Arab.
J. Chem. 4, 55–62, 2011. https://doi.org/10.1016/j.arabjc.2010.06.019.
161. Y. Yücel, Biodiesel production from pomace oil by using lipase immobilized
onto olive pomace, Bioresour. Technol. 102, 3977–3980, 2011. https://doi.
org/10.1016/j.biortech.2010.12.001.
162. A.F. Hsu, K.C. Jones, T.A. Foglia, W.N. Marmer, Transesterification activity
of lipases immobilized in a phyllosilicate sol-gel matrix, Biotechnol. Lett. 26,
917–921, 2004. https://doi.org/10.1023/B:bile.0000025903.11697.ae.
163. J.S. Miranda, N.C.A. Silva, J.J. Bassi, M.C.C. Corradini, F.A.P. Lage, D.B.
Hirata, A.A. Mendes, Immobilization of Thermomyces lanuginosus lipase on
mesoporous poly-hydroxybutyrate particles and application in alkyl esters
synthesis: Isotherm, thermodynamic and mass transfer studies, Chem. Eng.
J. 251, 392–403, 2014. https://doi.org/10.1016/j.cej.2014.04.087.
164. A.A. Mendes, R.C. Giordano, R.D.L.C. Giordano, H.F. De Castro,
Immobilization and stabilization of microbial lipases by multipoint cova-
lent attachment on aldehyde-resin affinity: Application of the biocatalysts in
biodiesel synthesis, J. Mol. Catal. B Enzym. 68, 109–115, 2011. https://doi.
org/10.1016/j.molcatb.2010.10.002.
165. S.V. Ranganathan, S.L. Narasimhan, K. Muthukumar, An overview of
enzymatic production of biodiesel, Bioresour. Technol., 2008. https://doi.
org/10.1016/j.biortech.2007.04.060.
166. H. Fukuda, S. Hama, S. Tamalampudi, H. Noda, Whole-cell biocatalysts for
biodiesel fuel production, Trends Biotechnol., 2008. https://doi.org/10.1016/j.
tibtech.2008.08.001.
167. M.S. Soares, A.L.L. Rico, G.S.S. Andrade, H.F. De Castro, P.C. Oliveira,
Synthesis, characterization and application of a polyurethane-based support
for immobilizing membrane-bound lipase, Brazilian J. Chem. Eng., 2017.
https://doi.org/10.1590/0104-6632.20170341s20140227.
168. B. Atkinson, G.M. Black, P.J.S. Lewis, A. Pinches, Biological particles of given
size, shape, and density for use in biological reactors, Biotechnol. Bioeng.,
1979. https://doi.org/10.1002/bit.260210206.
169. S. Ertuğrul, G. Dönmez, S. Takaç, Isolation of lipase producing Bacillus sp.
from olive mill wastewater and improving its enzyme activity, J. Hazard.
Mater., 2007. https://doi.org/10.1016/j.jhazmat.2007.04.034.
52 Biodiesel Technology and Applications

170. V. Ramakrishnan, L.C. Goveas, N. Suralikerimath, C. Jampani, P.M. Halami,


B. Narayan, Extraction and purification of lipase from Enterococcus faecium
MTCC5695 by PEG/phosphate aqueous-two phase system (ATPS) and its
biochemical characterization, Biocatal. Agric. Biotechnol., 2016. https://doi.
org/10.1016/j.bcab.2016.02.005.
171. M. Adamczak, W. Bednarski, Enhanced activity of intracellular lipases
from Rhizomucor miehei and Yarrowia lipolytica by immobilization on bio-
mass support particles, Process Biochem., 2004. https://doi.org/10.1016/
S0032-9592(03)00266-8.
172. L. Ping, X. Yuan, M. Zhang, Y. Chai, S. Shan, Improvement of extracellu-
lar lipase production by a newly isolated Yarrowia lipolytica mutant and its
application in the biosynthesis of L-ascorbyl palmitate, Int. J. Biol. Macromol.,
2018. https://doi.org/10.1016/j.ijbiomac.2017.08.016.
173. J. Huang, J. Xia, W. Jiang, Y. Li, J. Li, Biodiesel production from microalgae
oil catalyzed by a recombinant lipase, Bioresour. Technol. 180, 47–53, 2015.
https://doi.org/10.1016/j.biortech.2014.12.072.
174. D. Adachi, F. Koh, S. Hama, C. Ogino, A. Kondo, A robust whole-cell bio-
catalyst that introduces a thermo- and solvent-tolerant lipase into Aspergillus
oryzae cells: Characterization and application to enzymatic biodiesel produc-
tion, Enzyme Microb. Technol. 52, 331–335, 2013. https://doi.org/10.1016/j.
enzmictec.2013.03.005.
175. Y. Luo, Y. Zheng, Z. Jiang, Y. Ma, D. Wei, A novel psychrophilic lipase from
Pseudomonas fluorescens with unique property in chiral resolution and bio-
diesel production via transesterification, Appl. Microbiol. Biotechnol. 73,
349–355, 2006. https://doi.org/10.1007/s00253-006-0478-3.
176. J. Yan, X. Zheng, S. Li, A novel and robust recombinant Pichia pastoris yeast
whole cell biocatalyst with intracellular overexpression of a Thermomyces
lanuginosus lipase: Preparation, characterization and application in biodiesel
production, Bioresour. Technol. 151, 43–48, 2014. https://doi.org/10.1016/j.
biortech.2013.10.037.
177. G.-C. Lee, L.-C. Lee, V. Sava, J.-F. Shaw, Multiple mutagenesis of non-­
universal serine codons of the Candida rugosa LIP2 gene and biochemi-
cal characterization of purified recombinant LIP2 lipase overexpressed in
Pichia pastoris., Biochem. J. 366, 603–611, 2002. https://doi.org/10.1042/
BJ20020404.
178. A. Yoshida, S. Hama, N. Tamadani, H. Noda, H. Fukuda, A. Kondo,
Continuous production of biodiesel using whole-cell biocatalysts: Sequential
conversion of an aqueous oil emulsion into anhydrous product, Biochem.
Eng. J. 68, 7–11, 2012. https://doi.org/10.1016/j.bej.2012.07.002.
179. D. Adachi, S. Hama, T. Numata, K. Nakashima, C. Ogino, H. Fukuda, A.
Kondo, Development of an Aspergillus oryzae whole-cell biocatalyst coex-
pressing triglyceride and partial glyceride lipases for biodiesel produc-
tion, Bioresour. Technol. 102, 6723–6729, 2011. https://doi.org/10.1016/j.
biortech.2011.03.066.
Biocatalytic Processes 53

180. M. Raita, V. Champreda, N. Laosiripojana, Biocatalytic ethanolysis of palm


oil for biodiesel production using microcrystalline lipase in tert-buta-
nol system, Process Biochem. 45, 829–834, 2010. https://doi.org/10.1016/j.
procbio.2010.02.002.
181. T.P. Korman, B. Sahachartsiri, D.M. Charbonneau, G.L. Huang, M.
Beauregard, J.U. Bowie, Dieselzymes: development of a stable and metha-
nol tolerant lipase for biodiesel production by directed evolution, Biotechnol.
Biofuels. 6, 70, 2013. https://doi.org/10.1186/1754-6834-6-70.
182. S.H. Kim, S.J. Kim, S. Park, H.K. Kim, Biodiesel production using cross-
linked Staphylococcus haemolyticus lipase immobilized on solid polymeric
carriers, J. Mol. Catal. B Enzym. 85–86, 10–16, 2013. https://doi.org/10.1016/j.
molcatb.2012.08.012.
183. A. Li, T.P.N. Ngo, J. Yan, K. Tian, Z. Li, Whole-cell based solvent-free system
for one-pot production of biodiesel from waste grease, Bioresour. Technol.
114, 725–729, 2012. https://doi.org/10.1016/j.biortech.2012.03.034.
184. F. Guan, P. Peng, G. Wang, T. Yin, Q. Peng, J. Huang, G. Guan, Y. Li,
Combination of two lipases more efficiently catalyzes methanolysis of soy-
bean oil for biodiesel production in aqueous medium, Process Biochem. 45,
1677–1682, 2010. https://doi.org/10.1016/j.procbio.2010.06.021.
185. T. Takaya, R. Koda, D. Adachi, K. Nakashima, J. Wada, T. Bogaki, C. Ogino, A.
Kondo, Highly efficient biodiesel production by a whole-cell biocatalyst employ-
ing a system with high lipase expression in Aspergillus oryzae, Appl. Microbiol.
Biotechnol. 90, 1171–1177, 2011. https://doi.org/10.1007/s00253-011-3186-6.
186. D. Huang, S. Han, Z. Han, Y. Lin, Biodiesel production catalyzed by
Rhizomucor miehei lipase-displaying Pichia pastoris whole cells in an isooc-
tane system, Biochem. Eng. J. 63, 10–14, 2012. https://doi.org/10.1016/j.
bej.2010.08.009.
187. Z.L. Huang, T.X. Yang, J.Z. Huang, Z. Yang, Enzymatic production of bio-
diesel from Millettia pinnata seed oil in ionic liquids, Bioenergy Res. 7, 1519–
1528, 2014. https://doi.org/10.1007/s12155-014-9489-6.
188. T. Matsumoto, S. Takahashi, M. Kaieda, M. Ueda, A. Tanaka, H. Fukuda,
A. Kondo, Yeast whole-cell biocatalyst constructed by intracellular over-
production of Rhizopus oryzae lipase is applicable to biodiesel fuel produc-
tion, Appl. Microbiol. Biotechnol. 57, 515–520, 2001. https://doi.org/10.1007/
s002530100733.
189. B. Zhang, Y. Weng, H. Xu, Z. Mao, Enzyme immobilization for biodiesel pro-
duction, Appl. Microbiol. Biotechnol. 93, 61–70, 2012. https://doi.org/10.1007/
s00253-011-3672-x.
190. R. Pogaku, A. Bono, C. Chu, Developments in sustainable chemical
and bioprocess technology, Springer US, Boston, MA, 2013. https://doi.
org/10.1007/978-1-4614-6208-8.
191. S.K. Narwal, R. Gupta, Biodiesel production by transesterification using
immobilized lipase, Biotechnol. Lett. 35, 479–490, 2013. https://doi.
org/10.1007/s10529-012-1116-z.
54 Biodiesel Technology and Applications

192. D. Brady, J. Jordaan, Advances in enzyme immobilisation, Biotechnol. Lett.


31, 1639–1650, 2009. https://doi.org/10.1007/s10529-009-0076-4.
193. M. Kalantari, M. Kazemeini, A. Arpanaei, Evaluation of biodiesel production
using lipase immobilized on magnetic silica nanocomposite particles of var-
ious structures, Biochem. Eng. J. 79, 267–273, 2013. https://doi.org/10.1016/j.
bej.2013.09.001.
194. S. Shah, S. Sharma, M.N. Gupta, Biodiesel preparation by lipase-catalyzed
transesterification of Jatropha oil, Energy and Fuels. 18, 154–159, 2004.
https://doi.org/10.1021/ef030075z.
195. S. Shah, S. Sharma, M.N. Gupta, Enzymatic transesterification for biodiesel
production, Indian J. Biochem. Biophys. 40, 392–399, 2003. https://doi.
org/10.1039/C6RA08062F.
196. P. Shao, X. Meng, J. He, P. Sun, Analysis of immobilized Candida rugosa lipase
catalyzed preparation of biodiesel from rapeseed soapstock, Food Bioprod.
Process. 86, 283–289, 2008. https://doi.org/10.1016/j.fbp.2008.02.004.
197. K. Ban, S. Hama, K. Nishizuka, M. Kaieda, T. Matsumoto, A. Kondo, H.
Noda, H. Fukuda, Repeated use of whole-cell biocatalysts immobilized
within biomass support particles for biodiesel fuel production, J. Mol.
Catal. - B Enzym. 17, 157–165, 2002. https://doi.org/10.1016/S1381-1177
(02)00023-1.
198. M.G.M. Purwanto, M.V. Maretha, M. Wahyudi, M.T. Goeltom, Whole
Cell Hydrolysis of Sardine (Sardinella Lemuru) Oil Waste Using Mucor
Circinelloides NRRL 1405 Immobilized in Poly-urethane Foam, Procedia
Chem. 14, 256–262, 2015. https://doi.org/10.1016/j.proche.2015.03.036.
199. G.S.S. Andrade, A.K.F. Carvalho, C.M. Romero, P.C. Oliveira, H.F. De
Castro, Mucor circinelloides whole-cells as a biocatalyst for the production
of ethyl esters based on babassu oil, Bioprocess Biosyst. Eng. 37, 2539–2548,
2014. https://doi.org/10.1007/s00449-014-1231-4.
200. W. Li, W. Du, D. Liu, Optimization of whole cell-catalyzed methanolysis
of soybean oil for biodiesel production using response surface methodol-
ogy, J. Mol. Catal. B Enzym. 45, 122–127, 2007. https://doi.org/10.1016/j.
molcatb.2007.01.002.
201. S. Tamalampudi, M.R. Talukder, S. Hama, T. Numata, A. Kondo, H. Fukuda,
Enzymatic production of biodiesel from Jatropha oil: A comparative study
of immobilized-whole cell and commercial lipases as a biocatalyst, Biochem.
Eng. J. 39, 185–189, 2008. https://doi.org/10.1016/j.bej.2007.09.002.
202. S. Hama, H. Yamaji, T. Fukumizu, T. Numata, S. Tamalampudi, A. Kondo, H.
Noda, H. Fukuda, Biodiesel-fuel production in a packed-bed reactor using
lipase-producing Rhizopus oryzae cells immobilized within biomass sup-
port particles, Biochem. Eng. J. 34, 273–278, 2007. https://doi.org/10.1016/j.
bej.2006.12.013.
203. B. Balasubramaniam, A. Sudalaiyadum Perumal, J. Jayaraman, J. Mani, P.
Ramanujam, Comparative analysis for the production of fatty acid alkyl
esterase using whole cell biocatalyst and purified enzyme from Rhizopus
Biocatalytic Processes 55

oryzae on waste cooking oil (sunflower oil), Waste Manag. 32, 1539–1547,
2012. https://doi.org/10.1016/j.wasman.2012.03.011.
204. S. Athalye, R. Sharma-Shivappa, S. Peretti, P. Kolar, J.P. Davis, Producing
biodiesel from cottonseed oil using Rhizopus oryzae ATCC #34612 whole
cell biocatalysts: Culture media and cultivation period optimization, Energy
Sustain. Dev. 17, 331–336, 2013. https://doi.org/10.1016/j.esd.2013.03.009.
205. M.G. Devanesan, T. Viruthagiri, N. Sugumar, Transesterification of Jatropha
oil using immobilized Pseudomonas fluorescens, African J. Biotechnol. 6,
2497–2501, 2007. https://doi.org/10.4314/ajb.v6i21.58115.
206. Q. Li, J. Zheng, Y. Yan, Biodiesel preparation catalyzed by compound-
lipase in co-solvent, Fuel Process. Technol. 91, 1229–1234, 2010. https://doi.
org/10.1016/j.fuproc.2010.04.002.
207. R.C. Rodrigues, M.A.Z. Ayub, Effects of the combined use of Thermomyces
lanuginosus and Rhizomucor miehei lipases for the transesterification and
hydrolysis of soybean oil, Process Biochem. 46, 682–688, 2011. https://doi.
org/10.1016/j.procbio.2010.11.013.
208. J.K. Poppe, C.R. Matte, M. Do Carmo Ruaro Peralba, R. Fernandez-Lafuente,
R.C. Rodrigues, M.A.Z. Ayub, Optimization of ethyl ester production from
olive and palm oils using mixtures of immobilized lipases, Appl. Catal. A
Gen. 490, 50–56, 2015. https://doi.org/10.1016/j.apcata.2014.10.050.
209. K. Tongboriboon, B. Cheirsilp, A. H-Kittikun, Mixed lipases for efficient
enzymatic synthesis of biodiesel from used palm oil and ethanol in a solvent-­
free system, J. Mol. Catal. B Enzym. 67, 52–59, 2010. https://doi.org/10.1016/j.
molcatb.2010.07.005.
210. A. Salis, M. Pinna, M. Monduzzi, V. Solinas, Comparison among immo-
bilised lipases on macroporous polypropylene toward biodiesel synthe-
sis, J. Mol. Catal. B Enzym. 54, 19–26, 2008. https://doi.org/10.1016/j.
molcatb.2007.12.006.
211. A.R. Rodrigues, A. Paiva, M.G. Da Silva, P. Simões, S. Barreiros, Continuous
enzymatic production of biodiesel from virgin and waste sunflower oil in
supercritical carbon dioxide, in: J. Supercrit. Fluids, pp. 259–264, 2011.
https://doi.org/10.1016/j.supflu.2010.10.031.
212. J.S. Alves, N.S. Vieira, A.S. Cunha, A.M. Silva, M.A. Záchia Ayub, R.
Fernandez-Lafuente, R.C. Rodrigues, Combi-lipase for heterogeneous sub-
strates: a new approach for hydrolysis of soybean oil using mixtures of bio-
catalysts, RSC Adv. 4, 6863–6868, 2014. https://doi.org/10.1039/c3ra45969a.
213. M. Lee, J. Lee, D. Lee, J. Cho, S. Kim, C. Park, Improvement of enzymatic
biodiesel production by controlled substrate feeding using silica gel in sol-
vent free system, Enzyme Microb. Technol. 49, 402–406, 2011. https://doi.
org/10.1016/j.enzmictec.2011.06.020.
214. D. Šinkuniene, P. Adlercreutz, Effects of regioselectivity and lipid class spec-
ificity of lipases on transesterification, exemplified by biodiesel production,
JAOCS, J. Am. Oil Chem. Soc. 91, 1283–1290, 2014. https://doi.org/10.1007/
s11746-014-2465-7.
56 Biodiesel Technology and Applications

215. Y. Watanabe, T. Nagao, Y. Nishida, Y. Takagi, Y. Shimada, Enzymatic pro-


duction of fatty acid methyl esters by hydrolysis of acid oil followed by ester-
ification, JAOCS, J. Am. Oil Chem. Soc. 84, 1015–1021, 2007. https://doi.
org/10.1007/s11746-007-1143-4.
216. Y. Yan, L. Xu, M. Dai, A synergetic whole-cell biocatalyst for biodiesel pro-
duction, RSC Adv. 2, 6170, 2012. https://doi.org/10.1039/c2ra20974h.
217. A.P. de los Ríos, F.J. Hernández-Fernández, D. Gómez, M. Rubio, F. Tomás-
Alonso, G. Víllora, Understanding the chemical reaction and mass-­transfer
phenomena in a recirculating enzymatic membrane reactor for green
ester synthesis in ionic liquid/supercritical carbon dioxide biphasic sys-
tems, J. Supercrit. Fluids. 43, 303–309, 2007. https://doi.org/10.1016/J.
SUPFLU.2007.06.003.
218. P. Lozano, Enzymes in neoteric solvents: From one-phase to multiphase sys-
tems, Green Chem. 12, 555, 2010. https://doi.org/10.1039/b919088k.
219. L.A. Blanchard, D. Hancu, E.J. Beckman, J.F. Brennecke, Green process-
ing using ionic liquids and CO2, Nature. 399 (1999) 28–29. https://doi.
org/10.1038/19887.
220. P. Lozano, J.M. Bernal, M. Vaultier, Towards continuous sustainable pro-
cesses for enzymatic synthesis of biodiesel in hydrophobic ionic liquids/
supercritical carbon dioxide biphasic systems, Fuel. 90, 3461–3467, 2011.
https://doi.org/10.1016/j.fuel.2011.06.008.
221. D. Yu, C. Wang, Y. Yin, A. Zhang, G. Gao, X. Fang, A synergistic effect of
microwave irradiation and ionic liquids on enzyme-catalyzed biodiesel pro-
duction, Green Chem. 13, 1869, 2011. https://doi.org/10.1039/c1gc15114b.
222. B.M. Nogueira, C. Carretoni, R. Cruz, S. Freitas, P.A. Melo, R. Costa-F??lix,
J.C. Pinto, M. Nele, Microwave activation of enzymatic catalysts for bio-
diesel production, J. Mol. Catal. B Enzym. 67, 117–121, 2010. https://doi.
org/10.1016/j.molcatb.2010.07.015.
223. M.L.B. Queiroz, R.F. Boaventura, M.N. Melo, H.M. Alvarez, C.M.F. Soares,
Á.S. Lima, M.F. Heredia, C. Dariva, A.T. Fricks, Microwave activation of
immobilized lipase for transesterification of vegetable oils, Quim. Nova.,
2015. https://doi.org/10.5935/0100-4042.20150031.
224. Y. Zhang, X. Xia, M. Duan, Y. Han, J. Liu, M. Luo, C. Zhao, Y. Zu, Y. Fu,
Green deep eutectic solvent assisted enzymatic preparation of biodiesel from
yellow horn seed oil with microwave irradiation, J. Mol. Catal. B Enzym. 123,
35–40, 2016. https://doi.org/10.1016/J.MOLCATB.2015.10.013.
225. S. Shah, M. Gupta, The effect of ultrasonic pre-treatment on the catalytic
activity of lipases in aqueous and non-aqueous media, Chem. Cent. J. 2, 1,
2008. https://doi.org/10.1186/1752-153X-2-1.
226. G. Kumar, D. Kumar, Poonam, R. Johari, C.P. Singh, Enzymatic transesterifi-
cation of Jatropha curcas oil assisted by ultrasonication, Ultrason. Sonochem.
18, 923–927, 2011. https://doi.org/10.1016/J.ULTSONCH.2011.03.004.
227. D. Yu, L. Tian, H. Wu, S. Wang, Y. Wang, D. Ma, X. Fang, Ultrasonic
irradiation with vibration for biodiesel production from soybean oil by
Biocatalytic Processes 57

Novozym 435, Process Biochem. 45, 519–525, 2010. https://doi.org/10.1016/J.


PROCBIO.2009.11.012.
228. G. V. Waghmare, V.K. Rathod, Ultrasound assisted enzyme catalyzed hydro-
lysis of waste cooking oil under solvent free condition, Ultrason. Sonochem.
32, 60–67, 2016. https://doi.org/10.1016/j.ultsonch.2016.01.033.
229. C.M. Trentin, A.S. Popiolki, L. Batistella, C.D. Rosa, H. Treichel, D. de
Oliveira, J.V. Oliveira, Enzyme-catalyzed production of biodiesel by
ultrasound-assisted ethanolysis of soybean oil in solvent-free system,
­
Bioprocess Biosyst. Eng. 38, 437–448, 2015. https://doi.org/10.1007/s00449-
014-1316-0.
230. P.B. Subhedar, C. Botelho, A. Ribeiro, R. Castro, M.A. Pereira, P.R.
Gogate, A. Cavaco-Paulo, Ultrasound intensification suppresses the
need of methanol excess during the biodiesel production with Lipozyme
TL-IM, Ultrason. Sonochem. 27, 530–535, 2015. https://doi.org/10.1016/J.
ULTSONCH.2015.04.001.
231. H. Yu, H. Yue, P. Halling, Optimal experimental design for an enzymatic bio-
diesel production system, IFAC-PapersOnLine. 48, 1258–1263, 2015. https://
doi.org/10.1016/j.ifacol.2015.09.141.
232. P. Tufvesson, J. Lima-Ramos, N. Al Haque, K. V. Gernaey, J.M. Woodley,
Advances in the process development of biocatalytic processes, Org. Process
Res. Dev. 17, 1233–1238, 2013. https://doi.org/10.1021/op4001675.
233. Q.K. Beg, V. Sahai, R. Gupta, Statistical media optimization and alkaline pro-
tease production from Bacillus mojavensis in a bioreactor, Process Biochem.
39, 203–209, 2003. https://doi.org/10.1016/S0032-9592(03)00064-5.
234. B. Manohar, S. Divakar, Applications of surface plots and statistical designs
to selected lipase catalysed esterification reactions, Process Biochem. 39, 847–
853, 2004. https://doi.org/10.1016/S0032-9592(03)00192-4.
235. Q.K. Beg, R.K. Saxena, R. Gupta, Kinetic constants determination for an
alkaline protease from Bacillus mojavensis using response surface method-
ology, Biotechnol. Bioeng. 78, 289–295, 2002. https://doi.org/10.1002/bit.
10203.
236. D.S. Rana, K. Thèodore, G.S. Narayana Naidu, T. Panda, Stability and kinet-
ics of β-1,3-glucanse from Trichoderma harzianum, Process Biochem. 39,
149–155, 2003. https://doi.org/10.1016/S0032-9592(02)00323-0.
237. D. Ba, I.H. Boyaci, Modeling and optimization: Usability of response sur-
face methodology, J. Food Eng. 78, 836–845, 2007. https://doi.org/10.1016/j.
jfoodeng.2005.11.024.
238. G.-T. Jeong, D.-H. Park, Response surface methodological approach
for optimization of enzymatic synthesis of sorbitan methacrylate,
Enzyme Microb. Technol. 39, 381–386, 2006. https://doi.org/10.1016/J.
ENZMICTEC.2005.11.046.
239. J.-F. Shaw, H.-Z. Wu, C.-J. Shieh, Optimized enzymatic synthesis of propyl-
ene glycol monolaurate by direct esterification, Food Chem. 81, 91–96, 2003.
https://doi.org/10.1016/S0308-8146(02)00383-7.
58 Biodiesel Technology and Applications

240. H. V. Lee, R. Yunus, J.C. Juan, Y.H. Taufiq-Yap, Process optimization design
for jatropha-based biodiesel production using response surface methodol-
ogy, Fuel Process. Technol. 92, 2420–2428, 2011. https://doi.org/10.1016/j.
fuproc.2011.08.018.
241. C.J. Shieh, H.F. Liao, C.C. Lee, Optimization of lipase-catalyzed biodiesel by
response surface methodology, Bioresour. Technol. 88, 103–106, 2003. https://
doi.org/10.1016/S0960-8524(02)00292-4.
242. S.V. Ghadge, H. Raheman, Process optimization for biodiesel produc-
tion from mahua (Madhuca indica) oil using response surface method-
ology, Bioresour. Technol. 97, 379–384, 2006. https://doi.org/10.1016/J.
BIORTECH.2005.03.014.
2
Application of Low-Frequency
Ultrasound for Intensified
Biodiesel Production Process
Mohd Razealy Anuar, Mohamed Hussein Abdurahman, Nor Irwin Basir
and Ahmad Zuhairi Abdullah*

School of Chemical Engineering, Universiti Sains Malaysia,


Engineering Campus, Nibong Tebal, Penang, Malaysia

Abstract
Biodiesel is an energy source that is mostly derived from plant-based oils through
a transesterification and esterification reactions with alcohols. Due to the immis-
cibility of the reactants, the reaction is only limited to the interface and vigorous
mechanical stirring is often needed to homogenize the reactants. To achieve high
yields (>90%) in about 2 h, temperatures between 50°C and 65°C in the presence
of 0.5 to 3.0 wt. % of catalyst are usually required. The use of ultrasonic energy as
the source of mixing has become a great interest as it can increase the interface
area leading to higher biodiesel yield. During ultrasonication, the cavitation bub-
bles will be formed and subsequently collapse asymmetrically once the critical
sizes are reached. The boundary of the reacting phases will be disrupted by the
micro jets that result from these collapses. It will also result in a temperature rise
at the locality of the phase boundary. The use of ultrasonic energy in conjunction
with a heterogeneous catalyst can lead to shorter reaction times of 10–40 min with
two to three times lower amount of catalyst needed as compared to those expected
in normal mechanically stirred reactor. Thus, the process will require shorter reac-
tion time, lower catalyst amount and lower energy input to result in high methyl
ester yield.

Keywords: Biodiesel, ultrasound, mass transfer, immiscible reactant,


transesterification, process intensification, catalyst

*Corresponding author: chzuhairi@usm.my (https://orcid.org/0000-0001-6394-2917)

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (59–84) © 2021 Scrivener Publishing LLC

59
60 Biodiesel Technology and Applications

2.1 Current Fossil Fuel Scenario


Recently, the increasing demand on the unsustainable fossil fuels is a
challenging topic to be addressed. With the current rate of crude oil
consumption, the conventional energy sources could be depleted not
even 50 years down the line [1]. The high usage of such fuels to gen-
erate energy has also caused associated environmental effects such as
global warming [2], greenhouse gas emissions, and acid rain [3] and
other environmental pollutions such as particulate matters [4]. It is
also reported that carbon monoxide (CO), carbon dioxide (CO2), sul-
fur oxides (SOx), and nitrogen oxides (NOx) that result from fossil fuels
combustion are the key culprits for global warming [1]. It was antic-
ipated that the world’s overall energy consumption and thus the CO2
emissions to the atmosphere could show 60% increase in 2018 as com-
pared to 2008. Furthermore, oil and gas prices had also been increasing
in the recent years [5].
Due to the above-mentioned issues, the necessity of finding the envi-
ronmental-benign alternatives to fossil fuels to address the future energy
demand is a worthwhile effort [6, 7]. Alternatives to replace fossil fuels are
such as biofuels, wind, and solar energy [2]. The dependence of transpor-
tation sector on fossil fuels was about 96% of the current energy consumed
in the sector [5]. The current scenario of soaring fossil fuel prices and their
finite reserves therefore justify the quest for biofuels as the fifth energy
sources.

2.2 Biodiesel
Recently, methyl esters of fatty acids in vegetable oils, also known as
biodiesel, have been demonstrated to be effectively used to power die-
sel engines at certain blend ratios [8, 9]. Biodiesel fuel is a result of the
alcohol transesterification of triglycerides in vegetable oils or animal fats.
Alternatively, free fatty acids in those triglycerides can also be esterified
to yield methyl esters. This fuel has been demonstrated to show potentials
as a renewable fuel source based on several favorable specifications [6].
It consists of monoalkyl esters and produced through a reaction between
the triglycerides and a monohydric alcohol [10]. Biodiesel as a fuel prom-
ises the advantages of (1) minimum dependence on fossil, (2) renewability,
(3) positive energy balance, (4) better emission quality, (5) non-toxic, (6)
agricultural surplus as feedstock, and (7) better safety in handling [11].
Ultrasound-Assisted Biodiesel Production 61

The properties of biodiesel in strongly influenced by the characteristics


of feedstock as well as the production system used. There are several meth-
ods of biodiesel production reported such as homogeneous acid-mediated
process [12], supercritical fluid process [3], enzymatic process [13], and
homogenous base catalyzed process [14]. Base catalysts are usually more
favorable than acid catalysts or enzymes [15]. Research efforts toward
future use of heterogeneous catalysts for this conversion are also actively
investigated [16, 17].

2.3 Transesterification
Biodiesel is commonly produced by the catalytic reaction between tri-
glycerides (oils or fats) with simple alcohols (i.e., methanol, ethanol,
propanol and butanol). Glycerol is co-produced as the side-product.
The transesterification process is as schematically shown in Figure 2.1.
Stoichiometrically, the reaction involved 3 moles of alcohol and 1 mole
of triglyceride to yield 3 moles of methyl esters (biodiesel) and 1 mole of
glycerol. However, due to the reversible nature of this reaction, in prac-
tice, excess of alcohol is used to push the equilibrium toward favorable side
to achieve higher ester yield [9]. In this reaction, diglycerides and mono-
glycerides are undesired products that should be minimized for the sake of
better FAME yield.
Esters are the desired products of the transesterification reactions. At
the end of the reaction, glycerol is removed by centrifugation or settling

O O
CH2 O C Rx CH2 O C Rx

O CH2 O H CH2 O H
CH2 O C R1 O CH2 O H R1COOCH3
O CH2 O C Rx CH2 O H
CH2 O C R2 + 3 CH3OH CH2 O H + R2COOCH3
O CH2 O H
CH2 O H
CH2 O C R3 O O CH2 O H R3COOCH3
CH2 O C Rx CH2 O C Rx
O
Note : x=1, 2 or 3
CH2 O C Rx CH2 O H

Triglycerides Methanol Diglycerides Monoglycerides Glycerol Methyl esters


(Biodiesel)

Figure 2.1 Transesterification of triglycerides for the production of biodiesel.


62 Biodiesel Technology and Applications

for subsequent purification. Purified glycerol can be used in many con-


ventional applications (pharmaceutical, cosmetic, and food industries).
Newly developed applications for glycerol are also found in the fields of
animal feed, surfactants, polymers, lubricants, etc. [18, 19]. Glycerol recov-
ery is also crucial because of its several applications in different industrial
processes.

2.4 Challenges for Improved Biodiesel Production


Conventional biodiesel production processes are still more expensive as
compared to those of fossil diesel. The cost of the biodiesel as compared
to diesel does not make it favorable alternative to fossil fuels. One of the
important issues which can directly reduce the cost is the reduction in
the production costs. These expenses are such as the heating, labor work,
mixing, and raw material use. On the other hand, generally, biodiesel is
produced from various vegetable oils that have undergone extensive puri-
fication processes. These oils are relatively costly and at the same time more
suitable for edible purposes. There has been heated debate on the fuel ver-
sus food dilemma across the globe [20].
Homogeneous catalyst has many environmental impacts due to produc-
tion of wastewater during the separation process. Alcohol and oil are also
immiscible and the reaction only takes place in the interface of the reac-
tants. Thus, the production of biodiesel requires a rather long reaction time
and excess of catalyst to assure high yields. These conditions can increase
the overall biodiesel production costs. Therefore, as an ultimate objective,
alternative method to decrease the process costs and time should be iden-
tified. Long reaction time will directly and indirectly affect the produc-
tion cost of biodiesel while affecting it quality at the same time. Effective
transesterification process usually occurs at temperatures higher than
60°C. Lower temperatures often result in poor reaction rates while higher
temperatures will require a pressurized reactor considering the methanol
boiling point of 64.8°C. Thus, reduction in the reaction time can cause less
time of energy consumption for heating the system. This highlights the
importance of finding technologies which are less costly and less energy
consumption. Time reduction can also reduce the labor work for produc-
tion of the same amount of biodiesel.
Biodiesel also needs to fulfill the strict requirements of ASTM D6751-17
or EN 14212:2008 standards to be used as fuel. The consistent compliance
is an uphill task. In order to produce biodiesel that consistently meet the
technical standards, strict quality control measures for the feedstock and
Ultrasound-Assisted Biodiesel Production 63

the catalyst should be implemented. However, the severity of the produc-


tion process should be reduced bearing in mind the low stability of the
triglycerides used in the production process [21]. This technology should
also have the feasibility to be scaled up. Thus, the reaction rates and opti-
mized operating conditions are also of great concerns in this study.

2.5 Homogeneous Catalyst for Biodiesel Production


Biodiesel production process is usually catalyzed by enzymes, acidic or
basic catalysts. Sulfuric acid and hydrochloric acids are conventionally
used for the acid catalyzed processes. These catalysts are often used for
oils feedstock with high water and free fatty acid (FFA) contents. Acid-
catalyzed biodiesel production processes normally need longer reaction
time and higher alcohol to oil ratio [17]. As such, the production costs are
usually higher in this case. As such, biodiesel is more commonly produced
now from refined/edible type oils (with lower FFA and water contents)
in the presence of an alkaline catalyst. This reaction is relatively higher
to make this option more favorable. However, the difficulty of alkali-
catalyzed esterification of non-edible oils is excessive soap formation which
inhibits the effective separation of ester and glycerin. This phenomenon is
attributed to the residual amounts of FFA that present in the oil, particu-
larly in the case of non-edible/not refined oils [16]. For an alkali-catalyzed
transesterification process, the glycerides and alcohol must be substantially
anhydrous. In the presence of water, the undesired saponification reaction
between the catalysts with FFA will produce metal salts of the fatty acids
(soaps) [21].
There is also certain environmental impact associated with the catalyst
amount used. Excessive amounts will result in large amount of soap which
is undesirable as far as product separation and biodiesel quality are con-
cerned. Besides, the remaining catalyst can end up in the product leading to
increased biodiesel pH other biodiesel quality parameters such as ash con-
tent [2]. The typical catalyst used for the production of biodiesel is homo-
geneous base catalysts such as KOH and NaOH. This process can produce
high methyl ester yield under atmospheric pressure and temperature of
50°C–65°C which is usually carried out in 1–4 h [6]. In order to improve
the biodiesel quality to meet certain intentional standards, downstream
purification of the methyl ester is required prior to its use as biodiesel. This
could be achieved by removing impurities such as catalyst and glycerol
by hot water washing. It should be noted that such further purification
processes generate enormous amounts of wastewater [15]. Thus, replacing
64 Biodiesel Technology and Applications

conventional homogeneous catalysts with environmentally friendly het-


erogeneous catalysts seems to be essential for the benefit of lower produc-
tion costs and environmental considerations.

2.6 Heterogeneous Catalyst for Biodiesel Production


Due to environmental concerns, the interest in biodiesel production is cur-
rently been directed toward heterogeneous process. This approach allows
catalyst removal after the reaction without extensive downstream purifica-
tion processes. Other advantages are in the reusability potential of the cat-
alyst and the less corrosive character of heterogeneous catalysts [22]. They
can also be formulated to exhibit better activity, selectivity and stability
[13]. Solid catalyst can also be divided into two groups, i.e., acid catalysts
and base catalysts. It has been reported that acidic heterogeneous catalysts
usually show weak activity and react at relatively higher temperature with
longer reaction time as compared to basic heterogeneous catalysts [23].
Various types of heterogeneous catalysts for transesterification process
have been reported by several researchers. Alkaline and alkaline-earth
metal compounds were recently used in biodiesel production to replace
conventional homogenous catalysts [5]. Alkaline and alkaline-earth metals
supported on other materials such as alumina and silica can be consid-
ered as potential catalysts for the transesterification reaction. Zeolites are
another group of catalysts studied for the transesterification process [24].
Clay minerals such as hydrotalic, chrysotile, and sepiolite are considered to
be the next group. These minerals belong to the group of modest or weakly
basic catalysts.
In the recent years, heterogeneous base catalysts are gaining popularity
for biodiesel production due to their benefits. Yet, there are still a lot more
to understand about the natures of these catalysts. Clarifying the nature of
the interaction between the support and the catalytically active component
in the process is still worthy of further investigation. In this new process,
the catalyst is expected to promote the transesterification reaction without
experiencing catalyst loss. Yet, it is known that solid base catalysts can be
easily poisoned by water or carbon dioxide. Poisoning is a phenomenon
responsible for loss of catalytic activity that stems from strong chemisorp-
tion of certain substances on the active sites in the reaction system. When
solid bases are used in liquid phase organic reactions, the reaction rate
may decrease due to the diffusion of reactants and products. One of the
main drawbacks of heterogeneous catalysts is lower catalytic activity as
compared to that of homogeneous ones such as KOH [6]. Consequently,
Ultrasound-Assisted Biodiesel Production 65

higher reaction temperature and pressure than those in homogeneous pro-


cesses are required with an excess of methanol. This will certainly result in
higher production costs. Therefore, more research endeavor is required to
make this technology more economically interesting and environmentally
favorable.

2.7 Immiscibility of the Reactants


Alcohol and triglycerides as the reactants in the transesterification process
are mutually immiscible. This reaction is therefore a mass transfer-limited
one due to their immiscibility [25]. The rate of the reaction is very much
dependent on the mixing ability of the reactants providing a better contact
or interface area. The reason for poor rates of reaction of transesterification
process is the mass transfer limitations between the reactants that often
create immiscible phases in the reactor [26]. To overcome this problem,
researchers have recently introduced new methods to make rigorous mix-
ing in the system. One of the techniques is the use of cavitation which
results in conditions characterized by local intense turbulence. These effects
are capable of improving the rates of chemical reactions which are limited
by mass transfer resistances by increasing the interfacial area between the
two phases [27].
Ultrasonic irradiation is one of the approaches that have attracted much
attention in this respect to overcome the poor immiscibility of the reac-
tants [28, 29]. Ultrasound is defined as any sound pressure with a fre-
quency above the upper limit of hearing by human. As a matter of fact,
the upper limit varies depending on individual. However, about 20 kHz
has been established as the upper limit for healthy, young adults. As such,
this level usually serves as a lower limit for the definition of an ultrasound.
Figure 2.2 shows the approximate frequency ranges of sounds with rough
guide of some sources of the sound and common applications [30].
Sonochemical reacting system is an innovative approach in which ultra-
sonic wave is used to create an oxidative environment by means of cavi-
tation bubbles that are formed during the rarefaction period of the high
frequency sound. The creation, propagation, and implosion of cavitation
bubbles caused by the sonic waves will cause high local temperature and
pressure rises. These rises are considered the responsible mechanism for
the acceleration of chemical reactions in a sonochemical system. This
leads to the increase in the reaction rate and shortened the time required
for complete reaction or to reach equilibrium [31]. In an aqueous system,
three different regions have been postulated for the reaction to occur, i.e.,
66 Biodiesel Technology and Applications

20 kHz
Infrasonic Sonic Utrasonic Hypersonic

Hz
0 10 102 103 104 105 106 107 108 109 1010

Human hearing Plastic welding


Cleaning
Aircraft

Earthquake Dog Sonochemistry

Violin Bat Chemical analysis


Medical diagnostic
Dolphin Non-destructive testing

Figure 2.2 Approximate frequency ranges of sound with rough guide of some sources or
applications.

(i) the bulk solution, (ii) the interface between the bulk solution and the
cavitation bubbles, and (iii) the gaseous interiors of cavitation bubbles [32].

2.8 Ultrasound-Assisted Biodiesel Production Process


2.8.1 Fundamental Aspects of the Process
Lately, the potential of ultrasonic energy in biodiesel production process
has attracted the attention of many researchers [2, 27]. Because of the
poor immiscibility of a liquid-liquid heterogeneous system, the interaction
between alcohol and oil is just limited to the interface of the two phases
[9]. Vigorous mechanical stirring is conventionally used to homogenize
the reactants to enhance the interfacial area between the two phases [5].
In a conventional stirring method to achieve high yields (>90%), a tem-
perature of 50°C–65°C in presence of catalyst 0.5%–3.0% and the reaction
should also be extended to 2 h or more [6]. Kalva et al. [9] concluded that
the rate of reaction was influenced by the interfacial contact between the
two reacting phases. In addition, the creation of alcohol anions would ini-
tiate the methyl ester formation reaction.
The use of ultrasonic energy as the source of mixing has become a great
interest by two different ways. It will intensively create a virtually homoge-
neous mixture and at the same time, it can create heating effect to increase
the biodiesel production yield. The application of ultrasound in biodiesel
production process can be done in several options. It can be directly used
to mix the immiscible reactants while they are fed into a continuously flow
Ultrasound-Assisted Biodiesel Production 67

reactor as marked as Option 1 in Figure 2.3. This approach is highly effec-


tive as direct contact between the ultrasonic transducer with the reacting
mixture is allowed. Then, the desired effects of the ultrasonication are
passed directly to the reacting mixture. Ultrasonic cavitation will also
result in a localized temperature increase in the phase boundary leading to
enhanced reaction rate. However, it also causes unnecessary stress to the
ultrasonic system as high temperature could disrupt its normal operation.
In addition, metal transducer might also not so stable to the prolonged
exposure to acidic condition at high temperature. This is especially true in
the case of acidic catalysts used to process high FFA content feedstock [33].
Alternatively, ultrasonic processor can also be used to create micro-
droplets (~50 microns in diameter) of the alcohol and pre-mix them with
the oil to create a homogeneous dispersion to be fed into the reactor as given
in Option 2 in Figure 2.3. Ultrasonic irradiation has been demonstrated to

Ultrasonicator

Oil Methanol

Option 1

Ultrasonicator

Product
Separation

Option 2

Product
Separation

Figure 2.3 Different options of ultrasound application for biodiesel production.


68 Biodiesel Technology and Applications

be beneficial in accelerating an immiscible reaction system through the


generation of cavitation bubbles [2]. The disruption of the phase boundary
will then result from the asymmetric collapses of the cavitation bubbles to
create micro jets of the liquid. In turn, this will lead to the mixing effect
of the system, especially at the interphase. Consequently, the stable emul-
sion mixture is created to allow accelerated reaction rate in the reactor. It
has been reported that the use ultrasound in conjunction with homoge-
neous catalyst in the biodiesel process can lead to a shorter reaction times
of 10–40 min with two to three times lower amount of catalyst needed as
compared to those in normal mechanical stirring reactor [34].
The main principle of ultrasound that leads to an efficient mixing of
different viscosity reactants is the creation of cavitation bubbles at the
interfacial area between molecules of reactants [35]. The molecules would
experience a significant vibration which causes the weakening of the
attraction force to create gap between liquid. At an ultrasonic wave of >20
kHz, microscopic and macroscopic bubbles will be formed, and they are
beneficial in enhancing the mass transfer limitation [36]. The increased
vibration of the bubbles at the boundary layer eventually produces acoustic
cavitation that increases the mixing rate of the liquid reactants [37, 38].
The cavitation bubbles would then collapse due to vigorous breakdown
when exceeding the limitation point. This might generate localized pressure
increase to as high as 2,000 atm with the internal temperature approaching
5,000 K, thus initiating the transesterification and esterification reactions
[39]. The intense internal heat and pressure generation would extensively
provide severe mechanical agitation enticing a homogeneous mixing
between oil and alcohol. Thus, the immiscibility and emulsification lim-
itation would be minimized. Overall, the principle of ultrasound acoustic
cavitation bubbles is summarized in Figure 2.4.
Kalva et al. [9] concluded the desired effects of ultrasound-assisted bio-
diesel production system based on chemical and physical phenomena.
The chemical reaction would be accelerated by transient implosive of the
bubbles collapse and a homogenize mixture would formed due to radial
motion of the bubbles that can generate micro-turbulence. The collapse of
the bubbles can cause transient implosive that would eventually produce
H+ and OH− ions that will sequentially accelerate the chemical reaction.
Ultimately, the production of biodiesel will technically and economically
undergo a massive improvement with the intensification of ultrasound
irradiation processing with advanced catalytic system. With recent studies
on various types of catalytic system based on the use of ultrasound, a bio-
diesel process with an accelerated reaction rate is foreseeable in the near
future [36, 40].
Ultrasound-Assisted Biodiesel Production 69

Ultrasonication -Shockwaves
>20kHz -Microjets
-4,000++ K hot spots
-1,000++ atm

Emulsified
mixture

Bubble Bubble Critical Bubble Bubble


Formation Growth Size Implosion
Transesterification
reaction

Alcohol

Triglyceride Fatty acid methyl esters

Figure 2.4 Principle of ultrasound-induced acoustic cavitation bubbles.

2.8.2 Homogeneously Catalyzed Ultrasound-Assisted System


Studies on ultrasound-assisted biodiesel production process started with
the application of homogeneous catalysis system. Both acid and base
homogeneous catalysts have been used in previous works that are tabulated
in Table 2.1. Sulfuric acid is the most often used acid homogeneous catalyst
in homogeneous biodiesel production [46]. It is highly active during the
reaction and gives FAME results of up to 99% [41, 42]. In conventional
mechanical stirring method, acid homogeneous catalyst is 4,000 times
slower than base homogeneous catalyst [47]. However, it is highly bene-
ficial in achieving high FAME yield production from high FFA feedstock.
Acid catalyst is capable of handling FFA and triglycerides by conducting
simultaneous esterification and transesterification reactions [46]. Thus,
high FAME yield with high purity product is expected to be produced.
This is main attributed to milder reaction conditions needed to achieve
significantly high reaction rates. However, the resistance of the ultrasonic
transducer that has prolonged direct contact with the acidic medium could
be the drawback of the ultrasound-assisted process.
Mostly, basic homogeneous catalytic processes were carried out by
using alkaline metal hydroxides [42, 44]. The transesterification reaction
could occur at a sufficiently fast rate and the FAME yield might achieve
nearly 100% [42]. Alkaline earth metal hydroxides are also available at
low cost. However, they are also sensitive to the presence of FFA in feed-
stock. Saponification reaction would take place and impedes the formation
70

Table 2.1 Performance of ultrasound-assisted homogeneous biodiesel production systems.


Reaction
conditions
Alcohol to oil Time Frequency Biodiesel
Catalyst Feedstock ratio T (°C) Wcat (%) (min) (kHz) yield (%) Reference
Sulfuric acid, Soybean oil 9:1 (Methanol) 60 1 60 24 90 [41]
H2SO4
Sulfuric acid, FFA from 3:1–9:1 25 1 90 40 98 [42]
H2SO4 Oreochromis (Methanol)
niloticus
p-toluenesulfonic Rapeseed oil 3.3:1 45 1 180 20 12 [43]
acid (Methanol)
Biodiesel Technology and Applications

Pottasium Silybum 4:1–12:1 30–80 0.25–2 5–70 40 96 [44]


hydroxide, marianum (methanol
KOH oil and
ethanol)
Pottasium Soybean oil 9:1 60 1 60 24 99 [41]
hydroxide,
KOH
(Continued)
Table 2.1 Performance of ultrasound-assisted homogeneous biodiesel production systems. (Continued)
Reaction
conditions
Alcohol to oil Time Frequency Biodiesel
Catalyst Feedstock ratio T (°C) Wcat (%) (min) (kHz) yield (%) Reference
Sodium Soybean oil 3:1–9:1 29 1 30 40 100 [42]
hydroxide, (Methanol)
NaOH
NaOCH3 Peanut oil 3:1-9:1 25 1 60 20 90 [28]
(Methanol)
C2H5ONa Fish oil 6:1 (ethanol) 20–60 1 30 20 >98 [45]
Ultrasound-Assisted Biodiesel Production
71
72 Biodiesel Technology and Applications

of ester. Thus, the activity of the catalyst might decrease and poor reac-
tion yield will result. Besides alkaline earth metal hydroxides, some other
catalysts have also been reported as base homogeneous catalysts such as
NaOCH3 and C2H5ONa [28, 45]. The use of these catalysts usually showed
tremendous reaction yields of up to 98%.
Unfortunately, the use of homogeneous catalyst has many drawbacks
especially with regards to the product quality problem and environmen-
tal concerns. It also requires an additional complicated product separation
unit to obtain high purity product. To avoid such problem, another alter-
native has been explored in which heterogeneous catalysts would be the
best candidates to replace the role of homogeneous catalysts in biodiesel
production.

2.8.3 Heterogeneously Catalyzed Ultrasound-Assisted System


2.8.3.1 Heterogeneously Acid Catalyzed System
Some works focusing investigating the potential of solid acid catalyst in
an ultrasound-assisted biodiesel production are tabulated in Table 2.2.
Heteropolyacid, i.e., tungstophosporic acid (TPA) was used by Badday
et al. [33, 37, 48] to create acidic sites in different kinds of support includ-
ing gamma alumina, cesium, and activated carbon. The use of 25% TPA
loading in gamma alumina resulted in the FAME yield of about 64.3%.
After the reaction conditions were mathematically optimized (60% ultra-
sonic amplitude at 20 kHz, a methanol:oil ratio of 19:1, 65oC, 60 min), the
FAME yield successfully increased to 84%. Higher TPA loadings (above
25%) reduced the FAME yield and it was attributed to partial blockage of
catalytic pores by the acidic component that eventually reduced the acces-
sibility of reactants into pore sites. Small leaching of TPA was unfortunately
reported owing to the partial dissolution of the catalyst in the polar sub-
stances in the mixture and it was accelerated by the ultrasonic effect during
the reaction. Significantly higher yield was reported by Maneechakr et al.
[49] with higher ratio of alcohol to oil.
Next, a heteropolyacid-doped cesium (Cs) was used in the ultrasound-as-
sisted biodiesel production [37]. The effect of different Cs exchange levels
was investigated with the highest FAME yield obtained of about 81.3%. The
reaction was reported to be insensitive to the addition of FFA as the acid
sites in the catalyst dominantly converted FFA directly to FAME. However,
high content of FFA caused the reduction of FAME yield as water that was
produced from the esterification of FFA could also allow acid hydrolysis
Table 2.2 Ultrasound-assisted biodiesel production processes catalyzed by heterogeneous acid catalyst.
Reaction
conditions
Alcohol to Time Frequency Biodiesel
Catalyst Feedstock oil ratio T (°C) Wcat (%) (min) (kHz) yield (%) Reference
Heteropoly acid Jathropa oil 20:1 56 4 40–60 20 87.33 [33]
supported
on activated
carbon
Heteropoly acid Jathropa oil 5:1–25:1 65 2.5–4.5 10–50 20 64.3-84 [48]
supported on
alumina

Caesium doped Jathropa oil 5:1–25:1 65 2.5–4.5 10–50 20 81.3 [37]


heteropoly acid

Sulfonated carbon Waste cooking 20:1–40:1 50-150 5–15 2–14 25 90.8 [49]
oil

Tungstated Oleaginous 30:1–60:1 40-60 2–4 20 22.5 71% [50]


zirconia (WO3/ scenedesmus
ZrO2) sp.
Ultrasound-Assisted Biodiesel Production
73
74 Biodiesel Technology and Applications

of the ester. High porosity also caused the catalyst to become unstable and
eventually deactivated the catalyst.
Maneechakr et al. [49] attempted the use of carbon as the support for
acid heterogeneous catalyst in an ultrasound-assisted biodiesel produc-
tion. The carbon was previously sulfonated by contacting with H2SO4 and
it resulted in FAME yields of up to 90.8% under optimum conditions. The
sulfonated catalyst was very active, and it showed low activation energy
of 11.64 kJ/mol. However, the catalyst showed a significant reduction in
FAME yield after being used in successive reactions (12 cycles) without
regeneration with H2SO4. The reduction might due to the deactivation
caused by the deposition of polymerized products and appreciable acid
loss that might occur during the reaction. The disappearance of Na and K
elements after the fourth cycle was also reported. A successful work toward
activity improvement of phosphotangestic acid as the active catalytic com-
ponent has been demonstrated by Nikseresht et al. [51].
Guldhe et al. [50] highlighted the role of tungstated zirconia in produc-
ing biodiesel from Oleaginous scenedesmus sp oil. The extraction of lipid
was carried out in a single step together with transesterification to bio-
diesel under ultrasound system. In such system, ultrasonic effect was used
to extract lipid and a catalyst was used for converting lipid to FAME. The
use of acid catalyst often subjects to some difficulties in which the acid
strength could affect the activity of the catalyst and it was unlikely to cre-
ate a uniform porous catalyst [46]. As such, another alternative especially
basic heterogeneous catalyst is worth investigation for use to replace acidic
heterogeneous catalyst for biodiesel production.

2.8.3.2 Heterogeneous-Based Catalyzed Ultrasound-Assisted System


Recent biodiesel research works utilizing ultrasound-assisted systems cat-
alyzed by various heterogeneous base catalysts are tabulated in Table 2.3.
In their works, Mootabadi et al. [52] and Chen et al. [54] highlighted the
effectiveness of oxides of alkaline earth metals in converting palm oil to
biodiesel. CaO showed the highest activity with yields of up to 95% in 1 h.
Clearly, the alcohol to oil ratio requirement was lower than those com-
monly reported for acid-catalyzed biodiesel production.
Heterogeneous base catalysts have also shown their applicability for bio-
diesel production with a wide range of feedstock. These include rarely found
feedstock such as Silybum marianum oil [55] and sesame oil (Sesamum
indicum L.) [56]. Malani et al. [40] even investigated mixed non-edible oil.
In the study on biodiesel production from Silybum marianum oil, a series
of TiO2 doped C4H4O6HK catalysts differed by the catalyst preparation
Table 2.3 Performance of ultrasound-assisted heterogeneous base-catalyzed processes for biodiesel production.

Reaction condition
Alcohol to Time Frequency Biodiesel
Catalyst Feedstock oil ratio T (°C) Wcat (%) (min) (kHz) yield (%) Reference
CaO, SrO, BaO Palm oil 3:1-15:1 65 3 60 20 95.0 [52]

Calcium Used cooking 6:1–14:1 45–65 0.5–1.25 0–40 20 93.5 [53]


diglyceroxide oil
CaO Palm oil 3:1–15:1 60 3–10 60–180 20 92.7 [54]
TiO2 doped with Silybum 8:1–18:1 40–70 1–9 10-40 40 90.1 [55]
C4H4O6HK marianum
oil
Commercial Waste 4:1–8:1 30–60 1–4 0–120 50 92.0 [20]
potassium cooking oil
and sodium
phosphate
Barium hydroxide Sesame 4.5:1–7.5:1 25–35 1–2 20–40 20 98.6 [56]
(Sesamum
indicum
L.) oil
Ultrasound-Assisted Biodiesel Production

(Continued)
75
76

Table 2.3 Performance of ultrasound-assisted heterogeneous base-catalyzed processes for biodiesel production. (Continued)

Reaction condition
Alcohol to Time Frequency Biodiesel
Catalyst Feedstock oil ratio T (°C) Wcat (%) (min) (kHz) yield (%) Reference
Coal fly ash Used cooking 4:1–12:1 70 3–11 0.5–2.5 20 95.6 [57]
oil
Magnetic Jatropha and 5:1–7:1 48–8 1–9 20–100 20–25 Soybean: [58]
soybean oil 97.9
Biodiesel Technology and Applications

(Na2SiO3@
Fe3O4/C) Jathropa:
94.7
Ultrasound-Assisted Biodiesel Production 77

procedure were investigated [55]. In this study, TiO2 impregnated with


potassium bitartrate and calcined for 6 h at a temperature of 600°C showed
the highest activity by producing 90.1% of biodiesel yield. Barium hydrox-
ide was used in a similar reaction system to convert sesame oil into bio-
diesel [56]. In the optimization study performed using an artificial neural
network (ANN) approach, it was found that the catalyst loading was the
most influential parameter that significantly determined the FAME yield.
The maximum FAME yield recorded was 98.6%.
Pukale et al. [20] and Gupta et al. [53] investigated the production of
biodiesel with waste cooking oil as the feedstock and catalyzed by com-
mercial PO4 and calcium diglyceroxide, respectively. By using various
commercial PO4 catalysts doped with alkaline metals like potassium and
sodium, waste cooking oil was successfully converted to biodiesel with
K3PO4 demonstrated the best catalytic performance. The high catalytic
activity was retained until the fourth cycle and it only started to drop in the
fifth to the eighth cycles, observed as the gradual reduction in the FAME
yield to 65%. They also demonstrated significant leaching of the catalyst
during the reaction. Meanwhile, the use of calcium diglyceroxide was also
beneficial in biodiesel production with a maximum yield was 93.5% using
an ultrasound-assisted system [53]. However, the catalyst also experienced
a severe leaching problem and the FAME yield sharply reduced to 50.5% in
the third cycle. This might due to the partial solubility of calcium diglycer-
oxide in glycerol-methanol mixture in the successive reaction cycles.
Heterogeneous base catalysts have also been prepared using coal fly
ash as reported by Xiang et al. [57]. With nowadays requirement for the
advanced catalysis system for biodiesel production, such a material could
potentially conduct good transesterification reaction with low quality
feedstock such as used cooking oil. Coal fly ash possibly contains some
amount of silicon dioxide, calcium oxide and magnesium oxide. The opti-
mal FAME yield could achieve 95.57% under the right process conditions.
Excellent reusability potential was also demonstrated by the catalyst in
which it could withstand up to eight reaction cycles under ultrasonic con-
dition with only minimal reduction in the catalytic activity.
Zhang et al. [58] came up with a unique heterogeneous catalysis sys-
tem by combining ultrasound with magnetic stirring. Ultrasound served
to create emulsified mixture between the reactants while the conven-
tional stirring’s role was to ensure the emulsified mixture was distributed
to the whole reacting volume. The removal of the catalyst was achieved
by means of magnetic method. Soybean oil and Jathropa curcas oils were
tested and both feedstocks demonstrated high FAME yields of 97.9% and
94.7%, respectively. The combination of magnetic stirring and ultrasound
78 Biodiesel Technology and Applications

successfully improved the catalytic activity and prevented the reduction in


the catalytic activity after several successive reaction cycles.

2.8.3.3 Influence of Reaction Parameters


In an ultrasound-assisted system, the FAME yield can be influenced by a
few reaction parameters such as the ratio of alcohol to oil, reaction time,
reaction temperature, catalyst amount, and ultrasonic amplitude. These
parameters should be taken into consideration to run the reaction at its
optimum conditions to obtain highest possible FAME yield. Based on
mechanism of transesterification reaction, alkoxide ion produced from
alcohol is required to initiate the reaction. Lower alcohols (methanol, eth-
anol, etc.) are usually preferred for biodiesel production [59]. The alcohol
amount used should be stoichiometrically higher than oil to push the equi-
librium forward to form more FAME as the oil transesterification reaction
is in fact a reversible reaction [20, 53]. In excess alcohol condition, ade-
quate production of alkoxide ion is ensured and high reaction rate will
be achieved [55]. However, higher amount of alcohol might significantly
dilute the product leading to an initiation of reversed reaction to decrease
the FAME yield [20]. In this respect, intensified reaction with the use of
ultrasound could lower the alcohol to oil requirement to the benefit of the
high biodiesel yield.
Reaction temperature is another important factor influencing biodiesel
production. In an ultrasound-assisted system, only low reaction tempera-
ture is needed as the cavitation effect will also produce internal heating
effect during the reaction. An increase in the reaction temperature would
facilitate the solubility and miscibility of oil in alcohol and increase the dif-
fusivity of three-phase reaction mixture (oil, alcohol, and catalyst) [20, 53,
55]. Besides, it also increases the kinetic energy of molecules that will sig-
nificantly increase the collision between reactant molecules hence increas-
ing the reaction yield [53]. However, higher reaction temperature also
leads to the supersaturated formation of vaporized alcohol bubbles [20].
It causes cushioning of the collapse of bubbles resulting in low implosion
intensity and eventually decreases the mass transfer [53, 60].
Basically, ultrasound-assisted system effectively increases the rate of
biodiesel production through cavitation effect; hence, only short reaction
time is needed to achieve high reaction yield. Additional time is needed to
ensure the reversible reaction can achieve its equilibrium. In the first 10
min, the reaction was observed to start with slow reaction rate due to inad-
equate agitation effect and reactant mixture was not homogeneously mixed
yet [55]. Thus, sufficient time was required to complete the conversion
Ultrasound-Assisted Biodiesel Production 79

of triglycerides to yield FAME. However, the reaction needed is usually


significantly shorter than that required by conventionally stirred reactor
system.
In a biodiesel production reactor system, the amount of catalyst used in
biodiesel production should be correctly determined. The ideal amount of
catalyst might result in high FAME yield and minimizes wastage of pricy
catalyst. At the same time, it contributed to lower impurity level that could
be possibly introduced into the biodiesel product. Takase et al. [44] high-
lighted that less amount of catalyst would beneficially increase the FAME
yield while excess of catalyst might cause more saponification reaction to
occur to form excessive formation of soap. In addition, higher amount
of catalyst often leads to the formation of viscous reactant mixture that
requires higher energy to provide adequate stirring effect. Besides, it might
also give a negative impact of energy dissipation due to scattering on sound
wave [53]. With the intensified reaction rate made possible with the use
of ultrasound, same level of biodiesel yield will be achievable with signifi-
cantly less amount of catalyst. The impact will be significant on the per-
spectives of process economy, safety, and product quality.

2.9 Conclusions
Biodiesel is the most potential biofuel to replace petroleum diesel.
Chemically, it is made up of fatty acid methyl esters that are derived mainly
from many vegetable oils through either a transesterification or an esteri-
fication reaction with alcohols, particularly methanol. Slow reaction bio-
diesel production process is associated with rather small interfacial area
between the two immiscible reactants. Heterogeneous catalysts are greener
alternatives to conventionally homogeneous ones but they are often sub-
ject to low specific surface area for sufficient interaction with reactants and
poor internal mass transfer, leading to poor rate of reaction. Ultrasound-
assisted biodiesel production can be potentially used to enhance the
transesterification and/or esterification reactions with aid of advanced het-
erogeneous catalytic system. With the cavitation effect provided by ultra-
sonic irradiation, the emulsification and mass transfer limitation between
the two immiscibility reactants could be effectively minimized. Improved
internal mass transfer and hot spots generated during the ultrasonication
will allow high methyl ester to be made possible under milder process
conditions. Hence, the transesterification reaction might reach its equilib-
rium in shorter reaction time while lower catalyst loading and alcohol to
oil ratio are needed. Many aspects of this new process are yet to be fully
80 Biodiesel Technology and Applications

investigated such as behavior of the ultrasonic-mediated process, the roles


of ultrasound on the reactants and the catalysts, effects on biodiesel quality,
optimization of the reaction based on different aspects, and also the kinetic
modeling of the reaction. Influence of ultrasonication to the reaction in
terms of reaction temperature requirement, product separation and purifi-
cation, and methyl ester quality also deserves further research attention. It
is also necessary to evaluate the biodiesel quality parameters to make sure
that ultrasonic cavitations have no adverse effect on the biofuel quality to
make it not suitable engine use.

Acknowledgement
This oleochemical research work is made possible by a Fundamental
Research Grant Scheme (6071366) provided by the Ministry of Education
of Malaysia and a Research University grant (8014059) provided by
Universiti Sains Malaysia.

References
1. Rezania, S., Oryani, B., Park, J., Hashemi, B., Cho, J., Review on transesteri-
fication of non-edible sources for biodiesel production with a focus on eco-
nomic aspects, fuel properties and by-product applications. Energ. Conver.
Manag., 201, article ID 112155, 2019.
2. Santos, F.FP., Rodrigues, S., Fernandes, F.A.N., Optimization of the produc-
tion of biodiesel from soybean oil by ultrasound assisted methanolysis. Fuel
Proc. Technol., 90, 312-316, 2009.
3. Tan, K.T., Lee, K.T., Mohamed, A.R., Production of FAME by palm oil trans-
esterification via supercritical methanol technology. Biomass Bioenerg, 33,
1096-1099, 2009.
4. Trakarnpruk, W., Porntangjitlikit, S., Palm oil biodiesel synthesized with
potassium loaded calcined hydrotalcite and effect of biodiesel blend on elas-
tomer properties. Renew. Energ., 33, 1558-1563, 2008.
5. Arzamendi, G., Arguiñarena, E., Campo, I., Zabala, S., Gandía, L.M., Alkaline
and alkaline-earth metals compounds as catalysts for the methanolysis of
sunflower oil. Catal. Today, 133-135, 305-313, 2008.
6. Kawashima, A., Matsubara, K., Honda, K., Acceleration of catalytic activity
of calcium oxide for biodiesel production. Bioresour. Technol., 100, 696-700,
2009.
Ultrasound-Assisted Biodiesel Production 81

7. Monteiro, M.R., Ambrozin, A.R.P., Lião, L.M., Ferreira, A.G., Critical review
on analytical methods for biodiesel characterization. Talanta, 77, 593-605,
2008.
8. Abdullah, A. Z., Razali, N., and Lee, K. T., Optimization of mesoporous
K/SBA-15 catalyzed transesterification of palm oil using response surface
methodology. Fuel Proc. Technol. 90, 958-964, 2009.
9. Kalva, A., Sivasankar, T., Moholkar, V.S., Physical mechanism of ultra-
sound-assisted synthesis of biodiesel. Ind. Eng. Chem. Res., 48, 534-544,
2009.
10. Hanh, H.D., Dong, N.T., Okitsu, K., Maeda, Y., Nishimura, R., Effects of
molar ratio, catalyst concentration and temperature on transesterification of
triolein with ethanol under ultrasonic irradiation. J. Japan Petr. Inst., 50, 195-
199, 2007.
11. Abdullah, A. Z., Razali, N., Mootabadi, H., and Salamatinia, B., Critical tech-
nical areas for future improvement in biodiesel technologies. Environ. Res.
Lett., 2, 1-5, 2007.
12. Tan, S.X., Lim, S., Ong, H.C., Pang, Y.L., State of the art review on develop-
ment of ultrasound-assisted catalytic transesterification process for biodiesel
production. Fuel, 235, 886-907, 2019.
13. Liu, Y., Xin, H.I., Yan, Y.J., Physicochemical properties of stillingia oil:
Feasibility for biodiesel production by enzyme transesterification. Ind. Crops
Prod., 30, 431-436, 2009.
14. Jeong, G.T., Yang, H.S., Park, D.H., Optimization of transesterification of
animal fat ester using response surface methodology. Bioresour. Technol.,
100, 25-30, 2008.
15. Sakai, T., Kawashima, A., Koshikawa, T., Economic assessment of batch bio-
diesel production processes using homogeneous and heterogeneous alkali
catalysts. Bioresour. Technol., 100, 3268-3276, 2009.
16. Abukhadra, M.R., Salam, M.A., Ibrahim, S.M., Insight into the catalytic con-
version of palm oil into biodiesel using Na+/K+ trapped muscovite/phillipsite
composite as a novel catalyst: Effect of ultrasonic irradiation and mechanism.
Renew. Sust. Energ. Rev., 115, article ID 109346, 2019.
17. Wong, K.Y., Ng, J.H., Chong, C.T., Lam, S.S., Chong, W.T., Biodiesel process
intensification through catalytic enhancement and emerging reactor designs:
A critical review. Renew. Sust. Energ. Rev., 116, article ID 109399, 2019.
18. Ma, F., Hanna, M.A., Biodiesel production: A review. Bioresour. Technol., 70,
1-15, 1999.
19. Meher, L.C., Sagar, V., D., Naik, S.N., Technical aspects of biodiesel produc-
tion by transesterification-A review. Renew. Sust. Energ. Rev., 10, 248-268,
2006.
20. Pukale, D.D., Maddikeri, G.L., Gogate, P.R., Pandit, A.B., Pratap, A.P.,
Ultrasound assisted transesterification of waste cooking oil using heteroge-
neous solid catalyst. Ultrason. Sonochem., 22, 278-286, 2015.
82 Biodiesel Technology and Applications

21. Kojima, Y., Takai, S., Transesterification of vegetable oil with methanol using
solid base catalyst of calcium oxide under ultrasonication. Chem. Eng. Proc.,
136, 101-106, 2019.
22. Dossin, T.F., Reyniers, M.F., Marin, G.B., Kinetics of heterogeneously MgO-
catalyzed transesterification. Appl. Catal. B Environ., 62, 35-45, 2006.
23. Xie, W., Peng, H., Chen, L., Transesterification of soybean oil catalyzed by
potassium loaded on alumina as a solid-base catalyst. Appl. Catal. A Gen.,
300, 67-74, 2006.
24. Albuquerque, M.C.G., Cavalcante, C.L., Torres, A.E.B., Azevedo, D.C.S.,
Parente, E.J.S., Transesterification of castor oil using ethanol: Effect of water
removal by adsorption onto zeolite 3A. Energ. Fuel, 23, 1136-1138, 2009.
25. Cintas, P., Mantegna, S., Gaudino, E. C., Cravotto, G., A new pilot flow reac-
tor for high-intensity ultrasound irradiation. Application to the synthesis of
biodiesel. Ultrason. Sonochem., 17, 985-989, 2010.
26. Kelkar, M.A., Gogate, P.R., Pandit, A.B., Intensification of esterification of
acids for synthesis of biodiesel using acoustic and hydrodynamic cavitation.
Ultrason. Sonochem., 15, 188-194, 2008.
27. Deshmane, V.G., Gogate, P.R., Pandit, A.B., Ultrasound-assisted synthesis of
biodiesel from palm fatty acid distillate. Ind. Eng. Chem. Res., 48, 7923-7927,
2009.
28. Fan, X., Wang, X., Chen, F., Ultrasonically assisted production of biodiesel
from crude cottonseed oil. Int. J. Green Energ., 7, 117-127, 2010.
29. Hingu, S.M., Gogate, P.R., Rathod, V.K., Synthesis of biodiesel from waste
cooking oil using sonochemical reactors. Ultrason. Sonochem., 17, 827-832,
2010.
30. Novelline, R., Squire’s Fundamentals of Radiology, 5th/ed. Harvard University
Press, 1997.
31. Ghodbane, H., Hamdaoui, O., Intensification of sonochemical decolorization
of anthraquinonic dye Acid Blue 25 using carbon tetrachloride. Ultrason.
Sonochem., 16, 455-461, 2009.
32. Wongwuttanasatian, T., Jookjantra, K., Effect of dual-frequency pulsed ultra-
sonic excitation and catalyst size for biodiesel production. Renew. Energ.,
152, 1220-1226, 2020.
33. Badday, A.S., Abdullah, A.Z., Lee, K.T., Transesterification of crude Jatropha
oil by activated carbon-supported heteropolyacid catalyst in an ultra-
sound-assisted reactor system. Renew. Energ., 62, 10-17, 2014.
34. Thanh, L.T., Okitsu, K., Sadanaga, Y., Takenaka, N., Maeda, Y., Bandow, H.,
Ultrasound-assisted production of biodiesel fuel from vegetable oils in a
small scale circulation process. Bioresour. Technol., 101, 639-645, 2010.
35. Parkar, P.A., Choudhary, H.A., Moholkar, V.S., Mechanistic and kinetic
investigations in ultrasound assisted acid catalyzed biodiesel synthesis.
Chem. Eng. J., 187, 248-260, 2012.
36. Murillo, G., Ali, S.S., Sun, J., Yan, Y., Fantozzi, F., Ultrasonic emulsification
assisted immobilized Burkholderia cepacia lipase catalyzed transesterification
Ultrasound-Assisted Biodiesel Production 83

of soybean oil for biodiesel production in a novel reactor design. Renew.


Energ., 135, 1025-1034, 2019.
37. Badday, A.S., Abdullah, A.Z., Lee, K.T., Ultrasound-assisted transesteri-
fication of crude Jatropha oil using cesium doped heteropolyacid catalyst:
Interactions between process variables. Energ., 60, 283-291, 2013.
38. Quah, R.V., Tan, Y.H., Mubarak, N.M., Khalid, M., Nolasco-Hipolito, C. An
overview of biodiesel production using recyclable biomass and non-biomass
derived magnetic catalysts. J. Environ. Chem. Eng., 7(4), article ID 103219,
2019.
39. Zhao, S., Niu, S., Yu, H., Ning, Y., Han, K., Experimental investigation on bio-
diesel production through transesterification promoted by the La-dolomite
catalyst. Fuel, 257, article ID 116092, 2019.
40. Malani, R.S., Shinde, V., Ayachit, S., Goyal, A., Moholkar, V.S., Ultrasound-
assisted biodiesel production using heterogeneous base catalyst and mixed
non-edible oils. Ultrason. Sonochem., 52, 232-243, 2019.
41. Guo, W., Li, H., Ji, G., Zhang, G., Ultrasound-assisted production of bio-
diesel from soybean oil using Brønsted acidic ionic liquid as catalyst. Biores.
Technol., 125, 332-334, 2012.
42. Santos, F.F.P., Malveira, J.Q., Cruz, M.G.A., Fernandes, F.A.N., Production of
biodiesel by ultrasound assisted esterification of Oreochromis niloticus oil.
Fuel, 89, 275-279, 2010.
43. Lifka, J., Ondruschka, B., Influence of mass transfer on the production of
biodiesel. Chem. Eng. Technol., 27, 1156-1159, 2004.
44. Takase, M., Feng, W., Wang, W., Gu, X., Zhu, Y., Li, T., Yang, L., Wu, X.,
Silybum marianum oil as a new potential non-edible feedstock for biodiesel:
A comparison of its production using conventional and ultrasonic assisted
method. Fuel Proc. Technol., 123, 19-26, 2014.
45. Armenta, R.E., Vinatoru, M., Burja, A.M., Kralovec, J.A., Barrow, C.J.,
Transesterification of fish oil to produce fatty acid ethyl esters using ultra-
sonic energy. JAOCS, 84, 1045-1052, 2007.
46. Veljković, V.B., Avramović, J.M., Stamenković, O.S., Biodiesel production by
ultrasound-assisted transesterification: State of the art and the perspectives.
Renew. Sust. Energ. Rev., 16, 1193-1209, 2012.
47. Baêsso, R.M., Costa-Felix, R.P.B., Miloro, P., Zeqiri, B., Ultrasonic parameter
measurement as a means of assessing the quality of biodiesel production.
Fuel, 241, 155-163, 2019.
48. Badday, A.S., Abdullah, A.Z., Lee, K.T., Ultrasound-assisted transesterifica-
tion of crude Jatropha oil using alumina-supported heteropolyacid catalyst.
Appl. Energ., 105, 380-388, 2013.
49. Maneechakr, P., Samerjit, J., Uppakarnrod, S., Karnjanakom, S., Experimental
design and kinetic study of ultrasonic assisted transesterification of waste
cooking oil over sulfonated carbon catalyst derived from cyclodextrin. J. Ind.
Eng. Chem., 32, 128-136, 2015.
84 Biodiesel Technology and Applications

50. Guldhe, A., Singh, B., Rawat, I., Bux, F., Synthesis of biodiesel from
Scenedesmus sp. by microwave and ultrasound assisted in situ transesterifi-
cation using tungstated zirconia as a solid acid catalyst. Chem. Eng. Res. Des.,
92, 1503-1511, 2014.
51. Nikseresht, A., Daniyali, A., Ali-Mohammadi, M., Afzalinia, A., Mirzaie, A.,
Ultrasound-assisted biodiesel production by a novel composite of Fe(III)-
based MOF and phosphotangestic acid as efficient and reusable catalyst.
Ultrason. Sonochem., 37, 203-207, 2017.
52. Mootabadi, H., Salamatinia, B., Bhatia, S. Abdullah, A.Z., Ultrasonic-assisted
biodiesel production process from palm oil using alkaline earth metal oxides
as the heterogeneous catalysts. Fuel, 89, 1818-1825, 2010.
53. Gupta, A.R., Yadav, S.V., Rathod, V.K., Enhancement in biodiesel production
using waste cooking oil and calcium diglyceroxide as a heterogeneous cata-
lyst in presence of ultrasound. Fuel, 158, 800-806, 2015.
54. Chen, G., Shan, R., Shi, J., Yan, B., Ultrasonic-assisted production of bio-
diesel from transesterification of palm oil over ostrich eggshell-derived CaO
catalysts. Bioresour. Technol., 171, 428-432, 2014.
55. Takase, M., Chen, Y., Liu, H., Zhao, T., Yang, L., Wu, X., Biodiesel production
from non-edible Silybum marianum oil using heterogeneous solid base cat-
alyst under ultrasonication. Ultrason. Sonochem., 21, 1752-1762, 2014.
56. Sarve, A., Sonawane, S.S., Varma, M.N., Ultrasound assisted biodiesel pro-
duction from sesame (Sesamum indicum L.) oil using barium hydroxide
as a heterogeneous catalyst: Comparative assessment of prediction abilities
between response surface methodology (RSM) and artificial neural network
(ANN). Ultrason. Sonochem., 26, 218-228, 2015.
57. Xiang, Y., Wang, L., Jiao, Y., Ultrasound strengthened biodiesel production
from waste cooking oil using modified coal fly ash as catalyst. J. Environ.
Chem. Eng., 4, 818-824, 2016.
58. Zhang, F., Fang, Z., Wang, Y.T., Biodiesel production directly from oils with
high acid value by magnetic Na2SiO3@Fe3O4/C catalyst and ultrasound. Fuel,
150, 370-377, 2015.
59. Singh, A.K., Fernando, S.D., Transesterification of soybean oil using hetero-
geneous catalysts. Energ. Fuel, 22, 2067-2069, 2008.
60. Choudhury, H.A., Chakma, S., Moholkar, V.S., Mechanistic insight into
sonochemical biodiesel synthesis using heterogeneous base catalyst.
Ultrason. Sonochem., 21, 169-181, 2014.
3
Application of Catalysts in
Biodiesel Production
Anilkumar R. Gupta and Virendra K. Rathod*

Department of Chemical Engineering, Institute of Chemical Technology,


Matunga (E), Mumbai, India

Abstract
Biodiesel has emerged as a potential substitute for petroleum diesel due to its supe-
riority in terms of renewable, biodegradable, and non-toxic nature. It can be pro-
duced from a wide range of feedstock using acid and base catalyst or biocatalyst.
Catalysts play a vital role in the economical production of biodiesel, as base cata-
lysts are only useful for high-quality feedstock, and low-quality feedstock requires
acid pretreatment followed by base catalysis. While using biocatalyst, simultane-
ous esterification and transesterification reaction can be carried out; however, its
high cost is of primary concern. Therefore, this chapter describes different cat-
alysts such as homogeneous acid (e.g., H2SO4 and H3PO4) and base (e.g., KOH
and NaOH) catalysts, heterogeneous acid (e.g., sulfated zirconia and alumina, and
cation-exchange resins), and base (metal oxide, hydrotalcite, etc.) catalyst, and
immobilized biocatalyst (immobilized lipase onto magnetic nanoparticles, resins,
etc.) with respect to their synthesis, mechanism, catalytic activity, effect on bio-
diesel yield or conversion, and their reusability.

Keywords: Biodiesel, heterogeneous catalyst, homogeneous catalyst, biocatalysis,


catalyst synthesis, transesterification, esterification

3.1 Introduction
Globally, energy demand is continuously increasing due to economic
growth and the rapid rise in the population [1]. The liquid fossil fuel is

*Corresponding author: vk.rathod@ictmumbai.edu.in

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (85–136) © 2021 Scrivener Publishing LLC

85
86 Biodiesel Technology and Applications

one of the primary sources used in transportation, agricultural, and indus-


trial sectors to meet their energy demand [2]. However, the burning of
fossil fuel emits harmful gases like carbon dioxide, nitrogen oxide, volatile
organic compounds, and hydrocarbons, which are responsible for global
climate change [3]. Apart from the environmental problem, it also impacts
the economic growth of major developing countries because of the depen-
dence on the import of crude oil. Being non-renewable is another disad-
vantage as it will be exhausted soon. Considering all these points, biofuel,
such as biodiesel, may be found as one of the alternatives to liquid fossil
fuel.
Biodiesel is monoalkyl esters of long-chain fatty acids produce using
edible or non-edible oils feedstock and alcohol (methanol or ethanol) or
esters (methyl acetate, ethyl acetate, and dimethyl carbonate) in the pres-
ence of a catalyst [4–6]. It can be used in the compression ignition (diesel)
engine (CIG) without any significant modification. Biodiesel appears as
an attractive biofuel because of renewable and biodegradable nature, free
from sulfur, aromatic compounds, and other harmful chemicals that harm
the environment [7]. It shows a high flash point and cetane number, better
lubricating properties, and similarities in physicochemical properties with
conventional petroleum diesel [6] and can be used as B100 (pure biodiesel)
or as a blend with petroleum diesel (B5–B20) in CIG [8].
Generally, edible vegetable oils such as sunflower oil, palm oil, soybean
oil, peanut oil, linseed oil, and coconut oil [9] have been used for the bio-
diesel production, but, due to continuous rise in the price and competing
with the food, raise the concern for the economic feasibility of biodiesel.
Non-edible oils (Jatropha curcas, Karanja, Neem, Madhuca indica, Rubber
seed oil, etc.), waste cooking oil (WCO), palm fatty acid distillate (PFAD)
[10], and animal fats (tallow, chicken fat, and a by-product of fish oil) [9]
can be the possible options to reduce the price and eliminate the competi-
tion with food. Furthermore, it also helps in solving waste disposal (animal
fats and waste oil) problems and reduction in the usage of arable land.
In recent times, microalgae become third-generation feedstock due to
the effective conversion of sunlight, water, and carbon dioxide into algal
biomass. The yield of algae oil per hector is 15 times higher than the second-
best oil, i.e., palm oil [11]. Therefore, there are different options available
to select the best available feedstock to reduce the cost and eliminate the
competition with food. The different feedstock used by the country has
been listed in Table 3.1 [12–14].
Biodiesel production can be carried out either by using a catalytic or
non-catalytic process. In the non-catalytic process, transesterification of
vegetable oil occurs well above the critical temperature of alcohol (513-K
Catalyst for Biodiesel 87

Table 3.1 Different biodiesel feedstock used by the respective country


in the world.
Country Biodiesel feedstock or raw material
Argentina Soybean, Sunflower, Crambe abyssinica, Jatropha
oil, Macrocarpa
Australia Waste cooking oil, Animal fat
Bangladesh Pongamia Pinnata, Rubber seed oil
Brazil Soybean oil, Palm oil, Caster oil
Canada Canola, Sunflower, Soybean
China Rapeseed oil, Waste cooking oil
Cuba Jatropha oil, Neem oil, Moringa oil
Europe Rapeseed oil, Sunflower oil
Finland Rapeseed oil, Animal fat
France Rapeseed oil, Sunflower oil
Germany Rapeseed oil
Ghana Palm oil, Coconut oil
Greece Sunflower oil, Cotton seed oil, Rapeseed oil
India Jatropha oil
Indonesia Palm oil
Ireland Animal tallow, Waste cooking oil
Italy Rapeseed oil
Japan Waste cooking oil
Kenya Castor oil
Malaysia Palm oil
Mali Jatropha oil
Mexico Animal fat, Waste cooking oil
Peru Palm oil, Jatropha oil
Philippine Coconut oil, Jatropha oil
(Continued)
88 Biodiesel Technology and Applications

Table 3.1 Different biodiesel feedstock used by the respective country


in the world. (Continued)
Country Biodiesel feedstock or raw material
Russia Rapeseed oil, Soybean oil, Sunflower oil
South Africa Canola oil, Sunflower oil, Soybean oil
South Korea Waste cooking oil
Spain Sunflower oil
Sweden Rapeseed oil
Tanzania Jatropha oil
Thailand Palm oil, Jatropha oil,
Turkey Sunflower oil, Rapeseed oil
USA Soybean oil
Zimbabwe Jatropha oil

methanol and 514-K ethanol) and pressure of 10–45 MPa [15]. This
process is not economically viable as energy consumption is too high
due to the requirement of high temperature and pressure. Therefore,
catalytic biodiesel synthesis is the most preferred process over the non-
catalytic process [16]. In the catalytic process, chemical catalysts or biocat-
alysts are used to transform the triglyceride (TG) molecule into fatty acid
alkyl esters (FAAEs).
The chemical catalysts exist in homogeneous (alkali and acid) and het-
erogeneous (solid base and acid) form. Traditionally, homogeneous cata-
lysts like sodium hydroxide (NaOH) and potassium hydroxide (KOH) are
used as it shows high catalytic activity for the transesterification of veg-
etable oil. However, it generates an ample amount of wastewater during
downstream processing, and being corrosive also enhances the cost due
to the requirement of costly corrosive resistance equipment [17]. The
use of heterogeneous catalysts is an alternative to overcome the problem
associated with a homogeneous catalyst with an advantage of recyclabil-
ity. The use of biocatalysts—free or immobilized form—is another option
for the transesterification process. It is beneficial over the chemical cat-
alyst in terms of producing high purity biodiesel, along with being envi-
ronment-friendly [18]. However, its higher cost is a significant obstacle
to employ on a commercial scale. The different catalytic methods used for
Catalyst for Biodiesel 89

Homogeneous

Chemical
catalysis

Heterogeneous

Catalytic methods

Free enzyme

Biocatalysis

Immobilized enzyme

Figure 3.1 The different catalytic methods for the biodiesel production.

biodiesel production have been shown in Figure 3.1. This chapter concen-
trates on trends and current development in the field of catalysis for bio-
diesel production.

3.2 Homogeneous Catalysis for the Biodiesel


Production
In general, biodiesel is produced by using a homogeneous catalyst owing to
the number of benefits like faster reaction rate, high selectivity, ease of its
application, and optimization. The homogeneous catalysts are acids (sul-
furic acid, hydrochloric acid, phosphoric acid, aluminium chloride, zinc
chloride, etc.) and bases (sodium hydroxide, sodium methoxide, potas-
sium hydroxide, potassium methoxide, etc.). For scaling-up, the homo-
geneous catalyst is preferred as it requires milder reaction conditions.
However, the major drawback of the utilization of homogenous catalysts is
that it produces a vast amount of wastewater during the purification of the
crude biodiesel. However, studies have been carried out to solve this issue
by introducing newer technologies. Some of the selective homogeneous
catalysts used for biodiesel production with their reaction conditions have
been tabulated in Table 3.2.

3.2.1 Homogeneous Acid Catalyst


The low quality and non-edible feedstock content large amounts of free fatty
acids (FFAs) (5%–40%) [32] which are not suitable for the base-catalyzed
90
Table 3.2 List of some selective homogeneous catalyst used for biodiesel production with their reaction conditions.
Reaction conditions Yield/Conv.
Catalyst Feedstock a b c d (%) Ref.
KOH Madhuca indica oil 1.5 333 9:1 90 91.76 [19]
NaOH Waste oil 0.49 333 12.2:1 63 95.6 [20]
KOH Hempseed oil 1.2 316.4 6.4:1 98.5 [21]
NaOH Soybean oil 1.0 333 6:1 60 90.0 [22]
CH3ONa WCO assisted 0.75 MW power 6:1 3 97.9 [23]
750 W
KOH Soybean oil with US assisted 1 333 6:1 6 98.0 [24]
KOH WCO 1.2 338 6:1 60 93.2 [25]
KOH Karanja oil 1.25 333 9:1 120 88.7 [26]
Biodiesel Technology and Applications

KOH Fish oil 0.75 343 9:1 60 97.11 [27]


KOH WCO with micro reactor assisted 1.16 335.4 9.4:1 2 98.26 [28]
H2SO4 Mixed oil pretreatment 2.5 333 6:1 60 96.6 [29]
AlCl3 Canola oil with THF as co-solvent 5 383 24:1 1080 98.0 [30]
p-toluene sulfonic acid Corn oil with dimethyl ether as co-solvent 4 353 6:1 120 100.0 [31]
a, Catalyst loading (%); b, Temperature (K); c, Molar ratio; d, Reaction time (min).
Catalyst for Biodiesel 91

transesterification reaction as it leads to soap formation as shown in Scheme


3.1. This formed soap not only reduces the biodiesel yield but also prevents
the separation of FAAEs and glycerin and thus affects the quality of the
biodiesel [33]. Therefore, acid catalysts sulfuric acid (H2SO4), hydrochloric
acid (HCl), phosphoric acid (H3PO4), and sulfonic acid have been used to
carry out esterification reaction (Scheme 3.2). Among all these homoge-
neous acid catalysts, sulfuric acid was reported to give a better result for
the conversion of FFAs to biodiesel [34].
In most of the cases, acid catalyst only used as a pre-treatment step to
reduce the FFAs content ~1% as acid-catalyzed transesterification reaction
is 4,000 times slower than the base and requires a large amount of alcohol
[9, 35, 36]. The reason for the slower reaction rate for the acid-catalyzed
transesterification reaction as compared to base-catalyzed transesterifi-
cation reaction has been revealed by Lotero et al. [37] through studying
both, i.e., acid and base-catalyzed mechanism. In the case of acid-catalyzed
transesterification mechanism (Scheme 3.3), the first protonation of the
carbonyl oxygen of TG molecule is taken place by Brønsted acid which is
the crucial step for the interaction between TG and catalyst. This initial
step increases the electrophilicity of the carbonyl carbon, which becomes
more susceptible to the nucleophilic attack of the alcohol. In the final step,

O O
R C OH + MOH R C OM + H2O

FFAs Base catalyst Soap Water


M = Na+, K+, etc.

Scheme 3.1 Soap formation in the presence of free fatty acids with the base catalyst.

H
O O OH OH

OH + R C OH R -H R
R C H C OH C OH
O O
Acid R' H R'
FFAs R'—OH
catalyst
Alcohol
H
H
O O OH H OH
OR' -H
R C R C OR' R C O R C O
-H2O
H O H
FAAEs O
R' R'

Scheme 3.2 General esterification reaction mechanism in the presence of an acid catalyst.
92 Biodiesel Technology and Applications

+
O OH
O O
O R1 O R1
H+
(i) R2 O R2 O
O R3 O R3
O O

+
OH OH + R
O O
O
O R1 O
(ii)
ROH R H 1
R2 O R2 O
O R3 O R3

O O

O
OH + R
O O O R1 OR
O H OH
(iii) R1 R2 O +
R2 O
O R3 O R3 +
O H
O

R1, R2, R3: Long chain fatty acid; R: Atkyl group of alcohol

Scheme 3.3 Acid-catalyzed transesterification of TG (i) Protonation of carbonyl oxygen


by acid catalyst; (ii) Alcohol (ROH) attacks on electrophilic carbon; (iii) Proton migration
and formation of product (FAAEs). This mechanism repeated two more times to give
finally three moles of product.

intermediate break down with the release of the product (FAAEs) and
regeneration of catalyst, i.e., H+ ion.
Whereas, the base-catalyzed transesterification reaction takes a more
direct pathway in which active species alkoxide ion (OR− act as a strong
nucleophile) generated initially and attacked to the carbonyl group to
transform TG to FAAEs. Thus, the difference in the formation of elec-
trophilic species in the case of an acid catalyst and nucleophilic species
for base catalyst is ultimately responsible for the acid and base-catalyzed
transesterification reaction rate.
Because of aforesaid reason, the feedstock with high FFAs is subjected
two-step process, i.e., esterification followed by transesterification. Bouaid
et al. [38] carried out the conversion of crude Jatropha oil into biodiesel,
where they employed the two-step process. In the first step, Jatropha oil
was subjected to esterification using acid catalyst H2SO4 to reduce the %
FFAs content from 9.48% to less than 1%, under the optimum condition
Catalyst for Biodiesel 93

of molar ratio of 20:1, reaction temperature of 333 K, and 5 wt% H2SO4 in


60-min reaction time. In the second step, alkaline catalyzed transesterifi-
cation of Jatropha oil (FFAs < 1%) was performed and obtained more than
98% biodiesel yield.
In another study, crude algal oil (5.3 %FFAs) was converted into bio-
diesel using a similar approach, i.e., esterification followed by transester-
ification. In the beginning, acid (H2SO4) esterification was performed to
reduce the %FFAs from 5.3 to <1% with reaction parameter of 12:1 metha-
nol to oil molar ratio, reaction temperature of 333 K, 1 wt% of acid catalyst
(H2SO4), and agitation speed of 400 rpm in the reaction time of 90 min.
The esterified oil was then used for alkali (KOH) transesterification reac-
tion, where the maximum biodiesel yield of 83% was achieved in 20-min
reaction time [39].
Apart from Brønsted acid, homogeneous Lewis acid catalyst Aluminium
chloride (AlCl3) was also employed for the biodiesel production from
canola oil. The maximum conversion of 98% was achieved with optimized
conditions—methanol-to-canola oil molar ratio of 24:1, temperature
383 K, and reaction time of 1,080 min using THF as a co-solvent [30].
The acid catalysts are very sensitive to the water present in the feed-
stock as it deactivates the catalyst. It was investigated that a small portion
of water (0.1 wt%) in the raw material affects the yield of acid-catalyzed
biodiesel. When the water content becomes 5 wt%, it was observed that
acid-catalyzed transesterification of soybean oil was completely hindered
and esters conversion dropped drastically from 95.1% to 5.6% under the
similar reaction conditions—molar ratio of 6:1, reaction temperature of
333 K, and 3 wt% H2SO4 with reaction time of 5,760 min [40]. The acid-
catalyzed biodiesel synthesis has some disadvantages such as it generates
a large amount of acidic water during the neutralization process, corrode
the equipment, and slower reaction rate for the conversion of feedstock to
biodiesel. Therefore, the heterogeneous acid catalyst has been developed to
solve the problem associated with a homogeneous acid catalyst, which will
be discussed in Section 3.3.1.

3.2.2 Homogeneous-Base Catalyst


The alkali-based homogeneous catalysts are the most preferred catalysts
for commercial biodiesel production [34]. It includes alkali metal hydrox-
ides and alkali metal alkoxides. The metal alkoxides are the more reactive
catalyst in carrying out transesterification reaction as it readily dissociates
into CH3O− and metal ion (Na+ and K+). However, due to lower cost, alkali
hydroxides (NaOH and KOH) are preferred over alkoxides for industrial
94 Biodiesel Technology and Applications

biodiesel production [41]. NaOH has added an advantage over KOH due
to faster solubility in methanol and economically available [42]. The per-
formance of the alkali catalyst depends on the quality of the feedstock, e.g.,
feedstock with %FFAs <1 wt% and water content <0.5 wt% give higher
biodiesel yield. The homogeneous base-catalyzed transesterification reac-
tions require a relatively smaller amount of catalyst, lower temperature and
molar ratio, and shorter time to reach the completion of the reaction [43].
The general transesterification of TG with their mechanism has shown in
Scheme 3.4, and the process flow block diagram for base-catalyzed bio-
diesel production has been depicted in Figure 3.2.
The canola oil was transesterified with 3 wt% NaOH and gave 85% bio-
diesel yield with oil-to-methanol molar ratio of 6.5:1, reaction temperature

Mechanism:

(i) ROH + B RO +
+ BH


O O
O OR
O
O R1 O R1
(ii) R2 O + RO– R2 O
O R3 O R3
O O

O–
OR O O
O –
O R1 O
(iii) R2 O R1 OR + R2 O
O R3 O R3
O O

O O
O OH
+
(iv) R2 O + BH R2 O + B
O R3 O R3
O O
R1, R2, R3: Long chain fatty acid; R: Atkyl group of alcohol

Overall reaction:

B
Triglyceride + 3 ROH 3FAAEs + Glycerol

Scheme 3.4 Alkali catalyzed transesterification of TG (i) formation alkoxide ion (strong
nucleophile); (ii) attack of a nucleophile to the carbonyl carbon of TG to form tetrahedral
intermediate; (iii) break down of intermediate to form FAAE; (iv) regeneration of the
catalyst. This mechanism repeated two more times to give finally three moles of product.
Catalyst for Biodiesel 95

Oil or pretreated oil + Catalyst + MeOH

MeOH
Transesterification

Separation Evaporation

Lower Phase Upper Phase


Crude Biodiesel

Purification Washing followed by


drying

Alkaline
Glycerol wastewater
Pure Biodiesel
Saponified
product

Figure 3.2 The block diagram for base-catalyzed biodiesel production.

of 70°C, and agitation speed of 550 rpm, in the reaction time of 120 min
[44]. In one more study, different concentrations of NaOH were used and
reported their respective rapeseed biodiesel yield. The maximum yield of
96% was found under mechanical stirring with 2 wt% NaOH in 15-min
reaction time [45]. Verma et al. [26] employed KOH for Karanja biodiesel
synthesis using both methanol and ethanol as an acyl acceptor. The highest
yield of 88.7% and 77% were obtained for methanolysis and ethanolysis,
respectively, at the reaction temperature of 333 K, the molar ratio of 9:1, and
catalyst loading of 1.25 wt% in 120-min reaction time. Maddikeri et al. [6]
investigated the potassium methoxide (KOCH3) catalyzed ultrasound-as-
sisted interesterification of WCO using methyl acetate as an acyl acceptor.
Under the optimal reaction conditions such as 12:1 WCO to methyl acetate
molar ratio, 313 K, and catalyst concentration of 1 wt%, biodiesel yield of
90% was obtained.
The main drawback of homogenous base-catalyzed transesterification
process is the requirement of high-quality pretreated feedstock as a pres-
ence of a small amount of water turn TG molecule into the soap as shown
in Scheme 3.5. The other disadvantages are energy-intensive purification
step and generation of a large amount of wastewater during the neutral-
ization of alkaline biodiesel. Hence, the researchers have been worked to
96 Biodiesel Technology and Applications

O
O O
O R1 OH
R2 O + 3MOH + 3H2O 3R OM + OH –
O R3 OH

O Base Water Soap Glycerol


TG Catalyst
+ +
M = Na, K, etc

Scheme 3.5 Soap formation in the presence of water and base catalyst.

develop a heterogeneous base catalyst to resolve the issues linked with


homogeneous catalyst and described in detail in Section 3.3.2.

3.3 Heterogeneous Catalyst


The process of conversion of feedstock to biodiesel can be improved by
introducing heterogeneous catalysts, as it minimizes the downstream pro-
cessing costs and generation of hazardous waste, which is linked with the
homogeneous catalysts [42]. Heterogeneous catalyst has added advantages
as ease of recovery, reusability, and promoting cost-effective greener pro-
cesses [46]. Heterogeneous catalysts can tolerate high FFAs and water con-
tent in the feedstock. These catalysts can be operated in harsh conditions
like high temperature and pressure.
The heterogeneous catalysts used for biodiesel production are metal
oxides, mixed metal oxides, supported acid and base catalysts, zeolites,
hydrotalcite (HT), and ion exchange resin, derived from waste sources,
renewable biomass, immobilized biocatalysts, etc. The main problem asso-
ciated with the use of heterogeneous catalysts is an increase in the mass
transfer resistance due to a further rise in the heterogeneity by forming
three phases (feedstock + alcohol + catalyst) [4]. An increase in the hetero-
geneity results in reduced interaction between TGs and alcohol molecules.
These mass transfer limitations can be reduced by introducing nanocatalyst
or solid support/carrier to provide a larger surface area for the interaction.
The supports are designed to carry out entrapment or anchoring of active
molecules on its surface or inside pores [47]. From the decade, a continu-
ous effort has been made to improve the properties of support materials.
The efficient catalytic process can be made by improving catalytic activity
and selectivity, which depends on the tailoring of the catalytic materials
and distribution of active sites [48]. The mesoporous support materials
offer such type of characteristics to improve the catalytic activity [47].
Catalyst for Biodiesel 97

The heterogeneous catalysts for the FAAEs or biodiesel synthesis grouped


into acid and base catalysts.

3.3.1 Heterogeneous Acid Catalyst


Considering the obstacles allied with the usage of homogeneous acid cat-
alysts, there are many researchers who have been worked to develop het-
erogeneous acid catalysts as an alternative for biodiesel synthesis. These
catalysts are useful in the esterification of readily available low-quality feed-
stock. Also, it ables to carry out simultaneous esterification and transester-
ification of FFAs and TG molecules, respectively. The solid acid catalyst
shows high catalytic activity due to the presence of a variety of acidic sites
of different strength on its surfaces [49]. To date, different heterogeneous
acid catalysts, sulfated zirconia (SZ) and alumina, sulfated silica, tung-
stated zirconia [50], heteropolyacid (HPA)–derived catalyst [51], biomass
impregnated catalyst [52], H-form zeolites, cation-exchange resins [53],
etc., have been evaluated for the conversion of FFAs and TGs to biodiesel.
The SZ is a solid superacid catalyst and found a wide variety of appli-
cations in organic transformations like isomerization, alkylation, and
esterification reactions. The various methods have been evaluated for
the preparation of SZ such as impregnation of dried zirconium hydrox-
ide [Zr(OH)4] with H2SO4 or ammonium sulfate [(NH4)2SO4], co-
precipitation of Zr(OH)4 and sulfates, impregnation of crystalline zirco-
nium oxide (ZrO2) with H2SO4 or (NH4)2SO4 [54], using Zirconium oxy-
chloride octahydrate (ZrOCl2.8H2O) as a precursor to form Zr(OH)4, and
impregnating with chlorosulfonic acid (HSO3Cl) followed by calcination
as shown in Figure 3.3 [10].
It has been observed that SZ with low or medium sulfur content only
shows a tetragonal phase, whereas high sulfur content SZ shows minor
monoclinic phase along with primary tetragonal phase [55]. One of the
studies found that HClSO3 SZ contains three times more sulfur than H2SO4
SZ [56]. Zhang et al. [56] employed SZ as a superacid catalyst for one-
step synthesis of biodiesel from crude rice bran oil with 40% FFA and
reported yield above 92% for three consecutive cycles, under the reaction
conditions—molar ratio of 12:1, 6 wt% catalyst, 393 K, and 720 min of
reaction time. A similar super acid catalyst (SZ) was used to esterify the
PFAD into biodiesel. In this study, methanol to PFAD molar ratio of 8:1,
catalyst loading of 3.5 wt%, reaction temperature of 333 K, stirring speed
of 250 rpm, and molecular sieve of 5 wt% were used and obtained 91.5%
conversion in 150-min reaction time [10].
98 Biodiesel Technology and Applications

Synthesis of Zr(OH)4 precursor Synthesis of HCISO3-ZrO2

ZrOCl2.8H2O + Zr(OH)4 +
25% NH3 solution 0.5 M HCISO3
Hydrolysis at Stirred for 5 min
373 K pH 9-10
Mixture was heated to 393 K
Obtained solid was aged to remove solvent completely
for 720 min at 373 K

Washed with DI water Dried in oven at 393 K for 1440 min


till neutral pH

Absence of CI¯ detected


by AgNO3 test Calcined at 873 K for 180 min

Obtained Zr(OH)4 was dried


for 1440 min at 378 K HCISO3-ZrO2 catalyst obtained

Figure 3.3 The flow diagram of synthesis of SZ from zirconium oxychloride octahydrate
using chlorosulfonic acid.

The γ-alumina (Al2O3) intensively used for acid catalyst preparation due
to the high thermal stability and presence of Lewis acid sites on its surface
[57]. There have been various γ- Al2O3–derived heterogeneous acid cata-
lysts reported for the biodiesel synthesis. Vahid et al. [58] prepared alu-
mina supported SZ nanocatalyst by grinding ZrOCl2·8H2O and (NH4)2SO4
with different amount of aluminium sulfate [Al2(SO4)3·18H2O] followed
by calcination at 773 K for 600 min. The obtained catalyst was employed
for the conversion of WCO to biodiesel and reported the 93.5% yield in
93-min reaction time under the optimum parameters of 12.7:1 methanol
to WCO molar ratio, the temperature of 421.5 K, and catalyst loading (alu-
mina-supported SZ) of 2.9 wt%. Kashyap et al. [59] used γ-Al2O3 as a cata-
lyst for the interesterification of Karanja oil to produce biodiesel.
In another study, zirconia-alumina, the composite material, was synthe-
sized using the sol-gel method and then sulfated with wet impregnation
and hydrothermal method. In the wet impregnation method, Zr-Al2O3
impregnated with 1 N sulfuric acids followed by calcination at 823 K for
720 min, whereas, in the case of hydrothermal method, Zr-Al2O3 and
thiourea were fed to the steel reactor and autoclave to 453 K for 1,440 min.
The hydrothermally synthesized sulfated Zr-Al2O3 catalyst was used for
the Jatropha oil biodiesel production, and a maximum conversion of 96%
was obtained with reaction parameters of 20:1 methanol to oil molar ratio,
Catalyst for Biodiesel 99

1.5 wt% catalyst, and reaction temperature of 383 K in 180-min reaction


time [60].
The mesoporous silica, such as SBA-15, can be used as catalysts sup-
port because of high surface area (400–900 m2/g) and high mechanical
and hydrothermal stability [61]. SBA-15 is synthesized with a coopera-
tive self-assembly process using Pluronic 123 as a template and tetraethyl
orthosilicate as a silica source under acidic conditions [62]. After comple-
tion of synthesis, the template is removed to form hexagonal pores with a
diameter of 5 to 30 nm [63]. The SZ supported on SBA-15 was synthesized
by dispersing SBA-15 into zirconium (IV) propoxide and hexane solution
followed by stirring until all the solvent get evaporated. The obtained solid
was dried and immersed in 0.5 N H2SO4 for 30 min. The sulfated product
was dried and subjected to calcination for 180 min. The obtained SZ sup-
ported on SBA-15 was used for the transesterification of palm oil. Under
the optimal conditions (catalyst loading of 5 wt%, molar ratio of 20:1,
473 K, and reaction time of 300 min) biodiesel yield reached more than
95% [64].
The sulfated titania-silica catalyst was synthesized by two steps pro-
cess. In the first step, titania-silica support was incorporated by the sol-gel
method, and in the second step, sulfuric acid was impregnated on support.
This catalyst was tested for the conversion of waste oil into biodiesel. The
maximum conversion was found to be 94.7% using a molar ratio of 20:1,
reaction temperature of 353 K, catalyst loading of 10 wt% in the reaction
time of 180 min [65]. For biodiesel production, esterification of palmitic
acid was carried out using two sulfonated catalyst—ZrO2-TiO2-SO3H and
ZrO2 -TiO2@SO24−. These catalysts were synthesized in two steps. In the
initial step, ZrO2-TiO2 nanorods were prepared by wet chemical precip-
itation method, whereas in the next step, nanorods were functionalized
2−
with an active group as depicted in Figure 3.4. The ZrO2 -TiO2@SO4 cata-
lyst exhibited high acidity and surface area and gave 95.3% biodiesel yield
while ZrO2-TiO2-SO3H showed lower acidity and surface area with 93.1%
biodiesel yield. However, ZrO2-TiO2-SO3H exhibited relatively high stabil-
ity even after five cycles [66].
Tungstated zirconia (WO3/ZrO2) was prepared by a simple impregna-
tion method where Zr(OH)4 was impregnated with ammonium metatung-
state hydrate [(NH4)6(H2W12O40).nH2O] followed by calcination at 1,073 K
for 240 min. The synthesized solid acid catalyst (WO3/ZrO2) was subjected
to in situ transesterifications of Scenedesmus sp. dried biomass under soni-
cation with molar ratio of 60:1, reaction temperature of 333 K, and catalyst
amount of 4 wt%. The maximum conversion of 71.37% was obtained in
20-min reaction time [67]. Peruzzolo et al. [68] synthesized heterogeneous
100 Biodiesel Technology and Applications

(a)
ZrO2
400 0C

pH=11 3h

TiO2 nanorods ZrOCl2·8H2O ZrO2–TiO2 nanorods


(b)

500 0C
H2SO4
5h SO42-

ZrO2-TiO2 nanorods ZrO2-TiO2@SO42- nanorods

C2H5
(c)
O

C2H5
OH SH 90 0C
O Si (CH3)3
OH
24h N2
O

OH
C2H5
ZrO2-TiO2 nanorods
OEt
Si (CH3)3
O SO3H
H2SO2 OEt

HCI OH
OH

ZrO2-TiO2-SO3H nanorods

Figure 3.4 Synthesis of (a) ZrO2-TiO2 nanorods, (b) ZrO 2 − TiO 2 @SO 24− nanorods, and
(c) ZrO2-TiO2-SO3H. (Adapted from Ref. [66] with permission from Elsevier Inc.).

acid catalysts in which tungsten species were supported on commercial


mesoporous silica. The two tungsten species—ammonium para tung-
state (NH4)10H2(W2O7)6 and sodium tungstate (Na2WO4)—were used for
incipient-wetness impregnation on mesoporous silica Cariact Q-15. For
the impregnation of (NH4)10H2(W2O7)6, the aqueous solution was used
whereas for Na2WO4, ammonia aqueous solution. Both these catalysts
were then employed for the esterification of palmitic acid. The ammonium
Catalyst for Biodiesel 101

paratungstate impregnated catalyst gave 83.9% conversion. In comparison,


sodium tungstate impregnated catalyst gave 82.3% conversion under sim-
ilar reaction conditions, i.e., the temperature of 393 K, the molar ratio of
12:1, catalyst loading of 10 wt% in 360-min reaction time.
The heteropolyacid (HPA) belongs to the class of polyoxometalates and
generally soluble in the polar medium like water and methanol [51]. Hence,
it is required to impregnate on a solid support or modify to become hetero-
geneous in carrying out the transformation of feedstock to biodiesel. The
HPA was modified with surfactant (C16TA—cetyltrimethylammonium) to
get a heterogeneous catalyst with enhancing water tolerance capacity. The
obtained catalyst was then subjected to the esterification of palmitic acid
with methanol to give 94.7% conversion [69]. The HPA–phosphotungstic
acid (H3PW12O40) impregnated on mesoporous KIT-6 for biodiesel syn-
thesis from neem oil. For impregnation, dried support material KIT-6
was mixed with deionized (DI) water and stirred. The aqueous solution of
H3PW12O40 was added dropwise to support material under stirring in 360
min. After addition, the mixture was filtered and washed with DI water.
Then, the impregnated KIT-6 was dried and calcined at 573 K for 300 min.
The obtained catalyst was used for the transesterification of neem oil with
methanol, and under optimal reaction conditions, maximum conversion
of 90% was achieved [70]. de Oliveira’s group [71] also performed impreg-
nation of phosphomolybdic (H3PMo12O40) HPA on bentonite to carry out
esterification of the distillate of deodorization of palm oil. The maximum
conversion of 93.3% was found for the molar ratio of 30:1, catalyst loading
of 5 wt%, reaction temperature of 403 K, and reaction time of 120 min.
Waste biomass is an excellent choice for the synthesis of a low-cost het-
erogeneous acid catalyst. It can be synthesized using different techniques
like direct sulfonation or carbonization, followed by sulfonation [72]. The
direct sulfonation is a straightforward and most widely used process under
which sulfonation and carbonization of waste biomass are performed
using H2SO4 at a preset temperature and time. Savaliya and Dholakiya [73]
prepared the solid acid catalyst using sugarcane bagasse where in situ car-
bonization and sulfonation were carried with concentrated sulfuric acid.
The yield of bagasse derived catalyst was 25–35 wt% w.r.t. raw material.
Furthermore, the obtained catalyst was studied for the biodiesel pro-
duction from the soapstock oil, and the maximum conversion of 97.2%
was reported. Its reusability was also tested and recycled up to three con-
secutive cycles without hampering its activity. In another study, the direct
sulfonation of coconut meal residue was carried out for biodiesel synthe-
sis from waste palm oil. For the preparation of the catalyst, coconut meal
residue was mixed with concentrated sulfuric acid to bring out in-situ
102 Biodiesel Technology and Applications

incomplete carbonization. The prepared catalyst was used to transesterify


the waste palm oil into biodiesel. The maximum biodiesel yield of 92.7%
was achieved, and the catalyst recycled up to four batches with maintaining
high catalytic activity [74].
Guo et al. [75] investigated that directly sulfonated materials showed
reduced catalytic activity and low acid density and no recyclability. To
overcome this problem, phenol was used as an additive due to its ease of
sulfonation and enhancement of acid density. In the case of carbonization
followed by sulfonation process (Figure 3.5), different biomass such as
agricultural waste, lignin, bagasse, algae, and seed cake (DOWC, de-oiled
seed waste cake) was first carbonized to remove moisture and breakdown
the carbon-carbon bond followed by sulfonation [72]. The carbonization
can be carried out using different processes—pyrolysis, hydrothermal, sol-
vothermal, etc. The various biomasses used to derive solid acid catalysts
through different process conditions for biodiesel production have been
summarized in Table 3.3.
Zeolites are inorganic material—aluminosilicates occur naturally or can
be synthesized by varying Si/Al ratios [51]. Doyle et al. [88] prepared a
zeolite Y catalyst from kaolin and studied its catalytic activity in the ester-
ification of oleic acid. For the preparation of zeolite Y first NaY zeolite was
synthesized using kaolin clay then transformed into HY zeolite by cat-
ion exchange. The synthesized catalyst applied for the conversion of oleic
acid into biodiesel. The obtained conversion was 85% with molar ratio of
6:1, catalyst concentration of 5 wt%, and reaction temperature of 343 K
in 60-min reaction time. Patel and Narkhede [89] were synthesized the
zeolite derived catalyst in which HPA–12-tungstophosphoric acid (TPA)
was anchored to Zeolite Hβ. In the synthesis procedure, sodium zeolite β
with Si/Al ratio 10 was converted to NH+4 form via the conventional ion

COOH COOH
OH OH
O OH OH
O
HO3S SO3H
HOOC COOH H2SO4 (98%)
DOWC HOOC
COOH
Protein, HO
HO O O
carbohydrate HO3S
and lipid (trace) OH SO3H
HO OH
HO
HOOC SH
HOOC
Carbon source Activated Carbon Sulfonated Activated Carbon

Figure 3.5 Synthesis of sulfonated acid catalyst from the biomass (DOWC) using
carbonization followed by sulfonation route. (Adapted from Ref. [76] with permission
from Elsevier Inc.).
Table 3.3 Different biomass derived solid acid catalyst used for the biodiesel production.
Acid site density
Biomass a b c d Feedstock (mmol/g) Yield/Conv. (%) Ref.
Peanut shell 996 900 746 600 Cotton seed oil 6.85 90.20 [77]
Sugarcane bagasse 648 30 423 900 WCO 3.69 93.80 [78]
Coconut shell 695 240 373 900 Palm oil 3.21 88.15 [79]
Deoiled Jatropha 623 240 363 300 Jatropha curcas oil 2.24 99.13 [80]
curcas seed
Cassava stillage 873 180 433 120 WCO 1.60 94.30 [81]
residue 453 120 1.65 94.10
Deoiled Mesua ferrea 996 60 453 480 Oleic acid 2.01 42.00 [76]
L. seed
Hardwood 800 -- 423 1440 WCO 0.043 89.00 [82]
Vegetable oil asphalt 773–973 -- 483 600 Cotton seed oil 2.04 89.93 [83]
Deoiled Calophyllum 673 300 423 600 Calophyllum 2.8 36.40 [84]
inophyllum seed inophyllum oil
Catalyst for Biodiesel

(Continued)
103
104

Table 3.3 Different biomass derived solid acid catalyst used for the biodiesel production. (Continued)
Acid site density
Biomass a b c d Feedstock (mmol/g) Yield/Conv. (%) Ref.
De-alkaline lignin 533 1200 423 600 Jatropha and Soybean 5.05 92.20 [85]
673 120 423 600 blended oils 5.35 93.20
Sugarcane bagasse 723 300 423 300 PFAD -- 80.00 [86]
Wood waste 723 120 423 900 Canola oil 2.6 7.60 [87]
White wood bark 948 1.2 24.40
Biodiesel Technology and Applications

Wood shaving 1148 0.43 8.40


a, Carbonization temperature (K); b, Carbonization time (min); c, Sulfonation temperature (K); d, Sulfonation time (min).
Catalyst for Biodiesel 105

exchange process. Then, calcination was carried out at 823 K for 360 min
to get Hβ zeolite. After the preparation of zeolite Hβ, the impregnation
of TPA was performed to obtain the acid supported catalyst. When the
synthesized catalyst was subjected to the esterification of oleic acid, the
highest conversion of 85% was obtained and retained its catalytic activity
even after four cycles.
The cation exchange resins are organic solid acid catalysts that show
quite high catalytic activity under mild reaction conditions. It also utilized
for industrial biodiesel production from the feedstock with high FFAs
content [53, 90, 91]. The most commonly used resin for the esterifica-
tion and transesterification reactions are sulfonic ion-exchange resin and
commercially available as Amberlyst resins (Amberlyst-15, Amberlyst-45,
Amberlyst-46, etc.), Nafion resins (Nafion NR50 and SAC-13), and EBD
resins (EBD-100, EBD-200, and EBD-300) [90]. These ion-exchange resins
are usually cross-linked polymer on which strong sulfonic acid groups are
bonded, and it acts as an active site for the esterification reaction [92].
Amberlyst-15 is one of the macroporous reticular polystyrene-based
sulfonic ion-exchange resins, generally used to esterify the feedstock
for the biodiesel production. For example, Amberlyst-15 (surface area
53 m2/g) used in the esterification of crude Macaúba oil with ethanol
in the temperature range 333–373 K. The maximum yield of 32.8% was
found with 72% conversion [93]. Hykkerud and Marchetti [94] applied
Amberlyst-15 Wet catalyst for the esterification of oleic acid with etha-
nol, and the highest yield of 53% was achieved in the reaction time of
300 min. Ilgen [95] investigated oleic acid esterification with methanol
using Amberlyst-46 (surface area 75 m2/g) catalyst. With optimal reaction
conditions, i.e., molar ratio of 3:1, catalyst loading of 15 wt%, reaction
temperature of 373 K, and the reaction time of 120 min, the extent of con-
version of 98.23% was attained. The Amberlyst-45 retained its catalytic
activity even after 10 cycles.
Nafion NR50 is a super acidic ion exchange resin made up of sulfonic
acid supported fluorinated polymers. In contrast, Nafion SAC-13 is a com-
posite material (sulfonic acid functionalized Nafion polymer on porous
silica) with high surface area (more than 200 m2/g) and high-temperature
stability (443 K). Both the acidic ion-exchange resins showed pronounced
catalytic performance in carrying out esterification and transesterification
reactions. For instance, López and co-worker [96] examined the trans-
esterification of triacetin using Nafion acid ion-exchange resins. They
investigated the swelling properties of Nafion NR50 and found methanol
enhances the higher degree of polymer swelling that results in increased
exposure of active sites of the catalyst. The Nafion NR50 showed a higher
106 Biodiesel Technology and Applications

initial reaction rate as compared to Nafion SAC-13 due to the presence of


more acid site density.
EBD resins are polymeric matrix made of styrene and divinylbenzene
co-polymers and functionalized with strong sulfonic acid. EBD-100 is a
microporous gelular resin and has a low quantity of divinylbenzene (5%–
8%), while the other two, i.e., EBD-200 and EBD-300 are macroporous
resins have a significant portion of divinylbenzene (12%–20%). Russbueldt
and Hoelderich [97] were used aforesaid acid ion-exchange resins for the
pretreatment of FFAs containing vegetable oils. The catalytic performance
of the microporous gelular resins EBD-100 was superior to the macropo-
rous resins—EBD-200 and EBD-300. The obtained result might be due to
dense gel matrix and poor methanol uptake of the macroporous materials,
while high swelling of low cross-linked EBD-100 in methanol. The main
problem associated with the use of EBD resins for the pretreatment is its
fast catalytic deactivation due to the presence of the low amount of salts as
a contaminant in the feedstock. However, deactivation of EBD resins can
be reversed by simply washing with hydrochloric acid [90].
Another, acid resins catalyst Purolite (polystyrene cross-linked with
divinylbenzene) D5081 and D5082 were used to esterify the oleic acid
with methanol at 338 K. It has been observed that both Purolite D5081
and D5082 showed highest oleic acid conversion for the similar catalyst
loading as compared to other tested sulfonic ion-exchange resins such as
Amberlyst-15, Amberlyst-35, and Nafion SAC-13. The reason might be
explained on the basis of ease of accessibility of the reactant molecule to
the active sites of the catalyst [98].

3.3.2 Heterogeneous-Base Catalyst


As mentioned, the drawbacks of homogeneous base catalysts in Section
3.2.2, the various heterogeneous base catalysts, have been investigated for
the transesterification of TGs. For example, alkaline and alkaline earth
metal carbonates (e.g., Na2CO3 and CaCO3), alkaline earth metal oxide
(e.g., CaO and MgO), mixed metal oxide (e.g., Ca-Zn mixed oxides), HT
(e.g., Al-Ca HT), supported and doped solid base catalyst (e.g., CaO or
MgO supported on silica), etc. [99–101].
The alkali and alkaline earth metal carbonate—sodium carbonate
(Na2CO3), potassium carbonate (K2CO3), calcium carbonate (CaCO3),
magnesium carbonate (MgCO3), etc.—have been attempted for the trans-
esterification of vegetable oil. In K2CO3 catalyzed transesterification of
Jatropha curcas oil, potassium carbonate reacted with methanol and gen-
erated potassium methoxide, which was measurably responsible for the
Catalyst for Biodiesel 107

high biodiesel yield (ca. 98%) [102]. However, CaCO3 and MgCO3 were
not directly used to transesterify the feedstock as it has been less basic and
needs to transform into a more active catalyst, i.e., corresponding oxides
[99].
Alkaline earth metal oxides—MgO (magnesium oxide), CaO (calcium
oxide), SrO (strontium oxide), and BaO (barium oxide)—were extensively
studied as a solid base catalyst. These metal oxides show intense catalytic
activity in carrying out transesterification reaction due to the presence of
M2δ+–O2δ− ion pairs (Scheme 3.6) [103] and exhibit its effectiveness in the
order: MgO < CaO < SrO < BaO. Furthermore, Lee et al. [104] reported
that the basic strength of the metal oxides depends on the electronegativity
of the conjugated metal ion. The ionic radius of the Group II elements,
i.e., Mg2+, Ca2+, Sr2+, and Ba2+ in the periodic table, increases from top to
bottom due to which electronegativity decreases and that leads to decrease
the force of attraction on the electrons. Owing to this effect, the basic char-
acteristics of attached O2− enhances in alkaline earth metal oxides. Though
BaO is more effectively transesterified the feedstock, they are unsuitable

R–O– H+
2 + 2-
(i) ROH + M O
M O
O O–
O OR
O
O R1 O R1
– H
(ii) R2 O + R–O R2 O
O R3 O R3
O M O O

O
O OR O O
O R1 O–
(iii) R2 O R1 OR + R2 O
O R3 O R3
O O

O O
O– H+ OH
(iv) R2 O + R2 O +M O
O R3 M O O R3

O O
M-Ca2+, Mg2+, Sr2+, Ba2+
R1, R2, R3: Long chain fatty acid; R: Alkyl group of alcohol

Scheme 3.6 Alkaline earth metal oxides catalyzed transesterification of TG.


108 Biodiesel Technology and Applications

for biodiesel synthesis owing to higher solubility in methanol and noxious


property [105].
Among all alkaline earth metal oxides, CaO are the most investigated
catalyst for the conversion of vegetable oil into biodiesel due to its avail-
ability, lower cost, and ease of handling. It can be derived from different
sources in the form of carbonates such as chalk, limestone, marble, and
protective shells of a living organism such as eggs and seashells [106, 107].
Generally, the calcium carbonate is calcined in the range of 1,073–1,273 K
to decompose into CaO as shown in Eq. 3.1 [108].

−1273 K
CaCO3 ←1073
 → CaO + CO2 ­ (3.1)

To test the activity of seashell derived CaO, Buasri et al. [109] carried
out the decomposition of waste seashell at 1,273 K and obtained CaO sub-
jected to transesterification of palm oil. At the optimal reaction conditions,
the catalyst gave a maximum yield of 97.23%. A similar study was per-
formed by the Goli and Sahu [110] using waste chicken eggshell derived
CaO (7 wt%), and the maximum biodiesel yield of 93% was reported.
Different calcium oxide sources used for biodiesel production with their
yield have been shown in Table 3.4.
Many of the researchers have been attempted to improve the catalytic
activity of CaO through physical or chemical treatment. Calcination,
hydration, and dehydration (C-H-D) of CaO sources are one the physical
treatment method to increase the surface area and basic strength of the
catalysts [119]. For instance, Begum’s group [120] carried out C-H-D treat-
ment to eggshells and reported enhanced surface area (8.64 m2/g) and basic
strength (12.2<H_<15.0) as compared to commercial CaO (surface area,
3.0 m2/g; basic strength, 9.8<H-<12.2). The treated eggshells were applied
for the transesterification of waste frying oil, and the obtained result was
compared with commercial CaO. It has been observed that C-H-D treated
CaO gave 94.52% yield, while commercial CaO gave only a 67.57% yield
for the same reaction conditions. Similar treatment (C-H-D) was applied
to Polymedosa erosa seashells to obtain highly activated CaO nanocata-
lyst with diameter 66 nm and surface area 90.61 m2/g [121]. The obtained
nano-CaO was subjected to Jatropha biodiesel synthesis and the yield of
98.54% was achieved with reusability up to six cycles without significant
loss in catalytic activity.
Some studies also revealed that the catalytic activity of CaO could be
improved by treating with different chemicals. For example, Tang and
co-workers [122] used surface modifier octadecyltrichlorosilane to modify
Table 3.4 Recent examples of different sources of calcium oxide used for the biodiesel production.
Calcination Transesterification reaction
CaO sources temp. (K) Feedstock conditions Yield/Conv. (%) Ref.
Chicken Eggshell 1,173 Palm kernel oil Catalyst amount - 4 wt%, T - 328 97.10 [111]
K, M/O - 10:1, t - 60 min
Waste Eggshell 1,173 Eucalyptus oil Catalyst amount - 12 wt%, T - 70.50 [112]
328 K, M/O - 6:1, t - 90 min
Chicken Eggshell 1,123 WCO Catalyst amount - 3 wt%, T - 338 93.10 [113]
K, M/O - 12:1, t - 90 min
Waste venus 1,173 Palm oil Catalyst amount - 5 wt%, T - 338 97.00 [114]
clam K, M/O - 15:1, t - 360 min
Limestone 1,273 Pongamia oil Catalyst amount - 12 wt%, T - 97.28 [115]
338 K, M/O - 15:1, t - 180 min
Chicken manure 1,123 WCO Catalyst amount - 7.5 wt%, T - 90.00 [116]
338 K, M/O - 15:1, t - 240 min
Ostrich Eggshell 1,273 WCO Catalyst amount - 1.5 wt%, T - 96.00 [117]
Chicken Eggshell 338 K, M/O - 12:1, t - 120 min 94.00

Mussel shell 1,173 Camelina sativa oil Catalyst amount - 1 wt%, T - 338 95.00 [118]
Clam shell K, M/O - 12:1, t - 120 min 93.00
91.00
Oyster shell
Catalyst for Biodiesel
109
110 Biodiesel Technology and Applications

the CaO surface and applied for biodiesel production. It has been observed
that the surface-modified CaO performed very well (93.5% yield) as com-
pared to commercial CaO (80.6% yield) in the transformation of soybean
oil into biodiesel even in the presence of a small quantity of water. Lukic
et al. [123] carried out mechanochemical treatment of CaO using glycerol
to transform calcium oxide (9.3<H_<10) into more basic catalyst calcium
diglyceroxide (11.0<H_<15.0).
The mixed metal oxides are referred to as those which have more
than one metal oxide. The primary reason behind the synthesis of mixed
metal oxides is to improve the basic strength, surface area, and mechan-
ical strength as compared to single metal oxides [124]. The single metal
oxide like CaO leached out during transesterification reaction due to weak
mechanical strength and form suspension in the reaction mixture. The
formed suspension creates a problem during downstream processing. Lee
et al. [125] prepared the mixed metal oxide (MgO-ZnO) by co-precipitation
method and compared its effectiveness with corresponding individual
metal oxide (MgO and ZnO) for transesterification of Jatropha oil with
methanol. They found that MgO-ZnO mixed metal oxide is more active
catalyst for the conversion of TG into biodiesel and gave 83% yield while
individual metal oxide gave 64% for MgO and 41% for ZnO.
Taufiq-Yap and co-workers [126] also synthesized the calcium-based
mixed metal oxides (CaMgO and CaZnO) using the co-precipitation
method. For the synthesis of the catalyst, they used corresponding metal
nitrate, i.e., Ca(NO3)2, Mg(NO3)2, and Zn(NO3)2 in basic solution. The cal-
cination of the precipitates was performed at 1,073 K for CaMgO and at
1,173 K for CaZnO. The obtained catalysts showed higher basic strength
and reusability (up to six cycles) for the biodiesel production from crude
Jatropha curcas oil as compared to single metal oxide (CaO). The enhanced
catalytic activity of mixed metal oxides may be due to the synergy effect of
the interaction of multi-metal ion which results in increase basic strength
[127].
Hydrotalcite (HT) is a layer double hydroxide that occurs naturally or
it can be synthesized in the lab, which belongs to the class of basic and
anionic clays having general formula M12−+x M 3x+ (OH2)x+ (Ax/n)n-.yH2O. In
this formula, M+2 is a divalent metal cation and M+3 is a trivalent metal
cation, whereas An- is n- valent anions (CO3−, SO24−, Cl−, NO3−), and the value
of x in between 0.25 and 0.33 [128, 129].
HT was synthesized for the biodiesel production using simple co-
precipitation of aqueous solution magnesium and aluminum nitrates in
basic medium (NaOH and Na2CO3) followed by calcination. The prepared
Mg-Al HT was used for the transesterification of soybean oil. The maximum
Catalyst for Biodiesel 111

conversion of 90% was achieved with 5 wt% catalyst loading. However, the
reusability of the catalyst was very poor as it showed a sudden drop in
the conversion from 90.7% to 64% after the first cycle [130]. To enhance
the catalytic performance, Gandia and co-workers [131] first prepared the
Mg-Al HT and then rehydrated to get more basic and recyclable HT, as
shown in Figure 3.6. The obtained catalyst was applied for the transesterifi-
cation of sunflower oil and gave 96% conversion with 2 wt% catalyst.
Fatimah et al. [132] also tried to enhance the catalytic activity of HT
by increasing basic sites on the surface. It was reported that potassium
impregnation increases the basicity of the catalyst. Therefore, potassium
fluoride (KF) modified HT catalyst was synthesized for biodiesel produc-
tion. The synthesized catalyst with 30% KF loaded HT gave maximum bio-
diesel yield of 97%.
Catalyst support provides not only large surface area to which cat-
alyst particles can be anchored but also chemical stability. The different
catalyst support like alumina, silica, zinc oxide, magnetic nanoparticles
(MNPs), and activated carbon (AC) have been used to carry out biodiesel
production.
Aluminum oxide also known as alumina with the chemical formula
Al2O3 is structural materials exist in different transitional phases that
have significant industrial application. The most common usage is as cat-
alyst support owing to high thermal stability, inexpensive, and high spe-
cific surface area [133]. The alumina-supported potassium oxides (K2O/
Al2O3) have been synthesized for the biodiesel production. The different
potassium compounds such as KF, KI, KOH, KNO3, and K2CO3 have been
used to support on Al2O3 by incipient-wetness impregnation method.

Triglycerides + MeOH

CO32- Mg(OH)2
H2O [AI(OH)2]+
Calcination
Mixed Oxide Reconstructed Hydrotalcite
Mg-Al Hydrotalcite Mg2+(AI3+)Ox Mg2+ nAI3+m (OH)(2n+3m)•yH2O
Active OH¯
[Mg2+nAI3+m (OH)2(n+m)]m+[Ax-]m/x•yH2O Rehydration No-active OH¯

Biodiesel + Glycerol

Figure 3.6 Calcination and rehydration of Mg-Al hydrotalcite for the biodiesel
production. (Adapted from Ref. [131] with permission from Elsevier Inc.).
112 Biodiesel Technology and Applications

For example, Baz’s group [134] synthesized KF impregnated γ-Al2O3 to


transesterify the canola oil and reported maximum biodiesel yield of 97.7%
with 3 wt% catalyst loading. Similarly, Luengnaruemitchai and co-work-
ers [135] prepared KOH/Al2O3 catalyst by impregnation of Al2O3 by an
aqueous solution of KOH. The obtained catalyst was tested for the trans-
esterification of palm oil, where the maximum yield of 91.07% achieved
with 3 wt% catalyst loading. Among different potassium sources (KF, KI,
KOH, KNO3, and K2CO3) the best result obtained by KF loaded alumina,
i.e., KF/Al2O3 catalyst, giving 99.6% yield with 3 wt% catalyst loading. The
enhanced catalytic activity might be due to the formation of the K2O phase
and surface Al-O-K groups which results in increased basicity [136]. CaO/
Al2O3 catalyst also prepared by the impregnation technique using calcium
acetate as a precursor. The obtained catalyst was employed for palm oil
biodiesel production and achieved a maximum yield of 98.64% with 5.97
wt% catalyst [137].
Mesoporous silica such as SBA-15 and MCM-41 are most widely used
for the incorporation of the guest molecule. For example, Mota’s group
[138] used MCM-41 mesoporous material (surface area ~ 1,000 m2/g) as
a support for the synthesis of amine-functionalized basic catalysts for the
biodiesel production. For the functionalization, the activated MCM-41
[MCM-41 activated at 453 K to remove adsorbed water and expose silanol
groups (Si-O-H)] first functionalized with 3-chloro propyl triethoxy silane
to form chloro-functionalized MCM-41 (MCM-41-Cl). In the next step
nucleophilic substitution of MCM-41-Cl with guanidine-derived com-
pound (1,5,7-triazabicyclo [4,4,0] dec-5-ene) was carried out as shown in
Figure 3.7 and obtained catalyst labeled as MCM-41-guanidine. The pre-
pared catalyst was subjected to transesterification of soybean oil and it gave

N N
CI CI N
N
N
Si Si
~ 3.5 nm O OO O Si Si
O O N O OO O
O O
N / NaH / THF / reflux
HN
CI Si(OEt)3 /
PhCH3 / refux

MCM-41 MCM-41-CI MCM-41-guanidine

Figure 3.7 Functionalization of MCM-41 surface with guanidine derivative


(1,5,7-triazabicyclo [4,4,0] dec-5-ene). (Adapted from Ref. [138] with permission
from Elsevier Inc.).
Catalyst for Biodiesel 113

99% biodiesel yield in 180 min. However, the major drawback of the cat-
alyst was its reusability, as it shows a drastic drop in biodiesel yield (26%)
during the third cycle. The loss of catalytic activity of MCM-41-guanidine
might be due to the neutralization of basic sites by FFA present in the oil.
Samart et al. [139] prepared the KI supported mesoporous silica by
incipient wetness impregnation method applied for the biodiesel produc-
tion from soybean oil. The highest yield of 90.09% was achieved with 15%
KI loading on mesoporous silica and 5 wt% catalyst amount in reaction
time of 480 min. Furthermore, the catalyst was tested for their reusabil-
ity and it was found that after the first cycle, the yield was just 60%. The
mixed metal oxide–loaded mesoporous silica catalyst was prepared by Xie
and Zhao [140] for the soybean biodiesel production. They used incipient
wetness impregnation method under which SBA-15 was impregnated with
an aqueous solution of Ca(NO3)2 and (NH4)6Mo7O24, precursors followed
by drying at 373 K and calcination at 823 K. The obtained CaO–MoO3–
SBA-15 catalyst employed to transesterify the soybean oil. The optimized
conversion of 83.2% was achieved with 6 wt% catalyst.
ZnO supported calcium oxide catalyst was made in two steps. In first
step, ZnO was prepared from zinc oxalate, and in the second step, it was
impregnated with an aqueous solution of calcium acetate followed by
calcination at 873 K. The prepared catalyst further activated by thermal
treatment at 1,073 K for 60 min and used for the transesterification of
ethyl butyrate with methanol where it gave more than 90% yield [141].
Moholkar and co-workers [142], synthesized ZnO supported KI catalyst,
by a wet impregnation method. For the preparation of catalyst, 35 % aque-
ous solution of KI was used to impregnate ZnO to give KI/ZnO catalyst.
The obtained catalyst was applied for the ultrasound-assisted transesterifi-
cation of soybean oil to produce biodiesel. The highest conversion of 94.5%
was reported, while after the fifth cycle conversion was reduced to 54.7%.
The supported catalyst is lacking in the reusability generally due to leach-
ing of active catalytic species from the support, therefore, Kawi’s group
[143] calcium-doped Ce-incorporated SBA-15 catalyst was synthesized
and compared with SBA-15 supported mixed metal oxide (CaO-CeO2). For
the Ca-doped catalyst synthesis, the first Ce-incorporated SBA-15 was pre-
pared by dissolving P123, Ce(NO3)3, and tetraethylorthosilicate in acidic
solution (2 N HCl) followed by calcination at 823 K. After Ce-SBA-15
synthesis, it was doped with calcium by using Ca(NO3)2 aqueous solution
and finally calcined at 923 K to get Ca-doped-Ce-SBA-15. Whereas, for
mixed metal-supported SBA-15 (CaO-CeO2/ABA-15) was synthesized
using impregnation of an aqueous solution of calcium and cerium nitrate
followed by calcination at 923 K. Both these catalysts were tested for the
114 Biodiesel Technology and Applications

transesterification of palm oil with methanol. Ca-doped-Ce-SBA-15 cata-


lyst gave 96.4% yield while CaO-CeO2/ABA-15 catalyst gave only 71.2%.
Also, calcium doped catalyst reused up to 15 cycles without significant loss
in catalytic activity. The higher catalytic activity and reusability are shown
by calcium doped catalysts due to strong interaction between active species
and support.
MNPs recently received much attention as a support material because
of high surface area, simple synthesis and functionalization, and low cost
[144]. The main advantage of MNPs supported catalysts is their ease of
recovery under the magnetic field as compared to other energy-intensive
and time-consuming separation processes such as centrifugation and filtra-
tion [145]. The MNPs can be synthesized using different magnetic elements
like Fe, Ni, and Co into their oxides. The functionalization of MNPs can be
carried out by modification of their surface so that the active functional
group can be attached. For example, Chiang et al. [146] were synthesized
the MNPs, i.e., Fe3O4 by co-precipitation method using Fe+2 and Fe+3 in 1:2
molar ratio in basic medium then coated with silica using tetraethyl ortho-
silicate to form Fe3O4@Silica. The formed core-shell structure was function-
alized with organic super base 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD)
in two steps. First trimethoxysilyl propoxymethyl oxirane was anchored
onto MNPs coated silica and in the second step, TBD was functionalized to
get final catalyst (TBD-Fe3O4@silica). The synthesized basic catalyst was
tested for the Fatty acid methyl ester formation from algae oil and the high-
est yield of 97.1% was reported. Liu et al. [147] synthesized CaO coated
MgFe2O4 MNPs for soybean biodiesel production and with 1 wt% catalyst
loading, the maximum yield of 98.3% was achieved in 180 min.
AC exhibits high surface area, decent thermal stability, and low-cost
which makes it advantageous to use as catalyst support [148]. The highly
porous AC can be prepared from its precursor in two ways: the first way is
the physical activation where precursor thermally treated at 1,073–1,173
K in the presence of CO2 or steam, and in a second way, chemical activa-
tion is carried out by treating precursor with acid or base in the tempera-
ture range 723–923 K [149]. The support of the functional group on AC
was generally carried by impregnation. For instance, Narowska and co-
workers [150] were prepared AC from the beech tree and impregnated
with a saturated aqueous solution of KOH to get KOH/AC catalyst. This
catalyst was employed for the corn oil biodiesel production and it gave 92%
yield with 0.75 wt% catalyst amount. Fadhil et al. [151] also synthesized the
AC supported KOH catalyst using the wet impregnation method. The AC
was prepared from the precursor—polyethylene terephthalate waste. The
KOH/AC catalyst showed high basic strength and high catalytic activity in
Catalyst for Biodiesel 115

the transformation of WCO and waste fish oil (WFO) into biodiesel. The
WCO gave 92.66% yield while WFO gave 88.12% yield, also the catalyst
was reused up to five cycles. In another study, AC supported CaO was syn-
thesized by impregnation of an aqueous solution of calcium nitrate on AC
followed by calcination at 813 K. The CaO/AC catalyst successfully trans-
esterified the WCO into biodiesel and showed the yield of 77.32% with 5.5
wt% catalyst in 142 min [152].

3.4 Biocatalysts
Nowadays, biocatalysts have found significant applications in various
industrial processes due to their ability to carry out selectively chemical
transformation [52, 153]. The enzyme-catalyzed reactions are environment
friendly, required milder reaction conditions, and high-quality byproduct
(glycerol) [154]. These advantages lead to the utilization of the enzyme
as a biocatalyst for biodiesel production. Also, it aids in overcoming the
problem associated with chemical catalysis like the generation of a large
amount of acidic or basic wastewater. Owing to insensitive to FFAs and
water content, biocatalyst (lipase) can be used for the lower grade feed-
stock to carry out simultaneous esterification and transesterification [155].
However, one of the bottleneck for the utilization of enzyme as a bio-
catalyst is its high cost. Immobilization of enzymes can mitigate the cost
by improving stability and reusability along with catalytic activity [156].
Immobilization of enzymes has been carried out using the physical method
(adsorption and entrapment/encapsulation) and chemical method (cova-
lent attachment and cross-linking) as shown in Figure 3.8 [157].

Enzyme immobilization
methods

Physical Chemical
methods methods

Adsorption Entrapment Covalent Cross-linking


attachment
Gel/fiber Micro- MOFs Cross-linking to
entrapping encapsulation embedding support matrix CLEAs

Enzyme

Figure 3.8 Different methods for immobilization of enzymes. (Adapted from Ref. [157]
with permission from Elsevier Inc.).
116 Biodiesel Technology and Applications

Enzyme immobilization was carried out using the physical adsorption


method, in which solid support soaked in the enzyme solution prepared
in the buffer. During the soaking period, hydrophobic interaction, and salt
linkage happened which lead to adsorption [158]. For instance, Thakur
and co-workers [159] immobilized the extracellular lipase (isolated from
Pangong Lake followed by purification) on activated biochar prepared
from sugarcane bagasse. For the immobilization, activated biochar was
packed in the column, and then, lipase solution was loaded and elute recir-
culated. After one day, the immobilized column washed thrice with a phos-
phate buffer solution of pH 7 followed by drying. The immobilized lipase
used for the transesterification of bacterial oil and gave 92.23% yield. It was
reused and retained their 75% activity after the third cycle.
Another physical method, entrapment, was introduced for the immo-
bilization of lipase (from Pseudomonas cepacia) and utilized for biodiesel
synthesis from hybrid oil with high FFAs. In this technique, polyvinyl alco-
hol, sodium alginate, and lipase were mixed to get a clear solution and
added dropwise to CaB4O7 solution (prepared from H3BO3 and CaCl2)
to get bio-support catalyst. The transesterification of non-edible oil with
2-propanol was carried out in jacketed packed bed bioreactor which holds
a bio-support catalyst in a fixed position. The biodiesel yield of 84.58%
was reported and the catalyst retained their ca.70% activity with respect to
initial yield even after 10 cycles [160].
Encapsulation of enzyme was also reported for the physical immobili-
zation process under which lipase encapsulated in κ-carrageenan by co-
extrusion technique. Typically, lipase buffer solution of pH 7 (inner core)
and the κ-carrageenan aqueous solution of 3% w/v (outer shell) were coex-
truded from the coaxial needle at flow rate 5–50 ml/h and palm oil was
used as the continuous phase. After the encapsulation of lipase, it was hard-
ened in 2% KCl solution overnight, followed by drying. This encapsulated
lipase was tested for palm oil biodiesel production, where it gave 100%
conversion in 1440 min. The reusability of the biocatalyst was also exam-
ined and it was found that catalyst retained their 85% activity even after
the fifth cycle [161]. To further improve the activity of encapsulated lipase,
Nadar and Rathod [162] encapsulated ultrasound treated lipase within the
metal-organic framework as shown in Figure 3.9. This encapsulated lipase
shows the increased relative activity of 155% with high thermal stability.
Covalently immobilization of enzymes to the carrier is one of the
chemical methods, which provides strong interaction between carrier and
enzyme and thus the stability during biocatalytic processes. Lipase from
Rhizopus oryzae was immobilized on Fe3O4 MNPs with covalent linkage
using 3-aminopropyltriethylenesilane (APTES) and glutaraldehyde as a
Catalyst for Biodiesel 117

Sonication
Horn
Zn2+ + N NH

10 min
Room
Temperature

Lipase Sono-chemical Highly Activated Lipase-MOF


(5 mg/mL) Treatment Lipase

Native Lipase Activated Lipase Lipase-MOF

Figure 3.9 Encapsulation of ultrasound treated lipase within metal-organic framework


(MOF). (Adapted from Ref. [162] with permission from Elsevier Inc.).

linker. Furthermore, lipase was immobilized non-covalently on Fe3O4 and


compared its activity for the conversion of Chlorella Vulgaris microalgae oil
into biodiesel. It has been observed that covalently immobilized lipase gave
the highest conversion of 69.8% as compared to non-covalently immobi-
lized lipase - lipase/Fe3O4 (54.14%) and lipase/Fe3O4 -APTES (ca. 60%).
The lower conversion in case physically adsorbed catalyst, i.e., lipase/Fe3O4,
may be due to insignificant change in conformation along with leakage of
lipase from bare Fe3O4. However, lipase/Fe3O4-APTES showed enhanced
conversion, as amine group of APTES causes dipole interaction between
enzyme and carrier with less steric hindrance results in better conversion
than adsorbed catalyst—lipase/Fe3O4 [163].
Lipase from Candida rugosa was covalently immobilized onto micro-
porous biosilica-polymer composite, as shown in Figure 3.10. The immobi-
lized and free lipase tested for the conversion of algal oil into biodiesel. The
obtained results indicate that immobilized lipase gave higher yield (96.4%)
as compared to free lipase (85.7%) and also showed thermal and broader
pH range stability. Additionally, immobilized biocatalyst performed very
well during reusability and only showed a 17% decreased in catalytic activ-
ity after six cycles [164].
Cross-linked enzyme aggregates (CLEA) is an effective and low-cost
approach for the enzyme immobilization. For instance, Badoei-dalfard’s
group [165] prepared active CLEA by the addition of glutaraldehyde to the
Km12 lipase solution (5 mg/ml, pH 7) under stirring. Then, it was cova-
lently attached to amine-functionalized Fe3O4 (prepared from APTES)
118 Biodiesel Technology and Applications

O Br
O
OH OH
OH OH
OH
Br
Br
Br Br
Br H2C O CI
Br Br O
Bromoacetyl 2-Chloroethyl acrylate
bromide
Biosilica surface Br-end functionalization
biosilica

NH2
O O
NH
Cl Cl Cl O
Cl
Cl
Cl
Br Br Br H2N Br Br Br
Br Br NH2 Br O O
Br
Ethylendiamine Glutaraldehyde
Biosilica-g-chloroethyl acrylate

Br Br Br Br Br
Br Br Br
Br Br
Lipase

Figure 3.10 Lipase from Candida rugosa immobilization onto micro-porous biosilica.
(Adapted from Ref. [164] with permission from Elsevier Inc.).

with the aid of glutaraldehyde and termed as mCLEA-lip. The immobilized


CLEA transesterified the WCO and gave 1.45-fold higher yield (71%) as
compared to free lipase (49%). Moreover, it showed high stability terms of
temperature, pH, and storage with nine times reusability. L62 lipase from
Staphylococcus haemolyticus was used for the synthesis of two immobilized
biocatalysts: i) H-L62 [hydrophobic adsorption and entrapment onto poly
(methacrylate-co-divinyl benzene) resin] and ii) HC-L62 [hydrophobic
adsorption and entrapment with covalent cross-link onto poly (methac-
rylate-co-divinyl benzene) resin] (Figure 3.11). H-L62 biocatalyst showed
higher olive oil conversion (90%) than the HC-L62 (80%). However, the
reusability of H-L62 was significantly reduced after the fifth cycle, whereas
HC-L62 reused for more than 10 cycles without significant loss in the
activity. The reduction in the activity of H-L62 might be due to leaching
of active catalyst (lipase L62), while HC-L62 retained its activity owing
to cross-linking among each other [166]. Some selected reports in recent
Catalyst for Biodiesel 119

N C

N C
L62 lipase glutaraldehyde

adsorption/ cross-linking
entrapment

MA-DVB bead H-L62 HC-L62

Figure 3.11 L62 lipase immobilization using adsorption/entrapment and adsorption/


entrapment with cross-linking onto MA-DVB [poly (methacrylate-co-divinyl benzene)
resin]. (Adapted from Ref. [166] with permission from Elsevier Inc.).

years on different immobilization methods for biodiesel production are


tabulated in Table 3.5.
The different immobilization methods and their advantages and dis-
advantages are summarized in Table 3.6. There are several immobi-
lized enzyme available commercially for the biodiesel production such
as Novozym 435—Immobilized Candida Antarctica Lipase B (CALB)
on hydrophobic support acrylic resin, Lipozyme TL IM—Immobilized
Thermomyces lanuginosus lipase on silica gel support, and Lipozyme RM—
Immobilized Rhizomucor miehei lipase, on an anionic resin, etc. Lipozyme
TL IM is a regiospecific biocatalyst that preferentially catalyzed 1 and 3
position of TGs, whereas Novozym 435 is not shown preferential position
for catalyzation [178].

3.5 Conclusion
From the last decade, the price of petroleum diesel has been increased
tremendously. Hence, as an alternative, biodiesel production became a
subject of interest where numerous researches have been conducted for
the economical production of biodiesel. Biodiesel catalysis is one of the
areas which has been explored a lot with respect to each and every type
of catalyst (homogeneous, heterogeneous, and biocatalyst). Therefore,
this chapter reports detailed information regarding previous and recently
developed catalysts.
Through our analysis of this report, it has been observed that for the
higher conversion of feedstock into biodiesel, some of the catalysts required
high temperature or needed a high amount of catalyst or excess of alcohol,
120

Table 3.5 Some selected reports on different immobilization methods for biodiesel production.
Immobilization Yield/Conv.
methods Carrier Lipase Reaction conditions (%) Recycles Ref.
Adsorption silica aerogels Rhizopus oryzae Esterification: butanol/oleic acid 80 12 [167]
lipase - 1:1, T - 310 K, biocatalyst -
450 IU, t - 60 min
Adsorption Granular activated Candida Transesterification: Isobutanol/ 100 NA [168]
carbon (ACG-E) antarctica Palm oil - 6:1, biocatalyst - 50 (ACG–
and activated carbon lipase B mg, T - 313 K, stirring - 300 E), 82
cloth (ACC-E) rpm, t - 2400 min (ACC–
E)
Adsorption Polypropylene Pseudomonas Transesterification: Methanol/ 98 NA [169]
fluorescens Soybean oil - 8:1, biocatalyst
lipase - 600 mg, T - 303 K, water
Biodiesel Technology and Applications

content - 0.5 mg/g of catalyst,


t - 4,200 min
Adsorption hydrotalcite Lipozyme TL Transesterification: Methanol/ 92.8 2 [170]
WCO - 4:1, biocatalyst - 4
wt%, T - 303 K, stirring - 300
rpm, t - 6300 min
(Continued)
Table 3.5 Some selected reports on different immobilization methods for biodiesel production. (Continued)
Immobilization Yield/Conv.
methods Carrier Lipase Reaction conditions (%) Recycles Ref.
Covalent epoxy-functionalized Candida Transesterification: Methanol/ 68 16 [171]
binding silica antarctica canola oil - 3:1, biocatalyst
(CALB) - 13 wt%, T - 323 K, stirring -
250 rpm, t - 5760 min
Thermomyces 45 16
lanuginosus
(TLL)
Rhizomucor Transesterification: Methanol/ 98 16
miehei (RML) canola oil - 6:1, biocatalyst
- 13 wt%, T - 323 K, stirring -
250 rpm, t - 5760 min
Cross-linking MNPs Candida Transesterification: Methanol/ 90 10 [172]
antarctica FFAs - 10:1, biocatalyst - 1
(CALB) wt%, T - 303 K, t - 180 min
Covalent Amberlite IRA-93 Rhizopus Transesterification: Methanol/ 90.5 7 [173]
binding oryzae lipase soybean oil - 4.8:1, biocatalyst
- 24 U/g oil, T - 310 K, water
content - 60 wt% stirring -
Catalyst for Biodiesel

180 rpm, t - 2,880 min


(Continued)
121
122

Table 3.5 Some selected reports on different immobilization methods for biodiesel production. (Continued)
Immobilization Yield/Conv.
methods Carrier Lipase Reaction conditions (%) Recycles Ref.
Entrapment silica aerogels Candida Transesterification: Methanol/ ~90 5 [174]
antartica, sunflower oil - 1:1, biocatalyst
Lipozyme ~ 6 LU mg −gel1, T - 313 K, t -
3,000 min
Entrapment Celite supported Candida Transesterification: Methanol/ 60 NA [175]
sol-gels antarctica, triolein - 1:1, biocatalyst - 1 g,
Novozym T - 313 K, t - 360 min
435
Biodiesel Technology and Applications

Covalent polyporous magnetic Candida Transesterification: Methanol/ 92.3 5 [176]


binding cellulose support antarctica oil - 1.6:1, biocatalyst - 15
lipase B wt%, MW - 400 W, T - 333 K,
t - 120 min
Catalyst for Biodiesel 123

Table 3.6 Advantages and disadvantages of different immobilized methods.


(Adapted from Ref. [177] with permission from Elsevier Inc.).
Methods Advantages Disadvantages
Adsorption Preparing conditions are The interaction between the
mild and easy with low lipase and the carrier is
cost. The carrier can be weak, so the immobilized
regenerated for repeated lipase was sensitive to
use. pH, ionic strength, and
temperature, etc. The
adsorption capacity is small
and the protein might be
stripped off from the carrier.
Covalent The immobilized lipase The preparation conditions are
bond is rather stable because rigorous, so the lipase might
of the strong forces lose its activity during the
between the protein and immobilized process. Some
the carrier. coupling reagents are toxic.
Cross-linking The interaction between The cross-linking conditions
the lipase and the are intense and the
carrier is strong and the mechanical strength of the
immobilized lipase is immobilized lipase is low.
stable.
Entrapment The entrapment conditions This immobilized lipase
are moderate, and the always has the mass
immobilized method transfer restriction during
is applicable to a wide the catalytic process, so
range of carrier and the lipase is only effective
lipases. for low molecular weight
substrates.

whereas some of the solid catalysts were not stable and leached out during
the recycling process. There are also examples of many catalysts that have
been prepared or modified in such a way that its performance and catalytic
activity improved and gave higher biodiesel yield with the requirement of
lower temperature (<353 K), the molar ratio (4:1–6:1), and catalyst loading.
The heterogeneous catalysts, which are easy to prepare and exhibit high
performance under mild reaction conditions, can be considered econom-
ical for biodiesel production. Moreover, the selection of catalyst depends
on the types of biodiesel feedstock. Therefore, SZ, metal oxide from waste
sources, and AC-supported catalysts appear to be the most used catalysts.
124 Biodiesel Technology and Applications

References
1. Hoseini, S.S., Najafi, G., Ghobadian, B., Mamat, R., Ebad, M.T., Yusaf, T.,
Characterization of biodiesel production (ultrasonic-assisted) from evening-
primroses (Oenothera lamarckiana) as novel feedstock and its effect on Cl
engine parameters. Renew. Energy, 130, 50, 2019.
2. Caliskan, H., Environmental and enviroeconomic researches on diesel
engines with diesel and biodiesel fuels. J. Clean. Prod., 154, 125, 2017.
3. Atabani A.E., El-Sheekh, M.M., et al, Edible and nonedible biodiesel feed-
stocks: Microalgae and future of biodiesel, in: Clean Energy for Sustainable
Development, Mohammad G.R., Abul K.A., Subhash C. S., (Ed.), pp. 507–
556, Elsevier Inc., 2017.
4. Gupta, A.R., Yadav, S.V., Rathod, V.K., Enhancement in biodiesel production
using waste cooking oil and calcium diglyceroxide as a heterogeneous cata-
lyst in presence of ultrasound. Fuel, 158, 800, 2015.
5. Gharat, N., Rathod, V.K., Ultrasound assisted enzyme catalyzed transesteri-
fication of waste cooking oil with dimethyl carbonate. Ultrason. Sonochem.,
20, 900, 2013.
6. Maddikeri, G.L., Pandit, A.B., Gogate P.R., Ultrasound assisted interesteri-
fication of waste cooking oil and methyl acetate for biodiesel and triacetin
production. Fuel Process. Technol., 116, 241, 2013.
7. Mansir, N., Taufiq-Yap, Y.H., Rashid, U., Lokman, I. M., Investigation of
heterogeneous solid acid catalyst performance on low grade feedstocks
for biodiesel production: A review. Energy Convers. Manag., 141, 171,
2017.
8. Gupta, A.R., Jalan, A.P., Rathod, V.K., Solar energy as a process intensifica-
tion tool for the biodiesel production from hempseed oil. Energy Convers.
Manag., 171, 126, 2018.
9. Baskar, G., Aiswarya, R., Trends in catalytic production of biodiesel from
various feedstocks. Renew. Sustain. Energy Rev., 57, 496, 2016.
10. Gupta, A.R., Chiplunkar, P.P., Pratap, A.P., Rathod, V. K., Esterification of
Palm Fatty Acid Distillate for FAME Synthesis Catalyzed by Super-Acid
Catalyst HClSO3–ZrO2, Waste Biomass Valori, 2020.
11. Schenk, P.M. Thomas-Hall, S. R., Stephens, E., Marx, U.C., et al, Second
Generation Biofuels: High-Efficiency Microalgae for Biodiesel Production.
BioEnergy Res., 1, 20, 2008.
12. Avinash, A., Subramaniam D., Murugesan, A., Bio-diesel - A global scenario.
Renew. Sustain. Energy Rev., 29, 517, 2014.
13. Basumatary, S., Yellow Oleander (Thevetia peruviana) Seed Oil Biodiesel
as an Alternative and Renewable Fuel for Diesel Engines: A Review. Int. J.
ChemTech Res., 7, 2823, 2015.
14. Owolabi, R.U., Adejumo, A.L., Aderibigbe, A.F., Biodiesel: Fuel for the Future
(A Brief Review). Int. J. Energy Eng., 2, 223, 2012.
Catalyst for Biodiesel 125

15. Da Silva, C., Oliveira, J.V., Biodiesel production through non-catalytic super-
critical transesterification: current state and perspectives. Brazilian J. Chem.
Eng., 31, 2, 271, 2014.
16. Nasreen, S., Nafees, M.L., et al, Review of Catalytic Transesterification
Methods for Biodiesel Production, in: Biofuels - State of Development, B.
Krzysztof (Ed.), pp. 93-119, IntechOpen, 2018.
17. Gupta A.R., Rathod, V.K., Calcium diglyceroxide catalyzed biodiesel produc-
tion from waste cooking oil in the presence of microwave: Optimization and
kinetic studies. Renew. Energy, 121, 757, 2018.
18. Sebastian, J., Muraleedharan, C., Santhiagu, A., Enzyme catalyzed biodiesel
production from rubber seed oil containing high free fatty acid. Int. J. Green
Energy, 14, 687, 2017.
19. Muthukumaran, C., Praniesh, R., Navamani, P., Swathi, R., et al, Process
optimization and kinetic modeling of biodiesel production using non-edible
Madhuca indica oil. Fuel, 195, 217 2017.
20. Singhasiri T., Tantemsapya N., Production of biodiesel from food processing
waste using response surface methodology. Energy Sources, Part A Recover.
Util. Environ. Eff., 38, 2799, 2016.
21. Stamenković, O.S., Veličković, A.V., Kostić, M.D., Joković, N.M., et al,
Optimization of KOH-catalyzed methanolysis of hempseed oil. Energy
Convers. Manag., 103, 235, 2015.
22. Keera, S.T., El Sabagh, S.M., Taman, A.R., Transesterification of vegetable oil
to biodiesel fuel using alkaline catalyst. Fuel, 90, 42, 2011.
23. Chen, K.S., Lin, Y.C., Hsu, K.H., Wang, H. K., Improving biodiesel yields
from waste cooking oil by using sodium methoxide and a microwave heating
system. Energy, 38, 151, 2012.
24. Brito J.Q.A., Silva, C.S., Almeida, J. S., Korn, M.G.A., et al., Ultrasound-
assisted synthesis of ethyl esters from soybean oil via homogeneous catalysis.
Fuel Process. Technol., 95, 33, 2012.
25. Ouanji, F., Kacimi, M., Ziyad, M., Puleo, F., Liotta, L.F., Production of bio-
diesel at small-scale (10 L) for local power generation. Int. J. Hydrogen
Energy, 42, 8914, 2017.
26. Verma, P., Dwivedi, G., Sharma, M.P, Comprehensive analysis on potential
factors of ethanol in Karanja biodiesel production and its kinetic studies.
Fuel, 188 586, 2017.
27. Fadhil, A.B., Ahmed, A.I., Ethanolysis of fish oil via optimized protocol and
purification by dry washing of crude ethyl esters. J. Taiwan Inst. Chem. Eng.,
58,71, 2016.
28. Mohadesi, M., Aghel, B., Maleki, M., Ansari, A., Production of biodiesel
from waste cooking oil using a homogeneous catalyst: Study of semi-
industrial pilot of microreactor. Renew. Energy, 136, 677, 2019.
29. Farag, H.A., El-Maghraby, A., Taha, N.A., Optimization of factors affecting
esterification of mixed oil with high percentage of free fatty acid. Fuel Process.
Technol., 92, 507, 2011.
126 Biodiesel Technology and Applications

30. Soriano Jr., N.U., Venditti, R., Argyropoulos, D.S. Biodiesel synthesis via
homogeneous Lewis acid-catalyzed transesterification. Fuel, 88, 560, 2009.
31. Guan, G., Kusakabe, K., Sakurai, N., Moriyama, K., Transesterification of
vegetable oil to biodiesel fuel using acid catalysts in the presence of dimethyl
ether. Fuel, 88, 81, 2009.
32. Kumar, A., Sharma, S., Potential non-edible oil resources as biodiesel feed-
stock: An Indian perspective. Renew. Sustain. Energy Rev., 15, 1791, 2011.
33. Jitputti, J., Kitiyanan, B., Rangsunvigit, P., Bunyakiat, K., et al,
Transesterification of crude palm kernel oil and crude coconut oil by differ-
ent solid catalysts. Chem. Eng. J., 116, 61, 2006.
34. Lam, M.K., Lee, K.T., Production of biodiesel using palm oil, in: Biofuels:
Alternative Feedstocks and Conversion Processes, A. Pandey, C. Larroche,
S.C. Ricke, C.G. Dussap, E. Gnansounou, (Ed.), pp. 353-374, Elsevier Inc.,
2011.
35. Tan, C.H., Nagarajan, D., et al, Biodiesel from Microalgae, in: Biofuels:
Alternative Feedstocks and Conversion Processes for the Production of Liquid
and Gaseous Biofuels, A. Pandey, C. Larroche, C.G.Dussap, E. Gnansounou,
S.K. Khanal, S. Ricke, (Ed.), pp. 601–628 Elsevier Inc., 2019.
36. Avhad, M.R., Marchetti, J.M., Innovation in solid heterogeneous catalysis for
the generation of economically viable and ecofriendly biodiesel: A review.
Catal. Rev. Sci. Eng., 58, 157, 2016.
37. Lotero, E., Liu, Y., Lopez, D.E., Suwannakarn, K., et al, Synthesis of biodiesel
via acid catalysis. Ind. Eng. Chem. Res., 44, 5353, 2005.
38. Bouaid, A., El Boulifi, N., Martinez, M., Aracil, J., Optimization of a two-step
process for biodiesel production from Jatropha curcas crude oil. Int. J. Low-
Carbon Technol., 7, 331, 2012.
39. Rahman, M.A., Aziz, M.A., Al-khulaidi, R.A., Sakib, N., et al, Biodiesel
production from microalgae S pirulina maxima by two step process:
Optimization of process variable. J. Radiat. Res. Appl. Sci., 10, 140, 2017.
40. Canakci, M., Van Gerpen J., Biodiesel production via acid catalysis. Trans.
Am. Soc. Agric. Eng., 42, 1203, 1999.
41. Talha, N.S., Sulaiman, S., Overview of catalysts in biodiesel production.
ARPN J. Eng. Appl. Sci., 11, 439, 2016.
42. Thangaraj, B., Solomon, P.R., Muniyandi, B., Ranganathan, S., et al, Catalysis
in biodiesel production - A review. Clean Energy, 3, 2, 2019.
43. Kayode, B., Hart, A., An overview of transesterification methods for produc-
ing biodiesel from waste vegetable oils. Biofuels, 10, 419, 2019.
44. Hariprasath, P., Selvamani, S.T., Vigneshwar, M., Palanikumar, K., et al,
Comparative analysis of cashew and canola oil biodiesel with homogeneous
catalyst by transesterification method. Mater. Today Proc., 16, 1357, 2019.
45. Georgogianni, K.G., Katsoulidis, A.K., Pomonis, P.J., Manos, G., et al,
Transesterification of rapeseed oil for the production of biodiesel using
homogeneous and heterogeneous catalysis. Fuel Process. Technol., 90, 1016,
2009.
Catalyst for Biodiesel 127

46. Refaat, A.A., Different techniques for the production of biodiesel from waste
vegetable oil. Int. J. Environ. Sci. Technol., 7, 183, 2010.
47. Islam, A., Taufiq-Yap, Y.H., Chan, E.S., Moniruzzaman, M., et al, Advances
in solid-catalytic and non-catalytic technologies for biodiesel production.
Energy Convers. Manag., 88, 1200, 2014.
48. Islam, A., Taufiq-Yap, Y.H., Chu, C.M., Chan, E.S., et al, Synthesis and char-
acterization of millimetric gamma alumina spherical particles by oil drop
granulation method. J. Porous Mater., 19, 807, 2012.
49. Rashid, U., Soltani, S., et al, Metal oxide catalysts for biodiesel production, in:
Metal Oxides in Energy Technologies, Y. Wu (Ed.), pp. 303–319, Elsevier Inc.,
2018.
50. Park, Y.M., Lee, D.W., Kim, D.K., Lee, J.S., The heterogeneous catalyst system
for the continuous conversion of free fatty acids in used vegetable oils for the
production of biodiesel. Catal. Today, 131, 238, 2008.
51. Ramos, L.P., Cordeiro, C.S., et al, Applications of Heterogeneous Catalysts
in the Production of Biodiesel by Esterification and Transesterification, in:
Bioenergy Research: Advances and Applications, V.K.Gupta, M.G. Tuohy, C.P.
Kubicek, J. Saddler, F. Xu, (Ed.), pp. 255–276, Elsevier Inc., 2014.
52. Dhawane, S.H., Halder, G., Synthesis of Catalyst Support From Waste
Biomass for Impregnation of Catalysts in Biofuel Production, in: Advances
in Feedstock Conversion Technologies for Alternative Fuels and Bioproducts,
M. Hosseini (Ed.), pp. 199-220, Elsevier Inc., 2019.
53. Fu, J., Chen, L., Lv, P., Yang, L., Yuan, Z., Free fatty acids esterification for bio-
diesel production using self-synthesized macroporous cation exchange resin
as solid acid catalyst. Fuel, 154, 1, 2015.
54. Yan, G.X., Wang, A., Wachs, I. E., Baltrusaitis, J., Critical review on the active
site structure of sulfated zirconia catalysts and prospects in fuel production.
Appl. Catal. A Gen., 572, 210, 2019.
55. Farcasiu, D., Li, J.Q., Cameron, S., Preparation of sulfated zirconia catalysts
with improved control of sulfur content II. Effect of sulfur content on physi-
cal properties and catalytic activity. Appl. Catal. A Gen., 154, 173, 1997.
56. Zhang, Y., Wong, W.T., Yung, K.F., One-step production of biodiesel from
rice bran oil catalyzed by chlorosulfonic acid modified zirconia via simulta-
neous esterification and transesterification. Bioresour. Technol., 147, 59, 2013.
57. Echaroj, S., Santikunaporn, M., Chavadej, S., Oligomerization of 1-decene
over sulfated alumina catalysts for the production of synthetic fuels and
lubricants: modelling and verification. React. Kinet. Mech. Catal., 121, 629,
2017.
58. Vahid, B.R., Saghatoleslami, N., Nayebzadeh, H., Toghiani, J., Effect of alu-
mina loading on the properties and activity of SO42−/ZrO2 for biodiesel pro-
duction: Process optimization via response surface methodology. J. Taiwan
Inst. Chem. Eng., 83, 115, 2018.
59. Kashyap, S.S., Gogate, P.R., Joshi, S.M., Ultrasound assisted intensi-
fied production of biodiesel from sustainable source as karanja oil using
128 Biodiesel Technology and Applications

interesterification based on heterogeneous catalyst (γ-alumina). Chem. Eng.


Process., 136, 11, 2019.
60. Ramesh, A., Palanichamy, K., Tamizhdurai, P., Umasankar, S., et al, Sulphated
Zr–Al2O3 catalysts through jatropha oil to green-diesel production. Mater.
Lett., 238, 62, 2019.
61. Thielemann, J.P., Girgsdies, F., Schlögl, R., Hess, C., Pore structure and sur-
face area of silica SBA-15: Influence of washing and scale-up. Beilstein J.
Nanotechnol., 2, 110, 2011.
62. Yu, C., Fan, J., Tian, B., Zhao, D., Morphology Development of Mesoporous
Materials: A Colloidal Phase Separation Mechanism. Chem. Mater., 16, 889,
2004.
63. Zhao, D. Feng, J., Huo, Q., Melosh, N., et al, Triblock copolymer syntheses of
mesoporous silica with periodic 50 to 300 angstrom pores. Science, 279, 548,
1998.
64. Thitsartarn, W., Kawi, S., Transesterification of Oil by Sulfated Zr-Supported
Mesoporous Silica. Ind. Eng. Chem. Res., 50, 7857, 2011.
65. Shao, G.N., Sheikh, R., Hilonga, A., Lee, J. E., et al, Biodiesel production by
sulfated mesoporous titania-silica catalysts synthesized by the sol-gel process
from less expensive precursors. Chem. Eng. J., vol. 215–216, 600, 2013.
66. Fan, M., Si, Z., Sun, W., Zhang, P., Sulfonated ZrO2-TiO2 nanorods as effi-
cient solid acid catalysts for heterogeneous esterification of palmitic acid.
Fuel, 252, 254, 2019.
67. Guldhe, A., Singh, B., Rawat, I., Bux, F., Synthesis of biodiesel from
Scenedesmus sp. by microwave and ultrasound assisted in situ transesterifi-
cation using tungstated zirconia as a solid acid catalyst. Chem. Eng. Res. Des.,
92, 1503, 2014.
68. Peruzzolo, T.M., Stival, J.F., Baika, L.M., Ramos, L.P., et al, Efficient esteri-
fication reaction of palmitic acid catalyzed by WO 3-x/mesoporous silica.
Biofuels, 2020.
69. Zhao, J., Guan, H., Shi, W., Cheng, M., et al, A Brønsted-Lewis-surfactant-
combined heteropolyacid as an environmental benign catalyst for esterifica-
tion reaction. Catal. Commun., 20, 103, 2012.
70. Sudhakar, P., Pandurangan, A., Heteropolyacid (H3PW12O40)-impregnated
mesoporous KIT-6 catalyst for green synthesis of bio-diesel using transester-
ification of non-edible neem oil. Mater. Renew. Sustain. Energy, 8, 22, 2019.
71. de Oliveira, A. de N., de Lima, M.A.B., Pires, L.H.de O., da Silva, M.R.,
et al, Bentonites modified with phosphomolybdic heteropolyacid (HPMo)
for biowaste to biofuel production. Materials (Basel)., 12, 2019.
72. Pandian, S., Sakthi, S.A., et al, Application of heterogeneous acid catalyst
derived from biomass for biodiesel process intensification: a comprehensive
review, in: Refining Biomass Residues for Sustainable Energy and Bioproducts,
R.P. Kumar, E. Gnansounou, J K.Raman, G. Baskar, (Ed.), pp. 87–109,
Elsevier Inc., 2020.
Catalyst for Biodiesel 129

73. Savaliya, M.L., Dholakiya, B.Z., A simpler and highly efficient protocol for
the preparation of biodiesel from soap stock oil using a BBSA catalyst. RSC
Adv., 5, 74416, 2015.
74. Thushari, I., Babel, S., Sustainable utilization of waste palm oil and sulfon-
ated carbon catalyst derived from coconut meal residue for biodiesel produc-
tion. Bioresour. Technol., 248, 199, 2018.
75. Guo, M.L., Yin, X.Y., Huang, J., Preparation of novel carbonaceous solid
acids from rice husk and phenol. Mater. Lett., 196, 23, 2017.
76. Konwar, L.J., Mäki-Arvela, P., Salminen, E., Kumar, N., et al, Towards carbon
efficient biorefining: Multifunctional mesoporous solid acids obtained from
biodiesel production wastes for biomass conversion. Appl. Catal. B Environ.,
176, 20, 2015.
77. Zeng, D., Liu, S., Gong, W., Wang, G., et al, Synthesis, characterization and
acid catalysis of solid acid from peanut shell. Appl. Catal. A Gen., 469, 284,
2014.
78. Lou, W.Y., Guo, Q., Chen, W.J., Zong, M. H., et al, A highly active bagasse-
derived solid acid catalyst with properties suitable for production of bio-
diesel. ChemSusChem, 5, 1533, 2012.
79. Endut, A., Abdullah, S.H., Yasmin, S., Hanapi, N.H.M., et al., Optimization
of biodiesel production by solid acid catalyst derived from coconut shell via
response surface methodology. Int. Biodeterior. Biodegrad., 124, 250, 2017.
80. Mardhiah, H.H., Ong, H.C., Masjuki, H.H., Lim, S., et al, Investigation of
carbon-based solid acid catalyst from Jatropha curcas biomass in biodiesel
production. Energy Convers. Manag., 144, 10, 2017.
81. Wang, L., Dong, X., Jiang, H., Li, G., et al, Preparation of a novel carbon-based
solid acid from cassava stillage residue and its use for the esterification of free
fatty acids in waste cooking oil. Bioresour. Technol., 158, 392, 2014.
82. Dehkhoda, A.M., West, A.H., Ellis, N., Biochar based solid acid catalyst for
biodiesel production. Appl. Catal. A Gen., 382, 197, 2010.
83. Shu, Q., Zhang, Q., Xu, G., Nawaz, Z., et al, Synthesis of biodiesel from
cottonseed oil and methanol using a carbon-based solid acid catalyst. Fuel
Process. Technol., 90, 1002, 2009.
84. Dawodu, F.A., Ayodele, O.O., Xin, J., Zhang, S., Application of solid acid
catalyst derived from low value biomass for a cheaper biodiesel production.
J. Chem. Technol. Biotechnol., 89, 1898, 2014.
85. Huang, M., Luo, J., Fang, Z., Li, H., Biodiesel production catalyzed by highly
acidic carbonaceous catalysts synthesized via carbonizing lignin in sub- and
super-critical ethanol. Appl. Catal. B Environ., 190, 103, 2016.
86. Chin, L.H., Abdullah, A.Z., Hameed B.H., Sugar cane bagasse as solid cata-
lyst for synthesis of methyl esters from palm fatty acid distillate. Chem. Eng.
J., 183, 104, 2012.
87. Yu, J.T., Dehkhoda, A. M., Ellis, N., Development of Biochar-based Catalyst
for Transesterification of Canola Oil. Energy & Fuels, 25, 337, 2011.
130 Biodiesel Technology and Applications

88. Doyle, A.M., Albayati, T.M., Abbas, A. S., Alismaeel, Z.T., Biodiesel pro-
duction by esterification of oleic acid over zeolite Y prepared from kaolin,
Renew. Energy, 97, 19, 2016.
89. Patel, A., Narkhede, N., 12-tungstophosphoric acid anchored to zeolite Hβ:
Synthesis, characterization, and biodiesel production by esterification of
oleic acid with methanol. Energy and Fuels, 26, 6025, 2012.
90. Su, F., Guo, Y., Advancements in solid acid catalysts for biodiesel production.
Green Chem., 16, 2934, 2014.
91. Park, J.Y., Kim, D.K., Lee, J.S., Esterification of free fatty acids using water-
tolerable Amberlyst as a heterogeneous catalyst. Bioresour. Technol., 101,
S62, 2010.
92. Tesser, R., Casale, L., Verde, D., Serio, M.D., et al, Kinetics and modeling
of fatty acids esterification on acid exchange resins. Chem. Eng. J., 157, 539,
2010.
93. Pasa, T.L.B., Souza, G.K., Yamaguchi, N.U., Pomini, A.M., et al, Esterification
of crude ‘Macaúba’ oil (Acrocomia aculeata) using amberlyst 15 as a hetero-
geneous catalyst performed in reactor parr. Chem. Eng. Trans., 65, 571, 2018.
94. Hykkerud, A., Marchetti, J.M., Esterification of oleic acid with ethanol in the
presence of Amberlyst 15. Biomass and Bioenergy, 95, 340, 2016.
95. Ilgen, O., Investigation of reaction parameters, kinetics and mechanism of
oleic acid esterification with methanol by using Amberlyst 46 as a catalyst.
Fuel Process. Technol., 124, 134, 2014.
96. López, D.E., Goodwin, J.G., Bruce, D.A., Transesterification of triacetin with
methanol on Nafion acid resins. J. Catal., 245, 381, 2007.
97. Russbueldt, B.M.E., Hoelderich, W.F., New sulfonic acid ion-exchange resins
for the preesterification of different oils and fats with high content of free
fatty acids. Appl. Catal. A Gen., 362, 47, 2009.
98. Andrijanto, E., Dawson, E.A., Brown, D.R., Hypercrosslinked polystyrene
sulphonic acid catalysts for the esterification of free fatty acids in biodiesel
synthesis. Appl. Catal. B Environ., 115, 261, 2012.
99. Romero, R., Luz, S., et al, Biodiesel Production by Using Heterogeneous
Catalysts, in: Alternative Fuel, M. Manzanera (Ed.), pp. 2151–2161,
IntechOpen, 2011,
100. Coman, S.M. Parvulescu, V.I., Heterogeneous Catalysis for Biodiesel
Production, in: The Role of Catalysis for the Sustainable Production of Bio-
Fuels and Bio-Chemicals, K.S. Triantafyllidis, A.A. Lappas, M. Stöcker, (Ed.),
pp. 93-136, Elsevier Inc., 2013.
101. Marwaha, A., Dhir, A., Mahla, S.K., Mohapatra, S.K., An overview of solid
base heterogeneous catalysts for biodiesel production. Catal. Rev. - Sci. Eng.,
60, 594, 2018.
102. Baroi, C., Yanful, E.K., Bergougnou, M.A., Biodiesel Production from
Jatropha curcas Oil Using Potassium Carbonate as an Unsupported Catalyst.
Int. J. Chem. React. Eng., 7, 2009.
Catalyst for Biodiesel 131

103. Chew, K.Y., Tan, W.L., Abu Bakar, N.H.H., Abu Bakar, M., Transesterification
of palm cooking oil using barium-containing titanates and their sodium
doped derivatives. Int. J. Energy Environ. Eng., 8, 47, 2017.
104. Lee, H V., Juan, J.C., Taufiq-Yap, Y.H., Kong, P.S., et al, Advancement in het-
erogeneous base catalyzed technology: An efficient production of biodiesel
fuels. J. Renew. Sustain. Energy, 7, 2015.
105. Dall’Oglio, E.L., De Sousa Jr., P.T., De Jesus Oliveira, P.T., De Vasconcelos,
L.G., et al, Use of heterogeneous catalysts in methylic biodiesel production
induced by microwave irradiation. Quim. Nova, 37,411, 2014.
106. Mallick, P.K., Particulate and Short Fiber Reinforced Polymer Composites,
in: Comprehensive Composite Materials, A. Kelly, C. Zweben, (Ed.), pp. 291–
331, Elsevier Inc., 2000.
107. Reddy, A.N.R., Ahmed, A.S., Islam, M.D., Hamdan, S., Methanolysis of
Crude Jatropha Oil using Heterogeneous Catalyst from the Seashells and
Eggshells as Green Biodiesel. ASEAN J. Sci. Technol. Dev., 32, 16, 2017.
108. Widiarti, N., Ni’mah, Y.L., Bahruji, H., Prasetyoko, D., Development of CaO
from natural calcite as a heterogeneous base catalyst in the formation of bio-
diesel: Review. J. Renew. Mater., 7, 915, 2019.
109. Buasri, A., Chaiyut, N., Loryuenyong, V., Worawanitchaphong, P., et al,
Calcium oxide derived from waste shells of mussel, cockle, and scallop as the
heterogeneous catalyst for biodiesel production, Sci. World J., 2013, 2013.
110. Goli, J., Sahu, O., Development of heterogeneous alkali catalyst from waste
chicken eggshell for biodiesel production. Renew. Energy, 128, 142, 2018.
111. Ajala, E.O., Ajala, M.A., Odetoye, T.E., Aderibigbe, F.A., et al, Thermal mod-
ification of chicken eggshell as heterogeneous catalyst for palm kernel bio-
diesel production in an optimization process. Biomass Convers. Biorefinery,
2020.
112. Rahman, W.U., Fatima, A., Anwer, A.H., Athar, M., et al, Biodiesel synthesis
from eucalyptus oil by utilizing waste egg shell derived calcium based metal
oxide catalyst. Process Saf. Environ. Prot., 122, 313, 2019.
113. Gupta, A. R., Rathod,V. K., Waste cooking oil and waste chicken eggshells
derived solid base catalyst for the biodiesel production: Optimization and
kinetics. Waste Manag., 79, 169, 2018.
114. Syazwani, O.N., Teo S.H., Islam, A., Taufiq-Yap, Y.H., Transesterification
activity and characterization of natural CaO derived from waste venus clam
(Tapes belcheri S.) material for enhancement of biodiesel production. Process
Saf. Environ. Prot., 105, 303, 2017.
115. Anjana, P.A., Niju, S., Meera Sheriffa Begum, K.M., Anantharaman, N.,
Utilization of limestone derived calcium oxide for biodiesel production from
non-edible pongamia oil. Environ. Prog. Sustain. Energy, 35, 1758, 2016.
116. Maneerung, T., Kawi, S., Dai, Y., Wang, C.H., Sustainable biodiesel produc-
tion via transesterification of waste cooking oil by using CaO catalysts pre-
pared from chicken manure. Energy Convers. Manag., 123, 487, 2016.
132 Biodiesel Technology and Applications

117. Tan, Y.H., Abdullah, M.O., Nolasco-Hipolito, C., Taufiq-Yap, Y.H., Waste
ostrich- and chicken-eggshells as heterogeneous base catalyst for biodiesel
production from used cooking oil: Catalyst characterization and biodiesel
yield performance. Appl. Energy, 160, 58, 2015.
118. Perea, A., Kelly, T., Hangun-Balkir, Y., Utilization of waste seashells and
Camelina sativa oil for biodiesel synthesis. Green Chem. Lett. Rev., 9, 27,
2016.
119. Yoosuk, .B, Udomsap, P., Puttasawat, B., Krasae, P., Modification of calcite
by hydration-dehydration method for heterogeneous biodiesel production
process: The effects of water on properties and activity. Chem. Eng. J., 162,
135, 2010.
120. Niju S., Meera S. Begum, K.M., Anantharaman, N., Modification of egg shell
and its application in biodiesel production. J. Saudi Chem. Soc., 18, 702, 2014.
121. Anr, R., Saleh, A.A., Islam, M.S., Hamdan S., et al, Biodiesel Production from
Crude Jatropha Oil using a Highly Active Heterogeneous Nanocatalyst by
Optimizing Transesterification Reaction Parameters. Energy and Fuels, 30,
334, 2016.
122. Tang, Y., Wang, S., Cheng, X., Lu, Y., Efficient heterogeneous catalyst for bio-
diesel production from soybean oil over modified CaO. Prog. React. Kinet.
Mech., 39, 273, 2014.
123. Lukić, I., Kesić, Ž., Zdujić, M., Skala, D., Calcium diglyceroxide synthesized
by mechanochemical treatment, its characterization and application as cata-
lyst for fatty acid methyl esters production. Fuel, 165, 159, 2016.
124. Chang, F., Zhou, Q., Pan, H., Liu, X.F., et al, Solid Mixed-Metal-Oxide
Catalysts for Biodiesel Production: A Review. Energy Technol., 2, 865,
2014.
125. Lee, H.V., Taufiq-Yap, Y.H., Hussein, M.Z., Yunus, R., Transesterification of
jatropha oil with methanol over Mg-Zn mixed metal oxide catalysts. Energy,
49, 12, 2013.
126. Taufiq-Yap, Y.H., Lee, H.V., Hussein, M.Z., Yunus, R., Calcium-based mixed
oxide catalysts for methanolysis of Jatropha curcas oil to biodiesel. Biomass
and Bioenergy, 35, 827, 2011.
127. Lee, H.V., Juan, J.C., Binti Abdullah, N.F., Nizah MF, R., et al, Heterogeneous
base catalysts for edible palm and non-edible Jatropha-based biodiesel pro-
duction. Chem. Cent. J., 8, 2014.
128. Helwani, Z., Othman, M.R., Aziz, N., Kim, J., et al, Solid heterogeneous cat-
alysts for transesterification of triglycerides with methanol: A review. Appl.
Catal. A Gen., 363, 1, 2009.
129. Endalew, A.K., Kiros, Y., Zanzi, R., Inorganic heterogeneous catalysts for
biodiesel production from vegetable oils. Biomass and Bioenergy, 35, 3787,
2011.
130. Silva, C.C.C.M., Ribeiro, N.F.P., Souza, M.M.V.M., Aranda, D.A.G., Biodiesel
production from soybean oil and methanol using hydrotalcites as catalyst,
Fuel Process. Technol., 91, 205, 2010.
Catalyst for Biodiesel 133

131. Navajas, A., Campo, I., Moral, A., Echave, J., et al, Outstanding performance
of rehydrated Mg-Al hydrotalcites as heterogeneous methanolysis catalysts
for the synthesis of biodiesel. Fuel, 211, 173, 2018.
132. Fatimah, I., Rubiyanto, D., Nugraha, J., Preparation, characterization, and
modelling activity of potassium flouride modified hydrotalcite for micro-
wave assisted biodiesel conversion. Sustain. Chem. Pharm., 8, 63, 2018.
133. Osman, A.I., Abu-Dahrieh, J.K., Rooney, D.W., Halawy, S.A., et al, Effect of
precursor on the performance of alumina for the dehydration of methanol to
dimethyl ether. Appl. Catal. B Environ., 127, 307, 2012.
134. Boz, N., Degirmenbasi, N., Kalyon, D.M., Conversion of biomass to fuel:
Transesterification of vegetable oil to biodiesel using KF loaded nano-γ-
Al2O3 as catalyst. Appl. Catal. B Environ., 89, 590, 2009.
135. Noiroj, K., Intarapong, P., Luengnaruemitchai, A., Jai-In, S., A comparative
study of KOH/Al2O3 and KOH/NaY catalysts for biodiesel production via
transesterification from palm oil. Renew. Energy, 34, 1145, 2009.
136. Boz, N., Kara, M., Solid base catalyzed transesterification of canola oil, Chem.
Eng. Commun., 196, 80, 2009.
137. Zabeti, M., Daud, W.M.A.W., Aroua, M.K., Biodiesel production using
alumina-supported calcium oxide: An optimization study. Fuel Process.
Technol., 91, 243, 2010.
138. De Lima, A.L., Mbengue, A., San Gil, R.A.S., Ronconi C.M., et al, Synthesis
of amine-functionalized mesoporous silica basic catalysts for biodiesel pro-
duction. Catal. Today, 226, 210, 2014.
139. Samart, C., Sreetongkittikul, P., Sookman C., Heterogeneous catalysis of
transesterification of soybean oil using KI/mesoporous silica. Fuel Process.
Technol., 90, 922, 2009.
140. Xie, W., Zhao L., Heterogeneous CaO-MoO3-SBA-15 catalysts for biodiesel
production from soybean oil. Energy Convers. Manag., 79, 34, 2014.
141. Alba-Rubio, A.C., Santamaría-González, J., Mérida-Robles, J.M., Moreno-
Tost, R., et al, Heterogeneous transesterification processes by using CaO
supported on zinc oxide as basic catalysts. Catal. Today, 149, 281, 2010.
142. Malani, R.S., Singh, S., Goyal, A., Moholkar, V.S., Ultrasound-Assisted
Biodiesel Production Using KI-Impregnated Zinc Oxide (ZnO) as
Heterogeneous Catalyst: A Mechanistic Approach, in: Conference Proceedings
of the Second International Conference on Recent Advances in Bioenergy
Research, 2018, pp. 67–81.
143. Thitsartarn, W., Maneerung, T., Kawi, S., Highly active and durable Ca-doped
Ce-SBA-15 catalyst for biodiesel production. Energy, 89, 946, 2015.
144. Bilal, M., Mehmood, S., Rasheed, T., Iqbal, H.M.N., Bio-Catalysis and
Biomedical Perspectives of Magnetic Nanoparticles as Versatile Carriers.
Magnetochemistry, 5, 42, 2019.
145. Rossi, L.M., Costa, N.J.S., Silva, F.P., Wojcieszak, R., Magnetic nanomaterials
in catalysis: advanced catalysts for magnetic separation and beyond. Green
Chem., 16, 2906, 2014.
134 Biodiesel Technology and Applications

146. Chiang, Y.D., Dutta, S., Chen, C.T., Huang, Y.T., et al, Functionalized Fe3O4@
Silica Core-Shell Nanoparticles as Microalgae Harvester and Catalyst for
Biodiesel Production, ChemSusChem, 8, 789, 2015.
147. Liu, Y., Zhang, P., Fan, M., Jiang, P., Biodiesel production from soybean oil
catalyzed by magnetic nanoparticle MgFe2O4@CaO. Fuel, 164, 314, 2016.
148. Bahuguna, A., Kumar, A., Krishnan, V., Carbon-Support-Based
Heterogeneous Nanocatalysts: Synthesis and Applications in Organic
Reactions. Asian J. Org. Chem., 8, 1263, 2019.
149. Jain, A., Balasubramanian, R., Srinivasan, M.P., Hydrothermal conversion of
biomass waste to activated carbon with high porosity: A review. Chem. Eng.
J., 283, 789, 2016.
150. Narowska, B., Kułażyński, M., Łukaszewicz, M., Burchacka, E., Use of acti-
vated carbons as catalyst supports for biodiesel production. Renew. Energy,
135, 176, 2019.
151. Fadhil, A.B., Aziz, A.M., Altamer, M.H., Optimization of methyl esters pro-
duction from non-edible oils using activated carbon supported potassium
hydroxide as a solid base catalyst. Arab J. Basic Appl. Sci., 25, 56, 2018.
152. Wan, Z., Hameed, B.H., Mohammad Nor, N., Ali Bashah, N.A., Optimization
of methyl ester production from waste palm oil using activated carbon sup-
ported calcium oxide catalyst. Solid State Phenom., 280, 346, 2018.
153. Peng, F., Yin, H., et al, Enzyme nanocarriers, in: Advances in Enzyme
Technology, R.S. Singh, R.R. Singhania, A. Pandey, C. Larroche, (Ed.), pp.
153–168, Elsevier B.V., 2019.
154. Chapman, J., Ismail, A.E., Dinu, C.Z., Industrial applications of enzymes:
Recent advances, techniques, and outlooks. Catalysts, 8, 1, 2018.
155. Ondul, E., Dizge, N., et al, Biocatalytic Production of Biodiesel from
Vegetable Oils, in: Biofuels - Status and Perspective, K. Biernat (Ed.), pp.
21-37 IntechOpen, 2015.
156. Atadashi, I.M., Aroua, M.K., Aziz, A.R.A., Sulaiman, N.M.N., The effects of
catalysts in biodiesel production: A review. J. Ind. Eng. Chem., 19, 14, 2013.
157. Liu, D.M., Chen, J., Shi, Y.P., Advances on methods and easy separated sup-
port materials for enzymes immobilization. Trends Anal. Chem., 102, 332,
2018.
158. Moazeni, F., Chen, Y.C., Zhang, G., Enzymatic transesterification for bio-
diesel production from used cooking oil, a review. J. Clean. Prod., 216, 117,
2019.
159. Khosla, K., Rathour, R., Maurya, R., Maheshwari, N., et al, Biodiesel produc-
tion from lipid of carbon dioxide sequestrating bacterium and lipase of psy-
chrotolerant Pseudomonas sp. ISTPL3 immobilized on biochar. Bioresour.
Technol., vol. 245,743, 2017.
160. Kumar, D., Das, T., Giri, B.S., Verma, B., Optimization of biodiesel synthe-
sis from nonedible oil using immobilized bio-support catalysts in jacketed
packed bed bioreactor by response surface methodology. J. Clean. Prod., 244,
2020.
Catalyst for Biodiesel 135

161. Jegannathan, K.R., Jun-Yee, L., Chan, E.S., Ravindra, P., Production of bio-
diesel from palm oil using liquid core lipase encapsulated in κ-carrageenan.
Fuel, 89, 2272, 2010.
162. Nadar, S.S., Rathod, V.K., Encapsulation of lipase within metal-organic
framework (MOF) with enhanced activity intensified under ultrasound,
Enzyme Microb. Technol., 108, 11, 2018.
163. Nematian, T., Salehi, Z., Shakeri,A., Conversion of bio-oil extracted from
Chlorella vulgaris micro algae to biodiesel via modified superparamagnetic
nano-biocatalyst. Renew. Energy, 146, 1796, 2020.
164. Bayramoglu, G., Akbulut, A., Ozalp, V.C., Arica, M.Y., Immobilized lipase on
micro-porous biosilica for enzymatic transesterification of algal oil. Chem.
Eng. Res. Des., 95, 12, 2015.
165. Badoei-dalfard, A., Malekabadi, S., Karami, Z., Sargazi, G., Magnetic cross-
linked enzyme aggregates of Km12 lipase: A stable nanobiocatalyst for bio-
diesel synthesis from waste cooking oil. Renew. Energy, 141, 874, 2019.
166. Kim, S.H., Kim, S.J., Park, S., Kim, H.K., Biodiesel production using cross-
linked Staphylococcus haemolyticus lipase immobilized on solid polymeric
carriers. J. Mol. Catal. B Enzym., 85, 10, 2013.
167. Kharrat, N., Ben Ali, Y., Marzouk, S., Gargouri, Y.T., Immobilization of
Rhizopus oryzae lipase on silica aerogels by adsorption: Comparison with
the free enzyme. Process Biochem., 46, 1083, 2011.
168. Naranjo, J.C., Córdoba, A., Giraldo, L., García, V.S., et al, Lipase supported
on granular activated carbon and activated carbon cloth as a catalyst in the
synthesis of biodiesel fuel. J. Mol. Catal. B Enzym., 66, 166, 2010.
169. Salis, A., Pinna, M., Monduzzi, M., Solinas, V., Comparison among immo-
bilised lipases on macroporous polypropylene toward biodiesel synthesis.
J. Mol. Catal. B Enzym., 54, 19, 2008.
170. Yagiz, F., Kazan, D., Akin, A.N., Biodiesel production from waste oils by
using lipase immobilized on hydrotalcite and zeolites. Chem. Eng. J., 134,
262, 2007.
171. Babaki, M., Yousefi, M., Habibi, Z., Brask, J., et al, Preparation of highly reus-
able biocatalysts by immobilization of lipases on epoxy-functionalized silica
for production of biodiesel from canola oil. Biochem. Eng. J., 101, 23, 2015.
172. Picó, E.A., López, C., Cruz-Izquierdo, Á., Munarriz, M., et al, Easy reuse of
magnetic cross-linked enzyme aggregates of lipase B from Candida antarc-
tica to obtain biodiesel from Chlorella vulgaris lipids. J. Biosci. Bioeng., 126,
451, 2018.
173. Wang, Y. D., Shen, X.Y., Li, Z.L., Li, X., et al, Immobilized recombinant
Rhizopus oryzae lipase for the production of biodiesel in solvent free system.
J. Mol. Catal. B Enzym., 67, 45, 2010.
174. Nassreddine, S., Karout, A., Lorraine Christ, M., Pierre, A.C.,
Transesterification of a vegetal oil with methanol catalyzed by a silica fibre
reinforced aerogel encapsulated lipase. Appl. Catal. A Gen., 344, 70, 2008.
136 Biodiesel Technology and Applications

175. Meunier, S.M., Legge, R.L., Evaluation of diatomaceous earth as a support for
sol-gel immobilized lipase for transesterification. J. Mol. Catal. B Enzym., 62,
54, 2010.
176. Zhang, H., Liu, T., Zhu, Y., Hong, L., et al, Lipases immobilized on the modi-
fied polyporous magnetic cellulose support as an efficient and recyclable cat-
alyst for biodiesel production from Yellow horn seed oil. Renew. Energy, 145,
1246, 2020.
177. Tan, T., Lu, J., Nie, K., Deng, L., Wang, F., Biodiesel production with immo-
bilized lipase: A review. Biotechnol. Adv., 28, 628, 2010.
178 Zhang, B., Weng, Y., Xu, H., Mao, Z., Enzyme immobilization for biodiesel
production, Appl. Microbiol. Biotechnol., 93, 61, 2012.
4
Hydrogenolysis as a Means of Valorization
of Biodiesel-Derived Glycerol: A Review
Manjoro T.T.*, Adeniyi A. and Mbaya R.K.K.

Department of Chemical, Metallurgical and Materials Engineering,


Tshwane University of Technology, Pretoria, South Africa

Abstract
The hydrogenolysis of glycerol to 1,2-propanediol is one of the most promis-
ing routes for producing value-added chemicals from glycerol, a by-product of
­biodiesel manufacturing. This process will improve the economics of the bio-
diesel manufacturing process that is currently costly compared to conventional
diesel from petroleum. Successful full industrialization of such a process will
shift the dependency on petroleum-based chemicals, thereby promoting renew-
able and sustainable energy usage. The conversion of glycerol proceeds faster in
a basic medium compared to a neutral one, in a helium atmosphere as compared
to under a hydrogen atmosphere and temperature lower than 200°C. Rhenium
was seen as the most effective catalyst in the selective catalytic transformation
of glycerol to 1,2-DPO. However, matter organic non-glycerol, non-methanol
(MONG-NM) impurities, and sulfur compounds must be removed from the
feedstock because they are detrimental to catalyst performance. The reaction
time needs to be kept low as increase in the reaction time does not lead to higher
selectivity of 1,2-propanediol. The reaction mechanism still needs to be estab-
lished to facilitate an effective process optimization to improve the selectivity of
1,2-propanediol.

Keywords: Hydrogenolysis, biodiesel, glycerol valorization, 1,2-propanediol,


heterogeneous catalyst

*Corresponding author: tendaimanjoro038@gmail.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (137–166) © 2021 Scrivener Publishing LLC

137
138 Biodiesel Technology and Applications

4.1 Introduction
Climate change is one of the 21st century’s greatest challenges. The fourth
assessment report of the Intergovernmental Panel on Climate change
(IPCC) identified human everyday activity as one of the major causes
of global warming [1]. The business as usual approach adopted thus far
by some countries, apart from being unproductive, may lead to a disas-
trous transformation of planet earth, and recent scientific discoveries
emphasize the growing urgency to the issue of reducing greenhouse gas
emissions [2]. One of the ways for reducing global warming is through
the use of renewable and sustainable energy, thus moving away from the
combustion of fossil fuels. Biodiesel is one of the promising renewable,
liquid fuel, and energy sources which is poised to compete and to a large
extent replace petroleum-based diesel fuel. The challenge at the present
moment that is hindering a wider implementation of the biodiesel tech-
nology and its subsequent full-scale industrialization, thus competing
with the traditional petroleum-based (or Fischer-Tropsch–based) die-
sel fuel, is the high cost involved in the currently used manufacturing
process.
The development and distribution of suitable renewable energy technol-
ogies is of paramount importance for meeting the growing energy require-
ments for economic expansion and to improve human life in general. The
development and holistic propagation of renewable energy technologies
has been and still is being prioritized in the global village countries so as
to provide sustainable energy source options to meet demands. Significant
advances have been made, since the mid-1970s, in a number of tech-
nologies such as wind energy, solar systems, geothermal applications,
and the production of biogas, bio-ethanol, and biodiesel from biomass
feedstocks [4].
The full deployment of the biodiesel technology is being hampered
mainly by the high cost of production compared to the conventional die-
sel, which is petroleum-derived and is, therefore, dependent on the cur-
rent price of petroleum and the politics surrounding this. Research work is
currently progressing at appreciable pace to find means of overcoming the
economic barriers associated with the development and implementation of
advanced biodiesel-based energy technologies. Support from governments
is being provided in the form of subsidies as witnessed in Europe, some
Asian states, Brazil, the United States of America, and South Africa. All of
these countries are currently making significant funds available to meet
their policy targets as far as biodiesel technology deployment is concerned.
Hydrogenolysis as a Means of Valorization 139

O COR OH O
cat.
O COR + MeOH OH +
R OMe
O COR OH

vegetable oil glycerin fatty acid methy esters


(FAME)

Figure 4.1 Biodiesel production through transesterification process [4].

As one of the ways to reduce the cost of the overall biodiesel produc-
tion, finding ways to add value to the glycerol that is produced as the main
by-product in the biodiesel manufacturing process (Figure 4.1) should be
considered. Finding uses for glycerol as such or after converting it into
high-value-added products has been identified as one way to significantly
lower the cost of biodiesel production [5].
Intensive research efforts are currently expended to find economically
and environmentally viable ways to transform the glycerol into value-added
chemical products. Of these methods, hydrogenolysis is one of the utmost
promising catalytic routes of conversion of glycerol into value-added comod-
ities [6]. Unfortunately, most development work reported thus far displays
significant shortcomings, particularly with respect to the selectivity toward
the desired product, 1,2-propanediol. From the results discussed in the fol-
lowing sections, it may be concluded that there is a clear need to develop
processes characterized by an improved selectivity toward the desired prod-
uct, and higher conversion of the glycerol fed to the process, while maintain-
ing mild and environmentally benign reaction conditions.
A way this may be achieved is by synthesizing novel catalyst systems,
selective to the production of 1,2-propanediol (1,2-PDO) and thereafter
identifying the optimal reaction conditions for maximizing the reaction
yield. This review deliberates on recent investigative work relating to the
optimization of the process of producing 1,2-PDO, beginning with the
catalyst selection and the choice of the most suitable reaction parameters
for this reaction among other reaction pathways used for the valorization
of glycerol. Recent findings and conclusions from various researchers are
listed and critically reviewed.

4.2 Ways of Valorization of Biodiesel-Derived


Glycerol
The increasing demand for biodiesel is expected to lead to potentially
abundant quantities of crude glycerol on the market. As mentioned above,
140 Biodiesel Technology and Applications

glycerol valorization has much to offer in the cost reduction of biodiesel


production within the framework of an integrated biorefinery. To this
end, various chemical or biotechnological strategies have been developed
to yield value-added chemical products using crude glycerol as substrate
[7]. The vast majority of unprocessed glycerol is used as raw mate-
rial for the production of other valued chemicals, followed by its use in
animal feeds.
Orthodox catalysis and biotransformation are the two foremost routes
available for transforming raw glycerol to an array of valued chemicals [8].
In recent times, encouraging results have been obtained in processes for
crude glycerol transformation. As an example, the production of 1,3-pro-
panediol, docosahexaenoic acid-rich algae, poly (hydroxy alkanoates),
butanol, hydrogen, mono glycerides, lipids, citric acid, and syngas from
crude glycerol are promising [9].
However, several of the available technologies require further develop-
ment to optimize them toward being cost-effective and operationally feasi-
ble for incorporation into functional biorefineries.

4.2.1 Catalytic Conversion of Glycerol Into Value-Added


Commodities
The summary of work done so far on the catalytic conversion of glycerol, as
one of the main routes to transforming glycerol into great value commod-
ity chemicals, is shown in Figure 4.2. The work substantiated the claim that
the catalytic conversion of this by-product of biodiesel manufacturing into
a variety of products as a feasible and commercially attractive route.

4.2.1.1 Catalytic Oxidation of Glycerol


Over the years, various researchers found that precious metal catalysts that
include gold, silver, palladium, and platinum showed remarkable catalytic
activity in the partial oxidation of glycerol in the liquid phase [10–14]. In
the various studies, the effect of reaction parameters such as varying tem-
perature, the oxygen pressure, the glycerol concentration, and the synthe-
sis method used to prepare the catalysts was investigated.
Ketchie et al. [12] studied the effect of gold (Au) particle size ranging
from 5 to 42 nm supported on carbon in the oxidation of glycerol in the
aqueous phase. Oxidation of glycerol in the liquid phase was performed
at 59.85°C and a pH of 13.8. The results obtained showed that large par-
ticles (≥20 nm) were more selective to glyceric acid (83%) with the small
Hydrogenolysis as a Means of Valorization 141

O OH HO OH
OH
HO O O O
HO OH
OH OH O
dihydroxyacetone glyceraldegyde tartronic acid hydroxyethanoic acid
oxidation HO O
catalyst OH O O
O HO OH
HO OH HO HO OH formic acid
O OH O O O O
glyceric acid hydroxypyruvic acid mesoxalic acid oxalic acid
OH HO OH
hydrogenolysis HO HO OH ROH

catalyst
1,2-propanediol 1,3-propanediol ethylene glycol R=C1-C3

O
dehydration O
OH
catalyst
acetol acrolein

pyrolysis, CnH2n+2 (C+H2)


CnH2n ROH (CO+H2)
gasification
alkane olefin alcohol syngas carbon + hydrogen
catalyst
OH
HO OH OH O O
O R OH
transesterification,HO HO O O O
esterification OH
O O
catalyst monoglycides α-monobenzoyl glycerol glycerol dimethacrylate

OH
O
etherification O O O
OH
catalyst OH O O
OH O O O
OH OH
monoethers diethers triether

oligomerization, OH
polymerization O OH polyglycerol methacrylates
Ci
catalyst j
glycerol 1-monoethers
O
carboxylation O
catalyst O OH
glycerol carbonate

Figure 4.2 Catalytic processes for the conversion of glycerol into useful value-added
chemicals [5].

particles having a lower selectivity, 69%, at 50% conversion of glycerol. This


was attributed to the higher formation rate of hydrogen peroxide (H2O2)
during the oxidation process. The presence of the peroxide promotes
the cleavage of the C-C bond and thus the formation of other products
other than glyceric acid (GLYA), namely, glycolic (GLYCA), lactic, oxalic
(OXALA), and tartronic (TARAC) acids.
142 Biodiesel Technology and Applications

The effect of the catalysts’ preparation method, for instance, sol immo-
bilization versus deposition precipitation and the activation method used,
i.e., chemical reduction versus calcination on Au/TiO2 catalysts for liquid
phase oxidation of glycerol, was studied by Dimitratos et al. [11]. Their
study showed that catalytic performance was dependent on the method of
preparation used for the different catalysts. Catalysts prepared and treated
by different methods were tested for activity at 50°C and 0.304-MPa oxy-
gen pressure. The results in Table 4.1 show the dependence of the selectiv-
ity obtained on the method of preparation. The conclusion may be drawn
that with this type of catalysts, the deposition precipitation method fol-
lowed by calcination gives the best results as regards the selectivity of the
formation of glyceric acid.
Glycerol oxidation over palladium (Pd), gold (Au), and platinum (Pt)
nanoparticles supported on carbon was probed by Carrettin et al. [10]. A
batch reactor was used at 60°C and air at 0.1 MPa pressure as oxidant.
At stated conditions, Pd/C and Pt/C gave some selectivity to glyceric acid
even though the main products were the undesirable by-products, CO2,
HCHO, and HCOOH. The C–C cleavage reactions were propelled by the
strong acidic cites given by the metals and the formation of C1 compounds
was eliminated by the adding basic sodium hydroxide (NaOH) to the
reaction vessel. The use of pure oxygen at a raised pressure of 0.3 MPa
significantly improved the catalytic activity of Pd/C and Pt/C catalysts.
Supported Au catalysts were completely inactive at the initial operating
conditions though for 1 wt.% Au/C, 100% selectivity was readily achieved
when the basic promoter aid was used at 0.3 MPa pure oxygen pressure.

Table 4.1 Glycerol oxidation using different catalysts preparation methods at


50°C [11].
S50–S90
Entry Samples GLYA GLYCA TARAC d (nm) HRTEM
1 1% Au/TiO2 (DP/calcined) 81 13 4 5
2 1% Au/TiO2 (DP/NaBH4) 55–62 28–26 17–12 2–5
3 1% Au/TiO2 (THPC) 55–52 15–15 25–28 2–3
4 1% Au/TiO2 (PVA) 52–69 13–17 24–10 4–5
5 1.5% Au/TiO2 (WGC) 64–60 23–23 11–12 4
6a 1% Au/ACb (PVA) 64–52 12–12 9–10 5
DP, deposition precipitation; THPC, Tetrakis hydroxypropyl phosphonium chloride; PVA,
poly-vinyl alcohol; WGC, World Gold Council; HRTEM, high-resolution transmission
electron microscopy; S50 and S90, selectivity at 50% and 90% glycerol conversion, respectively.
a
: From Ref. [15]; b: activated carbon.
Hydrogenolysis as a Means of Valorization 143

The base added is assumed to facilitate the initial dehydrogenation via


H-abstraction of one of the primary OH groups of the glycerol. In this
way, the rate limiting step (initial dehydrogenation) of the process will be
overcome. Similar results have been obtained by several researchers for
gold and palladium supported catalysts used in the catalytic oxidation of
glycerol [13, 15–20].
Furthermore, Skrzyńska et al. [14] investigated the activity of Pd, Pt,
Au, and Ag supported on alumina with a loading of 0.95 wt.% Pt, 0.96
wt.% f Pd, 0.98 wt.% Au, and 1.13 wt.% Ag in the oxidation of crude glyc-
erol. Based on the initial rate of reaction, Au/Al2O3 was the most active
catalyst sample, though a comparison of selectivity and conversion after
1–2 h showed that Pd/Al2O3 was the most robust and unaffected catalyst
toward impurities in crude glycerol, whereas the gold-containing sam-
ples rapidly deactivated. A loss of less than 50% (relative to the compara-
ble reaction which used pure glycerol) and an almost constant selectivity
to glyceric acid of close to 80%–90% was observed with the palladium
catalysts. The other catalysts’ activities were greatly affected, with the
conversion being lowered by a factor of 4 and 10 for Pt and Au systems,
respectively, when using crude glycerol instead of pure glycerol as raw
material. In the case of the effect of specific impurities on conversion
and selectivity, the group found that matter organic non-glycerol and
non-methanol (MONG-NM) impurities as well as sulfur compounds
were the most detrimental to catalyst performance. Therefore, in agree-
ment to the findings by [21], these should be removed prior to the cata-
lytic reaction.

4.2.1.2 Catalytic Dehydration of Glycerol


Numerous solid and acidic catalysts that include phosphates, sulfates, zeo-
lites, and supported heteropoly acids have been tested in the dehydration
of glycerol in the gaseous phase to produce acrolein [22–24]. The highest
selectivities to acrolein obtained by these research groups range between
65% and 80%. Suprun et al. [25] reported that solid catalysts with Hammet
acidity (Ho) of between −10 and −16 were the ones more suitable for use
in the dehydration of glycerol to acrolein compared to those catalysts with
Ho between −2 and −6. Also, in their studies, Suprun et al. [25] found
out that mesoporous Al2O3–PO4 and TiO2–PO4 catalysts which have large
pores showed high activity but only restricted selectivity to acrolein, while
catalysts based on the zeolites SAPO-11 and SAPO-34 with considerably
smaller micropores (5–6 Å) were less active but more selective. On the
aspect of deactivation of the catalysts, they found out that the more acidic
144 Biodiesel Technology and Applications

SAPO catalysts demonstrated the greatest cocking effect, compared to the


supported phosphates, probably due to the smaller pore diameters of the
zeolites; deactivation of the SAPOs was, therefore, quicker. Due to the fact
that crude glycerol contains a significant amount of water, there is broad
agreement among the above-mentioned researchers that there is need for
a catalyst with high water tolerance because it will improve the economics
of raw glycerol conversion to acrolein by eliminating the drying stages in
the glycerol purification.

4.2.1.3 Pyrolysis of Bioglycerol


The pyrolysis of glycerol under different operating conditions, varying
temperature (650°C–800°C), varying carrier gas flow rate (30–70 ml/
min), different packing material (quartz, silicon carbide and sand), and
different particle size of the material was investigated by Valliyappan et
al. [26]. While varying flow rate at 800°C, they reported that increasing
the carrier gas flow reduced the yield of H2 owing to the reduced resi-
dence time in the reactor. Increasing the flow rate to 70 ml/min resulted
in a drastic increase of the yield of CO and hydrocarbons, suggesting that
these are the primary products of pyrolysis. The effect of temperature
on the pyrolysis of glycerol was explored at 50 ml/min of carrier gas and
these researchers reported an increase in CO/H2 production, amounting
to a rise from 71% to 93.5% with an increase in the temperature from
650°C to 800°C. This was accompanied by char formation. At the same
time, a decrease in liquid products production was observed. The results
listed above are similar to those reported by Fernández et al. [27] who
also investigated the pyrolysis of glycerol to produce syngas (CO/H2).
These effects were attributed to thermal cracking of glycerol at high
temperatures.
Analyzing the effect of packing material and particle size variation at
800°C and 50 L/min H2 flow, Valliyappan et al. [26] reported that decreas-
ing the particle size of material resulted in the increase of gaseous products
and char simultaneously with decrease in yield of liquid phase products.
Smaller particle size packing materials simulated plug flow behavior in the
reactor coupled with enhanced heat transfer and thereby increasing the
conversion of liquid to gas and char. Of the catalysts tested, silicon carbide
of particle diameter of 0.15 mm was reported to have the greatest gas prod-
uct percentage of 77.6% while quartz and sand recorded maximum gas
product percentages of 71.2% and 71.9% respectively.
Buffoni et al. [28] investigated the use of nickel catalysts in the steam
reforming of glycerol to produce hydrogen. Nickel supported on α-Al2O3
Hydrogenolysis as a Means of Valorization 145

and ZrO2 or CeO2 modified catalysts were used in the investigation and the
results obtained by the researchers indicated that the minimum tempera-
ture to obtain hydrogen with high selectivity is 550°C. They also reported
that the stability and activity of the supported catalysts was enhanced by the
inclusion of CeO2 in the structure of the catalysts. Lower coke formation
and higher stability was reported and was attributed to the highest basic
character of Ni-Ceα supported catalyst that inhibited the lateral dehydra-
tion, rearrangement, and condensation reactions that could have led to the
formation of intermediate compounds in coke formation.
The results correspond to the findings of Iriondo et al. [84], who also
investigated the effect of incorporating ceria in the preparation of Ni-Al2O3
catalysts for use in glycerol steam reforming. They found that low ceria
loadings enhance the catalyst activity and improved the selectivity of H2 as
the main product. The stability of the catalysts was reported to have been
caused by nickel-ceria interactions in the surface of the alumina support.
Furthermore, an increase in the ceria load on the catalyst was reported to
be a cause of reduced capacity of the catalyst to convert intermediate oxy-
genated hydrocarbons (OHCs) into H2.
Iriondo et al. [29] investigated the effect of incorporating promoters in
the making of Ni catalysts supported on alumina for use in steam reform-
ing of glycerol and found out that lanthana (La2O3) modified Ni catalysts
produced only gaseous products. This phenomenon was reported to have
been caused by the reduction of the acidity of the alumina support thereby
favoring the conversion of OHCs into CH4, CO, CO2, and H2. There was
the formation of more OHCs when Pt catalyst was used instead of the Ni,
a result similar to the outcome of the work of Pompeo et al. [30] on the
use of platinum catalysts in steam reforming of glycerol to produce hydro-
gen. The use of bimetallic catalysts of nickel and platinum was reported
to promote methanation of glycerol rather than forming the desired and
expected steam reforming products.

4.2.1.4 Glycerol Transesterification


The main high commodity chemical yield from the transesterification of
glycerol is glycerol carbonate (GCL) which is used as a precursor and inter-
mediate in many chemical reactions such as polymerization and in the
coatings industry. Literature has documented research work that has been
performed on the transesterification of glycerol using basic catalysts. Liu et al.
[31] investigated the use of Mg/Al mixed oxide catalysts in the transesterifi-
cation of glycerol using dimethyl carbonate. They reported that while vary-
ing the Mg:Al ratio between 2 and 6, it was found that a ratio of 2:1 gave
146 Biodiesel Technology and Applications

the highest activity with 56% conversion of glycerol. This reduced to 42%
as the ratio increased, and 100% GCL selectivity when experiments were
done at 100°C and retention time of 90 min. This phenomenon was also
reported by Alvarez et al. [32] who observed the same trend and attributed
it to the increased surface basicity which depends on the metal ratio in the
catalyst. The effect of calcination temperature was also investigated between
the range 300°C to 700°C, and it was found that increasing the calcination
temperature increases the activity of all the catalysts, and notably, the high-
est activity of the Mg/Al = 2 was obtained when this system was calcined
at 600°C. The basicity is said to improve with calcination temperature due
to the formation of more weak −OH basic sites and O2− strong cites on the
catalysts. The less Al3+ Lewis acidic cites are available, the more the overall
catalyst basicity and thus the activity of the catalysts improves [31].
This result is consistent with the findings of Kumar et al. [33], Zheng et
al. [34], and Zheng et al. [35] who also investigated the use of mixed metal
oxides in the transesterification of glycerol. In comparison with other basic
catalytic materials used for this reaction, such as NaOH, CaO, and MgO
with glycerol conversion of 70%, 79%, and 50%, respectively, they demon-
strated that mixed metal oxides of Mg/Al = 2 is very much competitive
with a conversion of 66% and GLC selectivity of 100%. CaO and NaOH
fall short on the GLC selectivity with values of 81% and 70%, respectively,
when reactions were run for 2 h [35].
Apart from the use of metal oxides as basic catalysts in the transesteri-
fication of glycerol, Lanjekar and Rathod [36] reported the use of enzymes
to catalyze the reaction. They reported a high conversion of glycerol of
94.85% after running the reactions for 14 h at a temperature of 60°C.
Commercially, this route will turn to be expensive due to the long retention
times, expensive enzymes and the easiness to poisoning of the enzymes.
Algoufi and Hameed [37] reported the use of K-zeolite derived from coal
fly ash in the transesterification of glycerol. The reported data showed that
4 wt.% catalyst loading, reaction temperature of 75°C, and retention time
of 90 min yielded 100% conversion of glycerol and 96% GC selectivity.

4.2.1.5 Glycerol Direct Carboxylation


The high-value end product of catalytic glycerol carboxylation is glyc-
erol carbonate (glycerine carbonate or 4-hydroxymethyl-2-oxo-1,3-di-
oxolane). The same product is also obtained by the transesterification of
glycerol using organic carbonates [38]. Glycerol carbonate is a versatile
building block for a variety of advanced materials. Little is known to
have been reported on the direct carboxylation of glycerol using carbon
Hydrogenolysis as a Means of Valorization 147

dioxide except the works of Aresta et al. [39] and Ezhova et al. [40] that
was reviewed by Sonnati et al. [38]. The work by these research groups is
consistent in that they all used high pressure CO2 in liquid phase (above
5.07 MPa) and high temperatures (over 180°C). The stringent reaction
conditions require specialized equipment and thus escalate the ultimate
production cost of glycerol carbonate. Notable is the research by Ezhova
et al. [40] who investigated the use of gaseous CO2 at lower tempera-
tures in the presence of rhodium complexes with phosphine ligands.
The group reported a conversion of only 0.24% of the glycerol, achieving
selectivity toward glycerol carbonate of 100% while operating at 5.07
MPa (4.05 MPa CO2 and 1.02 MPa H2), 140°C, and 17-h reaction time.
This result falls far lower than the maximum conversion reported by
Aresta et al. [39], who reported a conversion of 6.86% while operating
at 5 MPa, 180°C, and 15-h retention time in the presence of a tin oxide–
based heterogeneous catalyst. Reasons for the anomalies and discrepan-
cies in the results might be attributed to differences in catalytic activity
between the two systems and different operating conditions used by the
two research groups.
Having highlighted some advancements being made in other catalytic
routes for the conversion of glycerol to high value chemicals, the following
section focuses on the conversion of glycerol to 1,2-PDO via hydrogenoly-
sis using different reaction systems.

4.3 Hydrogenolysis of Glycerol


4.3.1 Definition of Hydrogenolysis
Hydrogenolysis is defined as the chemical reaction involving the cleav-
age of a C-C or C-X (X=heteroatom) bond of a hydrocarbon compound,
followed by the addition of hydrogen, forming two molecules as a result.
This process has widely been used in the crude oil refining industry for
a variety of processes. One example is the catalytic hydrodesulfurization
of sulfur-rich hydrocarbons obtained from petroleum feedstocks. The
hydrogenolysis of carbohydrate feedstocks has been extensively studied
and its history dates back to the 1930s [41].
In recent times, the process of hydrogenolysis has witnessed renewed
interest for the transformation and conversion of glycerol and carbohy-
drates feedstocks into value-added polyols that include propylene gly-
col, ethylene glycol, and propane diols, among other derived products
[42–48].
148 Biodiesel Technology and Applications

OH OH
OH 1-propanol C3H8
1,2-propanediol Propane
(1,2-PD)
H2 OH
HO OH 2-propanol
HO OH + + C2H6
OH Catalyst 1,3-propanediol Ethane
Glycerol (1,3-PD)
OH
Ethanol
CH4
HO OH Methane
Ethylene glycol CH3OH
(EG) Methanol
Dihydric alcohol Monobasic alcohol Alkane

Figure 4.3 Successive stages in the hydrogenolysis of glycerol [49].

4.3.2 Catalytic Hydrogenolysis of Glycerol


Catalytic hydrogenolysis of glycerol is a multistage chemical reaction, usu-
ally activated by supported heterogeneous transition metals together with
base metals as evident in the above cited publications. Figure 4.3 shows
the successive hydrogenolysis of glycerol (as a starting raw material) to a
variety of products at different stages of the reaction.

4.3.3 Product Spectrum from Hydrogenolysis of Glycerol


Several chemicals can be derived from glycerol via chemoselective hydrog-
enolysis of glycerol, including 1,2-PDO, 1,3-PDO, lactic acid, with a fur-
ther hydrogenolysis of the main products leading to the formation of
propanol and subsequently propane as secondary products [6, 50–53]. In
some instances, CO2, formic acid, and even methane were detected in the
product streams, owing to reaction conditions used that end up affecting
the truncation of reaction systems [44].
The main point raised from the observations of Guo et al. [44], was the
significant amounts of lower alcohols and hydrocarbons that are detected
in the product streams with elevated reaction temperature. Other aspects
that influence the product range are the catalysts used and the relative
acidity of the active sites, the reaction time and the pressure of the hydro-
gen in cases where hydrogen was introduced to the reaction mixture. In
some instances where hydrogen donors (methanol, 2 propanol, and formic
acid) were used as sources of the reactive hydrogen to the reaction mix-
ture, the effect of the different molecules, pumping rate, and concentration
Hydrogenolysis as a Means of Valorization 149

have been studied to also help explain the phenomenon of hydrogenolysis


product range [43].
1,2-Propanediol, also known as propylene glycol, is a major commodity
chemical and has seen a 4% year on year market growth [42]. Propylene
glycol has several commercial applications, for example, as antifreeze,
coolant, solvent and extractant, de-icing agent, precursor in pharmaceu-
ticals, cosmetics, animal feed, and tobacco industries. It also plays a role
in the petroleum production, sugar refining, paper making, toiletries, liq-
uid detergent, alkyl resins, printing inks, plasticizers, and hydraulic break
fluids [6].
The vast array of uses and market growth for 1,2-PDO has thus influ-
enced the focus of this review to be centred on the production of propylene
glycol via the hydrogenolysis of glycerol.

4.3.4 Hydrogenolysis of Glycerol to 1,2-PDO (Propylene


Glycol): Reaction Systems Overview
Key claims have been made with some of the findings being patented
regarding the process systems that can be used in the conversion of glyc-
erol to propylene glycol. Below is an overview of both heterogeneous and
homogeneous catalyst systems that have been reported to aid in the che-
moselective conversion of glycerol to 1,2-PDO.
Dalai et al. [54] described a process for hydrogenolysis of glycerol to
produce propylene glycol as the major product, comprising contacting the
glycerol with hydrogen in the presence of a heterogeneous catalyst under
the following reaction conditions: 240°C, 4 MPa H2, 80% glycerol solution,
3% w/w catalyst on support, and 1,000 rpm mixing in a batch process. The
heterogeneous catalytic systems used by these researchers consisted of Cu,
Cr, Zn, and Zr obtained by co-precipitation and calcination. High selectiv-
ities for propylene glycol formation in the range of 90% were achieved. In
comparison, Stankowiak and Franke [55] reported a 97.5% selectivity of
1,2-PDO from the hydrogenolysis of glycerol, while using a heterogeneous
catalytic system based on metal oxides (20%–60% CuO, 30%–70% ZnO,
and 1%–10% MnO). In their process, 95% purity glycerol was reacted with
2.0 to 10.0 MPa H2 pressure at 180°C.
Using Rh(CO)2 acetylacetonate and tungstenic acid (H2WO4) as a homo-
geneous catalyst system for the hydrogenolysis of glycerol, Che [56] reported
a yield of 30% 1,2-PDO and 20% 1,3-PDO. This was achieved while oper-
ating at 200°C, 30 MPa of 1:2 (CO:H2) synthesis gas and 28.6% solution of
glycerol in 1-methyl-2-pyrolidinone solvent as optimal reaction condition.
150
Table 4.2 Heterogeneously catalyzed reactions without the use of solvents.
Reaction
temperature Glycerol 1,2-PDO
Catalyst system (°C) H2 pressure MPa conversion% selectivity% Source
Raney Cu 240 3 100 86 [57]
Cu and Zn 240– 270 15 99.4 84.4 [58]
Co, Cu, Mn, Mo, and inorganic 250 25 100 95.8 [59]
polyacid
Cu, Pd, Rh on ZnO and C 183 8 19 100 [60]
Cu Chromite 200 1.38 54.8 85 [42]
Ru/C and Cation exchange resin 120 4 15 53.4 [61, 62]
(Ambient 15) 8 12.9 55.4
Biodiesel Technology and Applications

Raney Ni and Phosphonium Salt 150 1 12 93 [63]


Cu/Al nanocatalysts 220 7 38 91 [64]
5% Ru/Al2O3 and 5%Pt/Al2O3 220 In Situ Generated 50 47.2 [65]
Raney Cu 205 Trickle-bed Reactor 100 94 [66]
(H2/Liquid: 375/0.05)
Ru/Pt 200 In Situ Generated 20.6 53.1 [67]
Hydrogenolysis as a Means of Valorization 151

Similar results were also obtained in systems employing ruthenium, nickel,


palladium, or cobalt as the metal component instead of rhodium [57].
Adding titanium or molybdenum metal promoters with a group 8 metal in
the catalyst systems still did not significantly alter the composition of the
products [56]. Similarly, Drent and Jager [68] reported the hydrogenolysis
of glycerol in the presence of a homogeneous catalyst based on a platinum
group metal (PGM) or a compound of a PGM and methanesulfonic acid.
They achieved a selectivity of 21.8% toward 1,2-PDO from hydrogenolysis
of glycerol at 170°C and 2.0/4.0 (CO/H2) MPa. They suggested that further
optimization could possibly yield better results.
In addition to the findings highlighted above, a large number of other
publications demonstrated the supremacy of the heterogeneously cat-
alyzed hydrogenolysis of glycerol to produce propylene glycol. Table 4.2
summarizes results obtained by different researchers using heterogeneous
catalysts.
The major hindrances that severely dent the use of homogeneous cat-
alysts in this process are the use of toxic solvents (methanol and sulfur
containing acids) in the reaction systems and the problems related to the
catalysts recovery after the reaction completion. Coupled with the low
conversions and selectivity toward 1,2-PDO, the use as well as research
pertaining the use of homogeneous catalyst systems is limited to a few
publications.
Despite the publication of several investigation results by a number of
researchers, the heterogeneously catalyzed glycerol hydrogenolysis to 1,2-
PDO process still has several drawbacks, which prevent scaling up beyond
bench scale to the pilot plant level. Aspects of great concern include the
high operating H2 pressures (>1.03 MPa), high temperatures (> 300°C),
use of very dilute glycerol solution (10%–30%), poor catalyst reusability,
catalyst leaching coupled with low glycerol conversion, and/or low 1,2-
PDO selectivity. Also, another significant limitation is that of long reaction
times (>5 h in batch systems) presently required to achieve an acceptable
conversion of glycerol to 1,2-PDO, thus rendering the current methods
economically non-viable [5].
Having noted the above systems’ drawbacks in the conversion of glyc-
erol to 1,2-PDO, the following sections disclose recent research work find-
ings by different groups in quest of optimizing the process.

4.3.5 Catalyst Selection


The applicability of supported precious metal catalysts and/or nickel-
based systems in the hydrogenolysis of glycerol with relatively high
152 Biodiesel Technology and Applications

activity and selectivity toward 1,2-PDO has been reviewed critically by


various researchers [43, 69–71]. They went on to propose that further
work should be carried out using different catalysts based on ruthe-
nium or copper on various supports. Their proposals were derived from
the works of Montassier et al. [72] and Feng et al. [73] who investigated
the use of ruthenium (Ru) in the hydrogenolysis of glycerol to 1,2-PDO.
Their findings were compared with the results when other noble metal
catalysts (Pt, Pd, and Ni) were tested [72, 73]. They derived the conclu-
sion that supported Ru catalysts demonstrated the highest activity and
selectivity toward 1,2-PDO. Findings by Miyazawa et al. [53], Balaraju et
al. [74], and Lahr and Shanks [75] concurred with the outcomes of this
research work.
Furthermore, Checa et al. [76] performed work on the catalytic trans-
formation of glycerol using various metal systems supported on zinc
oxide (ZnO) to investigate the different activities toward the reaction
exhibited by these metals. The group performed their catalytic tests at
2.0 MPa hydrogen (H2) or helium (He) pressure, 180°C, and a run time
of 12 h under basic or neutral atmosphere. The results obtained led to
the conclusion that the conversion of glycerol proceeds faster in a basic
medium compared to a neutral one and in a He as compared to under
a H2 atmosphere. Of the metal systems tested, the highest conversions
were reported in the order of Pt > Rh > Pd > Au, consistent with the
findings of Auneau et al. [77] who investigated the effect of reaction
medium, with gold and palladium. They recorded zero in neutral phase
and 5% and 17% for rhenium and platinum, respectively. Under basic
conditions, the conversion levels were 100%, 97%, 32%, and 26% for Pt,
Rh, Pd, and Au, respectively. With regard to the selectivity toward 1,2-
PDO, there was minimum variation ranging between 14% for Au and Pd
whereas the values were 19% and 27% for Pt and Rh, respectively. These
figures however fall short of those reported by Vasiliadou et al. [78]
which were 40.5% glycerol conversion and 1,2-PDO selectivity of 60.5%
using Ru-based catalyst supported on ZrO2. Other results quoted that
dwell on the effectiveness of Ru in the selective catalytic transformation
of glycerol to 1,2-DPO include Huang et al. [50] who achieved 47.8%
selectivity while Dasari et al. [42] obtained a moderate 40.0% selectivity
toward 1,2-PDO.
Of late, copper (Cu)–based catalysts have since been reported to have
high selectivity toward 1,2-PDO in the selective hydrogenolysis of glycerol
owing to its ability to promote the cleavage of the C-O bonds in contrast
to C-C bonds. Bienholz et al. [69], Guo et al. [71], and Gandarias et al.
[43] concurred with respect to Cu catalyst systems activity toward glycerol
Hydrogenolysis as a Means of Valorization 153

hydrogenolysis with all of them reporting the selectivity of 1,2-PDO rang-


ing from 70% up to 98% depending on the reaction conditions. It was
observed however, that, though the selectivity of the formation of 1,2-PDO
was very high, the conversion was very low, hovering below 30% for cat-
alysts with a Cu loading of 60% CuO, despite the system being operated
at an elevated temperature of 200°C and a pressure of 5 MPa [69]. High
conversions of 100% were only realized for gaseous reactions, elongated
reaction times up to 24 h and high catalyst weight loaded into reactors as
highlighted by Akiyama et al. [79] and Guo et al. [71].
The aspects of reusability of catalyst systems and ultimate trade-offs of
running the selective hydrogenolysis of glycerol to yield 1,2-PDO using
the front runner catalysts have not yet been explored, and thus, it becomes
inconclusive to declare the prime system that can solely be applied for the
hydrogenolysis of glycerol process. This uncertain and inconclusive sce-
nario leaves a knowledge gap for further investigations to be performed
around the hydrogenolysis of glycerol to 1,2-PDO and to optimize the pro-
cess thereof.

4.3.6 Reaction Conditions That Influence the Hydrogenolysis


of Glycerol to 1,2-PDO
Reaction conditions are the main drivers of all the transformations from
raw materials to desired products. In the case of the selective hydrogenoly-
sis of glycerol to 1,2-PDO, the combination of catalyst systems and physical
parameters (temperature, H2 pressure, catalyst loading, and reaction time)
influence the reaction pathway and thus the product mix obtained. This
section reviews some recent published investigations done in the quest
to ascertain optimal operating conditions for the production of 1,2-PDO
from the heterogeneously catalyzed hydrogenolysis of glycerol.

4.3.6.1 Effect of Reaction Temperature


Akiyama et al. [79] investigated the effect of reaction parameters using
Cu-based catalysts and found out that elevated temperatures (from 170°C
and above) lead to complete glycerol conversion but also subsequently lead
to decomposition of glycerol into lower alcohols (2-propanol, 1-propanol,
ethanol, methanol, and ethylene glycol) as the temperature is raised above
190°C as evident by the drop in selectivity of 1,2-PDO. This pattern is
clearly shown in Figure 4.4. Shinmi et al. [80] and Dasari et al. [42] also
found that selectivity toward 1,2-PDO drastically drops with excessive
reaction temperature above 200°C.
154 Biodiesel Technology and Applications

100 100
Glycerol Conversion (%) 95 95

Selectivity 1,2-PDO (%)


90 90
85 85
80 80
75 75
70 70
65 65
60 60
55 55
50 50
150 170 190 210 230 250
Reaction Temp (ºC)

Glycerol Conversion Selectivity 1,2-PDO

Figure 4.4 Effect of temperature on glycerol conversion and selectivity of 1,2-PDO [79].

Dasari et al. [42] found that there was a great influence of temperature
on the overall yield of propylene glycol (1,2-PDO) from the experimen-
tal reactions carried out while varying temperature from 150°C to 260°C
and at an operating hydrogen pressure of 1.38 MPa in the presence of a
copper chromite catalyst. The results obtained by this group showed that
as the temperature was increased from 150°C to 260°C, there was a uni-
form rise in the glycerol conversion from a low 7.2% to 87%. In the case
of yield and selectivity of 1,2-PDO, similar trends were observed where an
increase in the reaction temperature resulted in an increase in yield and
selectivity only up to 200°C. Further rise in temperature to 260°C resulted
in a decrease in the yield and the selectivity of 1,2-PDO as show in Figure
4.5 below. The further hydrogenolysis of the products obtained to lower
alcohols with excessive temperatures (above 200°C) was verified by the
detection of lower alcohols in the product stream and the drop in yield and
selectivity.

4.3.6.2 Effect of H2 Pressure


Regarding the effect of pressure on the process of hydrogenolysis of glyc-
erol to 1,2-PDO, Dasari et al. [42] reported that as the pressure was raised
from 0.34 to 2.07 MPa, the conversion of glycerol increased from 25%
to 65.3%, respectively, while the selectivity to 1,2-PDO improved from
36.4% to a maximum of 89.6%. In contrast, Shinmi et al. [80] reported an
increase in glycerol conversion from 25% to 79% with a pressure rise from
Hydrogenolysis as a Means of Valorization 155

100
90
80
70
60
50
%

40
30
20
10
0
100 150 200 250 300
Reaction Temp (ºC)

% Conversion % Yield % Selectivity

Figure 4.5 Effect of varying temperature on conversion, yield, and selectivity of 1,2-PDO.
Reactions were performed using 80% glycerol solution at hydrogen pressure of 1.38 MPa
for 24 h [42].

2 to 8 MPa but found no significant change (between 40% and 43.5%) in


the selectivity with the variation of operating pressure. The discrepancy
could be explained due to the different nature of the catalytic systems used,
which exhibit different characteristics since Dasari et al. [42] used co-­
precipitated copper chromite catalysts while the findings of Shinmi et al.
[80] were based on rhodium supported systems. The claim is supported
by Bolado et al. [70] who noted that different catalyst systems with vary-
ing physical and chemical properties also perform differently in these
chemical reactions, even though the other parameters are held constant
[57, 73, 72].

4.3.6.3 Effect of Initial Water Concentration


The effect of initial water content in the reaction mixture on the conver-
sion of glycerol and also on the selectivity to 1,2-PDO was investigated
by Ma and He [81], Dasari et al. [42], and Guo et al. [71]. Ma and He
[81] reported that there was almost no change in the glycerol conversion
but an increase in 1,2-PDO selectivity from 22.2% to 34.3%. According to
these authors, this resulted in changing the glycerol concentration from
10% to 40%, while using the Ru/Al2O3 catalyst system. Contrary to that,
when reactions were performed over Ru/Al2O3 + Re2(CO)10 conversion
of glycerol increased from 36.4% to 53.4% together with the 1,2-PDO
156 Biodiesel Technology and Applications

selectivity rising from 36.5% to 50,1%. Similar trends with regard to ini-
tial water content were also observed previously by Dasari et al. [42] who
achieved 85.0% as highest selectivity to 1,2-PDO at 20% initial water con-
tent. Dasari et al. [42] also witnessed a decline in selectivity, as they reduce
the water content to zero, to 71.9% when tests were performed using pure
glycerol.
In contrast to the decline reported by Dasari et al. [42], Guo et al.
[71] observed a maximum of 93.3% selectivity to 1,2-PDO correspond-
ing to 41.9% glycerol conversion while using pure glycerol. Dasari et al.
[42] further went on to suggest that it was advisable to exclude water
from the feedstock in an effort to drive the equilibrium toward the
products of hydrogenolysis since the reaction itself produces water as
a by-product.

4.3.6.4 Effect of Reaction Time


The increase in reaction time has a positive effect on the conversion levels
of glycerol but causes insignificant change to the selectivity toward 1,2-
PDO [71]. The same can be said for the findings of Nakagawa et al. [82]
where conversion of glycerol approached 90% after 50 h of residence time
though selectivity showed a rather decreasing trend from about 10% to
4.5% with increasing reaction time. It is thus wiser to limit the retention
time in the reactor so as to shorten the turnover time of the process for
there seems to be no considerable gain in 1,2-PDO selectivity as the main
product of choice in this instance.

4.3.6.5 Effect of Catalyst Weight


The work of Dasari et al. [42] was extended to investigating the effect of
varying the catalysts weight where they found that conversion of glyc-
erol increased with the rise in catalyst weight loaded into the reactor.
Initially, as the catalyst weight percentage is increased from 1% to 5%,
the conversion increased from 28.3% to 54.8%, whereas selectivity rose
from 63.3% to 85.0%. Further, catalyst weight percentage increase to 20%
levels resulted in a decline of the selectivity to 62% though the conversion
continued to rise to a peak 78.5%. Dasari et al. [42] attributed the fall
in selectivity with rising catalyst percentage to the availability of more
hydrogenolysis sites on the surplus catalysts that promoted excessive
hydrogenolysis of the formed 1,2-PDO to lower alcohols and gaseous
products.
Hydrogenolysis as a Means of Valorization 157

4.3.6.6 Proposed Reaction Mechanisms for Glycerol Hydrogenolysis


to Produce 1,2-PDO
Glycerol hydrogenolysis is thought to proceed via a mechanism that
involves the dehydration of glycerol to acetol and 3-hydroxy-propanal pro-
moted by acid catalysis in the first stage, subsequently followed by hydro-
genation to the glycols by the metal catalysts used, as noted by Vasiliadou
et al. (2009). The scheme in Figure 4.6 shows the proposed reaction mech-
anism revealing products of over-hydrogenolysis, hydrolysis of propane
diols to 1-propanol, 2-propanol, ethanol, and methanol.
The scheme proposed by Vasiliadou et al. [78] is in sync to mech-
anisms also proposed by Furikado et al. [83] who found that there is
sequential hydrogenolysis of the propanediols formed to give either 1-
propanol and/or 2-propanol using supported Ru catalysts on carbon. To
further substantiate the reaction mechanism which suggests the forma-
tion of acetol as an intermediate product in the hydrogenolysis of glyc-
erol, Miyazawa et al. [62] investigated the hydration of acetol and reported
that there was 100% conversion of acetol at 8 MPa H2 pressure, 120°C
temperature and a reaction time of 10 h. Product composition is similar
to that mentioned above in the proposed reaction mechanism scheme of
Vasiliadou et al. [78] (Figure 4.6).
The work performed by Dasari et al. [42], Montassier et al. [72], and Lahr
and Shanks [75], together with those reported above, concurred with the
detection of trace amounts of acetol in the final product analysis, implying
that there was a stage at which it was produced in the system. To further sub-
stantiate their claim that acetol is an intermediate product in the proposed

+H2
CH3 C CH2 CH3 CH CH2 1-propanol
O OH OH OH
acetol 1.2-propanediol +H2
-H2O 2-propanol
-H2O

-H2O O +H2
CH2 CH CH2
C CH2 CH2 CH2 CH2 CH2 +H2
OH OH OH ethanol
H OH OH OH
methanol
3-hydroxypropionaldehyde 1.3-propanediol
+H2
ethylene glycol
ethanol
methane

Figure 4.6 The scheme for glycerol hydrogenolysis reaction mechanism [78].
158 Biodiesel Technology and Applications

reaction mechanism, the authors proceeded to perform the hydrogenolysis


of glycerol in a two-step process. In the first stage, the reaction was left to
proceed without hydrogen but in the presence of the metal catalysts tested for
hydrogenolysis of glycerol. The subsequent stage would then be performed
under hydrogen pressure with the resulting product selectivity showing that,
in the first stage, acetol was produced and, in the second stage, results showed
acetol depletion simultaneously with great yield of 1.2-PDO.
In contrast to the above-mentioned mechanism, there also exists another
mechanism proposed by Montassier et al. [57] shown in the Figure 4.7.
The mechanism that includes an initial dehydrogenation of glycerol is
said to lead to the formation of glyceric aldehyde in equilibrium with its eno-
lic tautomer. A nucleophilic reaction of water or an adsorbed OH function
(dehydroxylation) followed by the hydrogenation of the unsaturated alde-
hyde was proposed to be the explanation for the end product being 1,2-PDO.
Based on their results, multiple groups suggested otherwise that a differ-
ent mechanism was followed by the system [42, 72, 75, 78]. There seems to
be no evidence that supports the route proposed by Montassier et al. [57]
for the formation of 1,2-PDO from glycerol, owing to the lack of any trace
of the postulated intermediate compounds in the product mixture.
On the other hand, the mechanism suggested by Montassier et al. [72]
was shared by Maris and Davis [52] who proposed the very same route for
chemoselective hydrogenolysis of glycerol. These groups argued that the
path and selectivity toward certain products is dependent on the catalyst

Dehydrogenation Dehydroxylation by
of C-O H2O oradsorbedOH

Hydrogenation
O OHOH
C- C- CH2
H H “H O”
2 O OH OH OH
OH OH OH 2 H2
C- C= CH2 CH2- CH - CH3
CH2- CH - CH2
H
Glycerol Glyceraldehyde Propylene Glycol
“H2O”

H- O OH OH
C= C- CH2
H

Figure 4.7 Conversion of glycerol to 1,2-PDO, reaction mechanism [57].


Hydrogenolysis as a Means of Valorization 159

type and, particularly, the surface properties of the catalyst used, thus
implying that any proposed mechanism that is fully supported by valid
results might be the actual pathway for the reaction.

4.4 Conclusion
The hydrogenolysis of glycerol to 1,2-propanediol is one of the most
promising routes for producing value-added chemicals from glycerol, a
by-­product of biodiesel manufacturing. However, there is need to develop
processes characterized by an improved selectivity toward the desired
product, and higher conversion of the glycerol fed to the process, while
maintaining mild and environmentally benign reaction conditions.
There are other competing by-products which include glyceric acid but
there is a vast array of uses and increase in market growth for 1,2-PDO.
Advancements made in catalytic routes for the conversion of glycerol to
1,2-PDO via hydrogenolysis using different reaction systems was explored.
It was evident that organic non-glycerol and non-methanol (MONG-NM)
impurities as well as sulfur compounds were the most detrimental to cat-
alyst performance and should be removed prior to the catalytic reaction.
The results obtained led to the conclusion that the conversion of glycerol
proceeds faster in a basic medium compared to a neutral one and in a He
as compared to under a H2 atmosphere. The review pointed to the effec-
tiveness of rhenium in the selective catalytic transformation of glycerol
to 1,2-DPO and also to the fact that operating temperature must be kept
below 200°C. It is also very important to exclude water from the feedstock
in an effort to drive the equilibrium toward the products of hydrogenoly-
sis since the reaction itself produces water as a by-product. The retention
time in the reactor must be reduced so as to shorten the turnover time of
the process because higher reaction time do not lead to considerable gain
in 1,2-PDO selectivity as the main product of choice. Further investiga-
tions are still needed in order to establish the reaction mechanism for the
hydrogenolysis of glycerol to 1,2-PDO and so as to optimize the process.

References
1. Pegels, A. Renewable energy in South Africa: Potentials, barriers and options
for support. Energy Policy, 38(9), 4945–4954, 2010.
2. Nogueira, L. A. H. Does biodiesel make sense? Energy, 36(6), 3659–3666,
2011.
160 Biodiesel Technology and Applications

3. Kandpal, T. C. & Broman, L. Renewable energy education: A global status


review. Renewable and Sustainable Energy Reviews, 34(0), 300–324, 2014.
4. Gerpen, J. V. Biodiesel processing and production. Fuel Processing Technology,
86(10), 1097–1107, 2005.
5. Zhou, C.-H., Beltramini, J. N., Fan, Y.-X. & Lu, G. Q. Chemoselective cata-
lytic conversion of glycerol as a biorenewable source to valuable commodity
chemicals. Chemical Society Reviews, 37(3), 527–549, 2008.
6. Balaraju, M., Jagadeeswaraiah, K., Prasad, P. S. S. & Lingaiah, N. Catalytic
hydrogenolysis of biodiesel derived glycerol to 1,2-propanediol over
Cu-MgO catalysts. Catalysis Science & Technology, 2(9), 1967–1976, 2012.
7. Tan, H. W., Abdul aziz, A. R. & Aroua, M. K. Glycerol production and its
applications as a raw material: A review. Renewable and Sustainable Energy
Reviews, 27(0), 118–127, 2013.
8. Yang, F., Hanna, M. A. & Sun, R. Value-added uses for crude glycerol--a
byproduct of biodiesel production. Biotechnology for Biofuels, 5(1), 1–10,
2012.
9. Bloom, P. Hydrogenolysis of glycerol and products produced therefrom.
Google Patents, 2012.
10. Carrettin, S., Mcmorn, P., Johnston, P., Griffin, K., Kiely, C. J. & Hutchings,
G. J. Oxidation of glycerol using supported Pt, Pd and Au catalysts. Physical
Chemistry Chemical Physics, 5(6), 1329–1336, 2003.
11. Dimitratos, N., Villa, A., Bianchi, C. L., Prati, L. & Makkee, M. Gold on
titania: Effect of preparation method in the liquid phase oxidation. Applied
Catalysis A: General, 311(0), 185–192, 2006.
12. Ketchie, W. C., Fang, Y.-L., Wong, M. S., Murayama, M. & Davis, R. J.
Influence of gold particle size on the aqueous-phase oxidation of carbon
monoxide and glycerol. Journal of Catalysis, 250(1), 94–101, 2007.
13. Mimura, N., Hiyoshi, N., Fujitani, T. & Dumeignil, F. Liquid phase oxidation
of glycerol in batch and flow-type reactors with oxygen over Au-Pd nanopar-
ticles stabilized in anion-exchange resin. RSC Advances, 4(63), 33416–33423,
2014.
14. Skrzyńska, E., Zaid, S., Girardon, J.-S., Capron, M. & Dumeignil, F. Catalytic
behaviour of four different supported noble metals in the crude glycerol oxi-
dation. Applied Catalysis A: General, 499(0), 89–100, 2015.
15. Demirel-gulen, S., Lucas, M. & Claus, P. Liquid phase oxidation of glycerol
over carbon supported gold catalysts. Catalysis Today, 102–103(0), 166–172,
2005.
16. Chan-Thaw, C., Campisi, S., Wang, D., Prati, L. & Villa, A. Selective Oxidation
of Raw Glycerol Using Supported AuPd Nanoparticles. Catalysts, 5(1), 131,
2015.
17. Gil, S., Marchena, M., Fernandez, C. M., Sanchez-silva, L., Romero, A. &
Valverde, J. L. Catalytic oxidation of crude glycerol using catalysts based
on Au supported on carbonaceous materials. Applied Catalysis A: General,
450(0), 189–203, 2013.
Hydrogenolysis as a Means of Valorization 161

18. Porta, F. & Prati, L. Selective oxidation of glycerol to sodium glycerate


with goldon- carbon catalyst: an insight into reaction selectivity. Journal of
Catalysis, 224(2), 397–403, 2004.
19. Skrzyńska, E., Ftouni, J., Mamede, A.-S., Addad, A., Trentesaux, M., Girardon,
J.-S., et al. Glycerol oxidation over gold supported catalysts – “Two faces” of
sulphur based anchoring agent. Journal of Molecular Catalysis A: Chemical,
382(0), 71–78, 2014a.
20. Villa, A., Campisi, S., Mohammed, K. M. H., Dimitratos, N., Vindigni, F.,
Manzoli, M., et al. Tailoring the selectivity of glycerol oxidation by tuning
the acidbase properties of Au catalysts. Catalysis Science & Technology, 5(2),
1126–1132, 2015.
21. Skrzyńska, E., Wondołowska-grabowska, A., Capron, M. & Dumeignil, F.
Crude glycerol as a raw material for the liquid phase oxidation reaction.
Applied Catalysis A: General, 482(0), 245–257, 2014b.
22. Kim, Y. T., Jung, K.-D. & Park, E. D. Gas-phase dehydration of glycerol over
ZSM-5 catalysts. Microporous and Mesoporous Materials, 131(1–3), 28–36,
2010.
23. Wang, F., Dubois, J.-L. & Ueda, W. Catalytic performance of vanadium pyro-
phosphate oxides (VPO) in the oxidative dehydration of glycerol. Applied
Catalysis A: General, 376(1–2), 25–32, 2010.
24. Yoda, E. & Ootawa, A. Dehydration of glycerol on H-MFI zeolite investi-
gated by FT-IR. Applied Catalysis A: General, 360(1), 66–70, 2009.
25. Suprun, W., Lutecki, M., Haber, T. & Papp, H. Acidic catalysts for the dehy-
dration of glycerol: Activity and deactivation. Journal of Molecular Catalysis
A: Chemical, 309(1–2), 71–78, 2009.
26. Valliyappan, T., Bakhshi, N. N. & Dalai, A. K. Pyrolysis of glycerol for the
production of hydrogen or syn gas. Bioresource Technology, 99(10), 4476–
4483, 2008.
27. Fernandez, Y., Arenillas, A., Diez, M. A., Pis, J. J. & Menendez, J. A. Pyrolysis
of glycerol over activated carbons for syngas production. Journal of Analytical
and Applied Pyrolysis, 84(2), 145–150, 2009.
28. Buffoni, I. N., Pompeo, F., Santori, G. F. & Nichio, N. N. Nickel catalysts
applied in steam reforming of glycerol for hydrogen production. Catalysis
Communications, 10(13), 1656–1660, 2009.
29. Iriondo, A., Barrio, V. L., Cambra, J. F., Arias, P. L., Guemez, M. B., Navarro,
R. M., et al. Influence of La2O3 modified support and Ni and Pt active phases
on glycerol steam reforming to produce hydrogen. Catalysis Communications,
10(8), 1275–1278, 2009.
30. Pompeo, F., Santori, G. & Nichio, N. N. Hydrogen and/or syngas from steam
reforming of glycerol. Study of platinum catalysts. International Journal of
Hydrogen Energy, 35(17), 8912–8920, 2010.
31. Liu, P., Derchi, M. & Hensen, E. J. M. Synthesis of glycerol carbonate by
transesterification of glycerol with dimethyl carbonate over MgAl mixed
oxide catalysts. Applied Catalysis A: General, 467, 124–131, 2013.
162 Biodiesel Technology and Applications

32. Alvarez, M. G., Segarra, A. M., Contreras, S., Sueiras, J. E., Medina, F. &
Figueras, F. Enhanced use of renewable resources: Transesterification of
glycerol catalyzed by hydrotalcite-like compounds. Chemical Engineering
Journal, 161(3), 340–345, 2010.
33. Kumar, A., Iwatani, K., Nishimura, S., Takagaki, A. & Ebitani, K. Promotion
effect of coexistent hydromagnesite in a highly active solid base hydrotal-
cite catalyst for transesterifications of glycols into cyclic carbonates. Catalysis
Today, 185(1), 241–246, 2012.
34. Zheng, L., Xia, S., Hou, Z., Zhang, M. & Hou, Z. Transesterification of glyc-
erol with dimethyl carbonate over Mg-Al hydrotalcites. Chinese Journal of
Catalysis, 35(3), 310–318, 2014.
35. Zheng, L., Xia, S., Lu, X. & Hou, Z. Transesterification of glycerol with
dimethyl carbonate over calcined Ca-Al hydrocalumite. Chinese Journal of
Catalysis, 36(10), 1759–1765, 2015.
36. Lanjekar, K. & Rathod, V. K. Utilization of glycerol for the production of
glycerol carbonate through greener route. Journal of Environmental Chemical
Engineering, 1(4), 1231–1236, 2013.
37. Algoufi, Y. T. & Hameed, B. H. Synthesis of glycerol carbonate by transesteri-
fication of glycerol with dimethyl carbonate over K-zeolite derived from coal
fly ash. Fuel Processing Technology, 1265-11, 2014.
38. Sonnati, M. O., Amigoni, S., Taffin De givenchy, E. P., Darmanin, T., Choulet,
O. & Guittard, F. Glycerol carbonate as a versatile building block for tomor-
row: synthesis, reactivity, properties and applications. Green Chemistry,
15(2), 283–306, 2013.
39. Aresta, M., Dibenedetto, A., Nocito, F. & Pastore, C. A study on the carbox-
ylation of glycerol to glycerol carbonate with carbon dioxide: The role of the
catalyst, solvent and reaction conditions. Journal of Molecular Catalysis A:
Chemical, 257(1–2), 149–153, 2006.
40. Ezhova, N. N., Korosteleva, I. G., Kolesnichenko, N. V., Kuz’min, A. E.,
Khadzhiev, S. N., Vasil’eva, M. A., et al. Glycerol carboxylation to glycerol
carbonate in the presence of rhodium complexes with phosphine ligands.
Petroleum Chemistry, 52(2), 91–96, 2012.
41. Smith patrick, B. & Gross richard, A. (Eds.). Biobased Monomers, Polymers,
and Materials. Vol. 1105. American Chemical Society. 2012.
42. Dasari, M. A., Kiatsimkul, P.-P., Sutterlin, W. R. & Suppes;, G. J. Low-pressure
hydrogenolysis of glycerol to propylene glycol. Applied Catalysis A: General,
281 225–231, 2005.
43. Gandarias, I., Arias, P. L., Fernandez, S. G., Requies, J., El doukkali, M. &
Guemez, M. B. Hydrogenolysis through catalytic transfer hydrogenation:
Glycerol conversion to 1,2-propanediol. Catalysis Today, 195(1), 22–31, 2012.
44. Guo, X., Li, Y., Shi, R., Liu, Q., Zhan, E. & Shen, W. Co/MgO catalysts for
hydrogenolysis of glycerol to 1, 2-propanediol. Applied Catalysis A: General,
371(1–2), 108–113, 2009b.
Hydrogenolysis as a Means of Valorization 163

45. Huang, Z., Cui, F., Kang, H., Chen, J. & Xia, C. Characterization and cat-
alytic properties of the CuO/SiO2 catalysts prepared by precipitation-gel
method in the hydrogenolysis of glycerol to 1,2-propanediol: Effect of resid-
ual sodium. Applied Catalysis A: General, 366(2), 288–298, 2009.
46. Kurosaka, T., Maruyama, H., Naribayashi, I. & Sasaki, Y. Production of
1,3-propanediol by hydrogenolysis of glycerol catalyzed by Pt/WO3/ZrO2.
Catalysis Communications, 9(6), 1360–1363, 2008.
47. Luo, G., Yan, S., Qiao, M., Zhuang, J. & Fan, K. Effect of tin on Ru-B/γ-Al2O3
catalyst for the hydrogenation of ethyl lactate to 1,2-propanediol. Applied
Catalysis A: General, 275(1–2), 95–102, 2004.
48. Pouilloux, Y., Piccirilli, A. & Barrault, J. Selective hydrogenation into oleyl
alcohol of methyl oleate in the presence of Ru•SnAl2O3 catalysts. Journal of
Molecular Catalysis A: Chemical, 108(3), 161–166, 1996.
49. Wang, Y., Zhou, J. & Guo, X. Catalytic hydrogenolysis of glycerol to propane-
diols: a review. RSC Advances, 5(91), 74611–74628, 2015.
50. Huang, L., Zhu, Y. L., Zheng, H. Y., Li, Y. W. & Zeng, Z. Y. Continuous pro-
duction of 1,2-propanediol by the selective hydrogenolysis of solvent-free
glycerol under mild conditions. Journal of Chemical Technology and
Biotechnology, 83(12), 1670–1675, 2008.
51. Ma, L. & He, D. H. Hydrogenolysis of Glycerol to Propanediols Over Highly
Active Ru-Re Bimetallic Catalysts. Topics in Catalysis, 52(6-7), 834–844,
2009.
52. Maris, E. P. & Davis, R. J. Hydrogenolysis of glycerol over carbonsupported
Ru and Pt catalysts. Journal of Catalysis, 249(2), 328–337, 2007.
53. Miyazawa, T., Koso, S., Kunimori, K. & Tomishige, K. Glycerol hydrogenoly-
sis to 1,2-propanediol catalyzed by a heat-resistant ion-exchange resin com-
bined with Ru/C. Applied Catalysis A: General, 329(0), 30–35, 2007.
54. Dalai, A. K., Sharma, R. V. & Kumar, P. Process for hydrogenolysis of glyc-
erol. Google Patents, 2014.
55. Stankowiak, A. & Franke, O. Method for preparing 1,2-propanediol by
hydrogenolysis of glycerol. Google Patents, 2011.
56. Che, T. M. Production of propanediols. Google Patents, 1987.
57. Montassier, C., Giraud, D. & Barbier, J. Polyol Conversion by Liquid Phase
Heterogeneous Catalysis Over Metals. In: M. Guisnet, J. B. C. B. D. D. C. M.
& PEROT, G. (Eds.). Studies in Surface Science and Catalysis (Vol. Volume
41). Elsevier:165–170, 1988.
58. Casale, B. & Gomez, A. M. Method of hydrogenating glycerol. Google
Patents, 1993.
59. Schuster, L. & Eggersdorfer, M. Preparation of 1,2-propanediol. Google
Patents. 1997.
60. Chaminand, J., Djakovitch, L. A., Gallezot, P., Marion, P., Pinel, C. & Rosier,
C. Glycerol hydrogenolysis on heterogeneous catalysts. Green Chemistry,
6(8), 359–361, 2004.
164 Biodiesel Technology and Applications

61. Kusunoki, Y., Miyazawa, T., Kunimori, K. & Tomishige, K. Highly active
metal– acid bifunctional catalyst system for hydrogenolysis of glycerol
under mild reaction conditions. Catalysis Communications, 6(10), 645–649,
2005.
62. Miyazawa, T., Kusunoki, Y., Kunimori, K. & Tomishige, K. Glycerol conver-
sion in the aqueous solution under hydrogen over Ru/C + an ion-exchange
resin and its reaction mechanism. Journal of Catalysis, 240(2), 213–221,
2006.
63. Perosa, A. & Tundo, P. Selective Hydrogenolysis of Glycerol with Raney
Nickel. Industrial & Engineering Chemistry Research, 44(23), 8535–8537,
2005.
64. Mane, R. B., Hengne, A. M., Ghalwadkar, A. A., Vijayanand, S., Mohite, P.
H., Potdar, H. S., et al. Cu:Al Nano Catalyst for Selective Hydrogenolysis of
Glycerol to 1,2-Propanediol. Catalysis Letters, 135(1), 141–147, 2010.
65. Roy, D., Subramaniam, B. & Chaudhari, R. V. Aqueous phase hydrogenolysis
of glycerol to 1,2-propanediol without external hydrogen addition. Catalysis
Today, 156(1–2), 31–37, 2010.
66. Schmidt, S. R., Tanielyan, S. K., Marin, N., Alvez, G. & Augustine, R. L.
Selective Conversion of Glycerol to Propylene Glycol Over Fixed Bed Raney®
Cu Catalysts. Topics in Catalysis, 53(15), 1214–1216, 2010.
67. Chaudhari, R. V., Roy, D. S. & Subramaniam, B. Polyol hydrogenolysis by
in-situ generated hydrogen. Google Patents, 2011.
68. Drent, E. & Jager, W. W. Hydrogenolysis of glycerol. Google Patents, 2000.
69. Bienholz, A., Hofmann, H. & Claus, P. Selective hydrogenolysis of glycerol
over copper catalysts both in liquid and vapour phase: Correlation between
the copper surface area and the catalyst’s activity. Applied Catalysis A:
General, 391(1–2), 153–157, 2011.
70. Bolado, S., Trevino, R. E., Garcia-cubero, M. T. & Gonzalez-benito, G.
Glycerol hydrogenolysis to 1, 2 propanediol over Ru/C catalyst. Catalysis
Communications, 12(2), 122–126, 2010.
71. Guo, L., Zhou, J., Mao, J., Guo, X. & Zhang, S. Supported Cu catalysts for
the selective hydrogenolysis of glycerol to propanediols. Applied Catalysis A:
General, 367(1–2), 93–98, 2009a.
72. Montassier, C., Menezo, J. C., Hoang, L. C., Renaud, C. & Barbier, J. Aqueous
polyol conversions on ruthenium and on sulfur-modified ruthenium. Journal
of Molecular Catalysis, 70(1), 99–110, 1991.
73. Feng, J., Xiong, W., Xu, B., Jiang, W., Wang, J. & Chen, H. Basic oxide-sup-
ported Ru catalysts for liquid phase glycerol hydrogenolysis in an addi-
tive-free system. Catalysis Communications, 46(0), 98–102, 2014.
74. Balaraju, M., Rekha, V., Prasad, P. S. S., Devi, B. L. A. P., Prasad, R. B. N. &
Lingaiah, N. Influence of solid acids as co-catalysts on glycerol hydrogeno-
lysis to propylene glycol over Ru/C catalysts. Applied Catalysis A: General,
354(1–2), 82–87, 2009.
Hydrogenolysis as a Means of Valorization 165

75. Lahr, D. G. & Shanks, B. H. Effect of sulfur and temperature on ruthenium-


catalyzed glycerol hydrogenolysis to glycols. Journal of Catalysis, 232(2),
386–394, 2005.
76. Checa, M., Auneau, F., Hidalgo-carrillo, J., Marinas, A., Marinas, J. M., Pinel,
C., et al. Catalytic transformation of glycerol on several metal systems sup-
ported on ZnO. Catalysis Today, 196(1), 91–100, 2012.
77. Auneau, F., Noel, S., Aubert, G., Besson, M., Djakovitch, L. & Pinel, C. On
the role of the atmosphere in the catalytic glycerol transformation over
­iridium-based catalysts. Catalysis Communications, 16(1), 144–149, 2011.
78. Vasiliadou, E. S., Heracleous, E., Vasalos, I. A. & Lemonidou, A. A. Ru-based
catalysts for glycerol hydrogenolysis—Effect of support and metal precursor.
Applied Catalysis B: Environmental, 92(1–2), 90–99, 2009.
79. Akiyama, M., Sato, S., Takahashi, R., Inui, K. & Yokota, M. Dehydration–
hydrogenation of glycerol into 1,2-propanediol at ambient hydrogen pres-
sure. Applied Catalysis A: General, 371(1–2), 60–66, 2009.
80. Shinmi, Y., Koso, S., Kubota, T., Nakagawa, Y. & Tomishige, K. Modification
of Rh/SiO2 catalyst for the hydrogenolysis of glycerol in water. Applied
Catalysis B: Environmental, 94(3–4), 318–326, 2010.
81. Ma, L. & He, D. Influence of catalyst pretreatment on catalytic properties
and performances of Ru–Re/SiO2 in glycerol hydrogenolysis to propanedi-
ols. Catalysis Today, 149(1–2), 148–156, 2010.
82. Nakagawa, Y., Shinmi, Y., Koso, S. & Tomishige, K. Direct hydrogenolysis
of glycerol into 1,3-propanediol over rhenium-modified iridium catalyst.
Journal of Catalysis, 272(2), 191–194, 2010.
83. Furikado, I., Miyazawa, T., Koso, S., Shimao, A., Kunimori, K. & Tomishige,
K. Catalytic performance of Rh/SiO2 in glycerol reaction under hydrogen.
Green Chemistry, 9(6), 582–588, 2007.
84. Iriondo, A., Barrio, V. L., Cambra, J. F., Arias, P. L., Guemez, M. B.,
Sanchezsanchez, M. C., et al. Glycerol steam reforming over Ni catalysts
supported on ceria and ceria-promoted alumina. International Journal of
Hydrogen Energy, 35(20), 11622–11633, 2010.
5
Current Status, Synthesis, and
Characterization of Biodiesel
Akshay Garg1, Gaurav Dwivedi2*, Prashant Baredar2 and Siddharth Jain1

Department of Mechanical Engineering, College of Engineering Roorkee,


1

Roorkee, India
2
Energy Center, Maulana Azad National Institute of Technology, Bhopal, India

Abstract
Biodiesel is an alkyl ester produced by transesterification of vegetable oils and ani-
mal fats. It can be used in the diesel engine without any changes in the existing tech-
nology of the engine. Biodiesel is proved as an environment-friendly fuel with low
emissions. Thus, this chapter focuses on the policies formulated by the Indian gov-
ernment to promote the production and commercialization of biodiesel in India.
Indian government is focusing mainly on the cultivation of non-edible sources on
agricultural waste lands. The chapter covers all the aspects of biodiesel feedstocks,
production techniques, and optimization techniques. The main emphasis is given
on the analytical methods to measure the biodiesel properties. The characterization
of biodiesel as a fuel is based on its properties. Thus, development various analytical
methods used in biodiesel analysis have been discussed in this chapter.

Keywords: Biodiesel, feedstocks, catalyst, optimization, analytical methods

5.1 Introduction
India is the second most populated country and one of the fastest-growing
economies in the world. Being a developing country, the energy consump-
tion is very high. Since past few decades, conventional fuels are fulfilling
this energy demand. Due to large consumption, these sources are depleting
at a high rate. This consumption is estimated to increase with an increase

*Corresponding author: gdiitr2005@gmail.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (167–194) © 2021 Scrivener Publishing LLC

167
168 Biodiesel Technology and Applications

in population in the near future. The utilization of fossil fuels to produce


energy contributes to the emission of greenhouse gases increasing global
warming in the environment. In 2017–2018, coal accounted for 70.69%
of the total energy generation and crude for 10.34% and natural gas for
8.7% [1]. In India, the amount of energy consumption is expected to rise
to 2280 BkWh by 2021–2022 and about 4,500 BkWh by 2031–2032 [2].
India’s crude oil reserves is about 0.92% of total crude oil production in
the world, while it consumes about 4.96% of total world consumption [1].
The crude price has been increasing rapidly in the past few years. This rise

Table 5.1 Consumption of crude oil in India (MT) [1, 4].


S. no. Year Native production Import Total % of import
1 1971 6.8 11.7 18.5 63
2 1981 10.5 16.2 26.7 61
3 1991 33.0 20.7 53.7 39
4 2001 32.0 57.9 89.9 64
5 2003–2004 33.4 90.4 123.8 73
6 2004–2005 33.98 100.0 133.98 75
7 2005–2006 32.19 105.0 137.19 77
8 2006–2007 33.99 120.0 153.99 78
9 2007–2008 36.68 148.0 184.68 80
10 2008–2009 33.51 132.78 166.28 80
11 2009–2010 33.69 159.26 192.95 82
12 2010–2011 37.68 163.60 201.28 81
13 2011–2012 38.09 171.73 209.82 82
14 2012–2013 37.86 184.80 222.66 83
15 2013–2014 37.79 189.24 227.03 83
16 2014–2015 37.46 189.43 226.90 83
17 2015–2016 36.4 202.85 239.79 84
18 2016–2017 36.01 213.93 249.94 85
19 2017–2018 35.68 220.43 256.12 86
Synthesis and Characterization of Biodiesel 169

in crude prices is severely affecting the developing countries like India.


Conventional oil meets about 95% of the demand for the transport sector
and has been steadily rising [3]. The demand for crude oil and its produc-
tion capacity (MT) of India is given in Table 5.1.
Table 5.1 indicates that large amount of crude oils is imported in India
for its domestic consumption.
In the present energy scenario, sustainable production of energy in an
effective manner is a huge concern, to mitigate the GHG emissions. Solar,
wind, biomass, and hydro are renewable. The government of India with the
help of policymakers is trying to balance the energy demand and environ-
mental protection with fossil fuels and other sources of energy.

5.2 Status of Biodiesel in India


In recent decades, the Indian Government is focusing on promotion of
biofuel production. The government agencies have introduced certain pol-
icies to encourage investment in this field.
The country has recognized biofuels as the most promising factor in fulfill-
ing the energy demand of the nation. Biofuels are non-toxic renewable fuels
and their production would lead to less import of crude oil and the global
concern of containment of carbon emissions. The transport sector plays the
major role in oil consumption and emissions. The government has intro-
duced various policies regarding vehicle emission standards. Thus, the use of
alternative fuels has become a priority to meet the standards to control air pol-
lution. Biofuels can meet energy demands in an environmentally and cost-
effective manner while reducing the dependency on the import of fossil fuels.
As per the Indian government policy of 2019, more attention is to be
given on the waste agricultural lands for the cultivating trees bearing
non-edible oil seeds for the biofuel production. The blend percentage of
biofuel in diesel is yet to be decided because of uncertainty in the availabil-
ity of biofuel. Utilization of the by-products of the biodiesel production
process is also promoted to maximize the use of waste [3].

5.3 Biodiesel Production in India


5.3.1 Feedstocks Popular in India
As per the Indian government policy, the study focuses on different feed-
stocks of biodiesel from the non-edible oils. The Fuel properties of differ-
ent oils are mentioned in Table 5.2.
170

Table 5.2 Fuel properties of oils [18, 21–36].


Density at Heating value Flash point Viscosity at 40°C Cetane Iodine Acid
Oil 15°C (MJ/Kg) (°C) (mm2/s) number number value
Diesel 820 43.8 70 3.5-5 45-55 220 0.07
Jatropha oil 916 38.96 211.7 37.28 46-55 97.9 11.89
Pongamia oil 933 35.992 222 39.9 32 90.7 4.73
Mahua oil 920 36.85 232 24.58 57 71 38
Neem oil 929 39.84 214 38.875 41 75.295
Linseed oil 924 39.3 241 26.24 34.6 188 1
Rubber seed oil 920 38.64 257 11.22 49.73 134.51 1.68
Biodiesel Technology and Applications

Tobacco oil 918 39.4 220 27.7 38.7 135 -


Castor oil 960 37.4 228 231.22 42.3 80.5 4.57
Waste cooking oil 937 - 235 50 52 - 21.84
Algal oil 860 41 102 21 59 119.1 0.45
Synthesis and Characterization of Biodiesel 171

5.3.1.1 Jatropha (Jatropha curcas) Oil


Jatropha includes 175 species and belongs to the Euphorbiaceae family. It
is a drought-resistant perennial, growing well in marginal areas, tropical
areas with rainfall of 1,000–1,500 mm per year [5]. It grows quickly and
produces seeds for 50 years and lives about 60 million years. The tree can
have a height ranging from 6 to 12 m with 2.5- to 3-cm length of seeds
[6]. One of the most recognized biodiesel production sources in India is
jatropha tree [4].

5.3.1.2 Pongamia Oil


Humid and subtropical environments are most suitable for pongamia tree
cultivation, with annual rainfall of 500 to 2,500 mm in its natural habitat.
Karanja is mainly found along the coasts and river banks in India [7]. A sin-
gle tree contributes about 8–24 kg of kernels. The maximum height of the
plant is 15–25 m. The pongamia trees are in abundance stage in india, but
in spite of that, pongamia oil has been proven potential oil in India [8, 9].

5.3.1.3 Mahua Oil


Mahua is a forest-based tree bearing non-edible oil sources with a pro-
duction potential of 60 MT per annum in India. Total oil content in the
kernel of mahua is 50% and about 34%–37% can be obtained by the small
expeller. Mahua trees are found in northern tropical parts of India with
a maximum height of 20 m [10]. In addition, 20- to 200-kg seeds can be
produced yearly from a single mahua plant [11].

5.3.1.4 Neem Oil


Since early times in India, neem has been used for protecting food from
pests, insects, and curing some diseases. Neem is grown all over India from
south to north spreading from tropical to subtropical regions. Neem plant
can possess maximum height of 18 m. Neem plant requires 120- to 140-cm
rain yearly for cultivation. Neem plant possess highest productivity after 15
years of plantation with oil content of 20%–30% [13].

5.3.1.5 Linseed Oil


Linseed oil (Linum usitatissimum) belongs to the family Linaceae and has
100 species. It is an herbaceous type crop containing 35–45 wt% oil. Its
172 Biodiesel Technology and Applications

height ranges from 0.3 to 1 m. Adequate moisture and cool temperature


is required during maturity of the plant which leads to an increase in oil
content and oil quality. Each branch of linseed tree consists of 1–10 seeds.
Madhya Pradesh and Uttar Pradesh produce about 70% of linseed oil in
India [14].

5.3.1.6 Rubber Seed Oil


Rubber tree (Hevea brasiliensis) belongs to Euphorbiaceous family. It
grows to a height of up to 34 m. The tree requires heavy rainfall and non-
frost climate condition for growth. The oil content in kernel of rubber seed
is 40%–50% and 50%–60% in seeds [15].

5.3.1.7 Tobacco Oil


Tobacco (Nicotiana tabacum) belongs to Solanaceae family. This plant is
mainly grown to produce cigarettes but due to the presence of the char-
acteristics similar to vegetable oils, this plant can be used for producing
biodiesel. It contains oil content of 35%–49% by weight [16].

5.3.1.8 Castor
Castor (Ricinus communis) belongs to the Eurphorbiaceae family. It is cul-
tivated in hot temperature in tropics and subtropic areas. About 1.8 MT
of castor oil is produced per year in all over the world. The height of the
castor tree can reach upto 3 m [17]. The oil content present in castor seeds
is 46%–55% [16].

5.3.1.9 Waste Cooking Oil


The WCO can be obtained from various sources. The waste oil produced
in restaurants, food processing industries, and fast food industries have
been found to degrade with usage time. Waste cooking oils are cheap and
possess poor food quality. WCO needs acidic pretreatment before using it
for biodiesel production to remove the unwanted residues [18, 19].

5.3.1.10 Algae Oil


Algal oils are found to be potential sources of biofuels to replace conven-
tional fuels to promote sustainability in the world. These are aquatic plants
which remove toxic components from water and can be used in producing
Synthesis and Characterization of Biodiesel 173

biodiesel. These plants take very less time for cultivation and consist of
50% oil by weight [20, 21].

5.3.2 Advantages of Non-Edible Oils


Non-edible oilseed crops have various advantages in concern of their
growth, feedstock, and adaptability which promote their use for biodiesel
production in India.

(1) Non-edible oil sources eliminate the need of fertile land as


they can be cultivated in waste agricultural land with low
fertility and low moisture demand.
(2) They eliminate the use of edible oils and food scarcity for
biodiesel production. Non-edible oils are harmful for human
consumption as there are many toxics present in these oils.
(3) During the conversion process, non-edible feedstocks pro-
duce useful by-products which can also be used power
generation.
(4) Non-edible oils are abundant and biodegradable. They possess
good fuel properties like high heat content, low sulfur content,
and lower aromatic content and are biodegradable [37, 38].

There are many problems faced during the direct use of oils as the fuel in
engines. Due to high viscosity of oils, overhauling of the engine takes place.
Thus, there is a need to perform some modification techniques to make the
oil ready to be used as a fuel in engine. Some commonly used methods are
mentioned in the following section.

5.3.3 Modification Techniques


5.3.3.1 Blending
Direct mixing of oil with diesel fuel to decrease the fuel viscosity and mak-
ing a blend is known as blending technique. Utilization of 100% oil is also
possible, some modifications are required in the engine fuel system [39,
40]. To prevent the modifications in the engine, the oil can be treated by
various techniques to decrease its viscosity.

5.3.3.2 Micro-Emulsification
Blending of oils with alcohols have been reported by many researchers
to decrease viscosity of oil as a fuel. Micro-emulsification improves spray
174 Biodiesel Technology and Applications

characteristics of fuel by explosive vaporization. All micro-emulsions with


higher alcohols like butanol, hexanol, and octanol remain within the range
limit of the properties of the fuel for diesel engines.

5.3.3.3 Cracking
Conversion of a substance into some useful form by application of pyrolysis
process is known as cracking. The pyrolysis can be performed on vegeta-
ble oils and animal fats to yield biodiesel with favorable fuel properties. The
pyrolysis process has been studied worldwide by various researchers from
more than a century, especially in the areas which lack in petroleum deposits.

5.3.3.4 Transesterification
Transesterification is the most commonly applied method to convert oil
into biodiesel of low viscosity. The produced biodiesel has the ability to run
the existing engine technology without any modifications. Triglyceride is
reacted with alcohol in the presence of a catalyst to produce alkyl ester and
glycerol as a by-product. This process results in a noticeable reduction of
viscosity of triglycerides [8, 9]. The method of transesterification is shown
in Figure 5.1.

5.3.4 Biodiesel Production Methodology


5.3.4.1 Catalytic Transesterification
The catalytic transesterification is classified into homogeneous catalytic
transesterification and heterogeneous catalytic transesterification.

Transesterification

Non Catalytic Catalytic

Super Critical
Homogeneous Heterogeneous
Methanol

Figure 5.1 Methods of transesterification [18–21].


Table 5.3 Different parameters in biodiesel production from non-edible oils [18–21] and [27–42].
Methanol-
Type of Catalyst to-oil Reaction Temperature Biodiesel
Oil transesterification concentration ratio time (°C) yield (%) References
Jatropha Homogeneous 1% H2SO4 3:7 (v/v) 6h 65 21.2 [41]
Oil 1% NaOH 3:7 (v/v) 6h 50 90.1
Heterogeneous 5% Na2ZrO3 1:16 8h 65 99.9 [44]
Supercritical - 43:1 4 min 320 100 [45]
Pongamia Homogeneous 1.43% KOH 11.06:1 81.43 56.6 98.4 [27]
Oil min
Heterogeneous 3% CaO/ZnO 10:1 6 min 92.61 [46]
Supercritical - 43:1 90 min 300 81% [47]
Mahua Oil Homogeneous 1.24% H2SO4 0.32 1.26 h 60 [48]
0.7% KOH 0.25 30 min 60 98
Heterogeneous 8% ZnO/Mg 1:7 50 min 50 97 [49]
Supercritical - 50:1 10 min 425 99 [50]
Synthesis and Characterization of Biodiesel

(Continued)
175
176
Table 5.3 Different parameters in biodiesel production from non-edible oils [18–21] and [27–42]. (Continued)
Methanol-
Type of Catalyst to-oil Reaction Temperature Biodiesel
Oil transesterification concentration ratio time (°C) yield (%) References
Neem Oil Homogeneous 0.08 H2SO4 1:8 1h 60 85 [51]
1% NaOH
Heterogeneous 6% eggshells 15:1 2h 65 97 [52]
Supercritical - 50:1 15 min 425 83 [50]
Linseed Homogeneous 0.5% NaOH 9:1 40 min 60 95.99 [31]
Oil
Heterogeneous 0.98% CaO 9.41:1 60 min 30 98.77 [53]
Supercritical - 48:1 15 min 573 98 [54]
Rubber Homogeneous 0.5% H2SO4 1:2 120 min 55 91.05 [55]
Biodiesel Technology and Applications

seed Oil
Heterogeneous 5% CaO 9:1 4h 65 97.84 [56]
Supercritical - 42:1 9 min 350 90.86 [57]
Tobacco Homogeneous 1% KOH 20:1 30 min 60 91 [58]
Oil
Heterogeneous
Supercritical - 43:1 90 min 303.4 92.8 [59]
(Continued)
Table 5.3 Different parameters in biodiesel production from non-edible oils [18–21] and [27–42]. (Continued)
Methanol-
Type of Catalyst to-oil Reaction Temperature Biodiesel
Oil transesterification concentration ratio time (°C) yield (%) References
Castor Oil Homogeneous 1% KOH 9:1 30 min 60 95 [36]
Heterogeneous 11% ZnO/Ni 1:8 60 min 55 95.2 [36]
Supercritical - 43:1 90 min 300 96.5 [60]
Waste Homogeneous 1% H2SO4 3:7 (v/v) 3h 65 21.5 [18]
Cooking 1% NaOH 3:7 (v/v) 3h 50 90.6
Oil
Heterogeneous 10% RS-SO3H 20:1 6h 70 97.71 [61]
Supercritical - 10:1 15 min 300 65 [62]
Algal Oil Homogeneous 3.36% H2SO4 8:1 60.44min 50 89.583 [63]
3.49% NaOH 8:1 73.63min 50 87.421
Heterogeneous 1.56% CaO. 3.2:10 125 min 50 88.89 [21]
Al2O3
Supercritical - 1:9 25 min 255 85.75 [64]
Synthesis and Characterization of Biodiesel
177
178 Biodiesel Technology and Applications

Homogeneous catalysts are mainly used in commercial production


of biodiesel as they have high catalytic activity. Homogeneous catalysts
mainly include KOH, NaOH base catalyst during transesterification, and
HCL, H2SO4 as the acid catalyst during esterification. Esterification is a
pretreatment method performed in case of high FFA oils [42].
Heterogeneous are less active but are reusable. These catalysts have less
catalytic activity as compared to homogeneous catalysts and require high
temperature for complete transesterification [43]. The different parameters
of transesterification process are shown in Table 5.3.

5.3.4.2 Non-Catalytic Transesterification


Supercritical method is based on the process of transesterification in the
absence of any catalyst. This process includes the reaction of oils with alco-
hol at very high temperature, thus increasing high energy consumption
and high equipment cost [43].
Various researchers have produced biodiesel from homogeneous, het-
erogeneous catalyst, and supercritical method. This section focuses on the
advantages and disadvantages of the homogeneous catalyst, heterogeneous
catalysts, and supercritical method. The advantages and disadvantages of
different biodiesel process methods is shown in Table 5.4.

Table 5.4 Advantages and disadvantages of different processes [42, 43, 65–67].
Type Advantages Disadvantages
Homogeneous Low cost, good catalytic Pre-treatment required,
activity, the low low FFA required,
temperature required, the more wastewater
high reaction rate from purification,
saponification
Heterogeneous No pretreatment is required Low catalyst activity and
in case of high FFA, the reaction rate
catalyst is recyclable,
ecofriendly
Supercritical Catalyst is not required The high reaction
in reaction, no effect of temperature and pressure
water and FFA in oil, is required, excess alcohol
high biodiesel quality, the required amount in the
high reaction rate reaction, high equipment
cost.
Synthesis and Characterization of Biodiesel 179

5.3.5 Optimization Methodology for Biodiesel


5.3.5.1 Central Composite Design Technique
Response surface methodology is the most frequently used technique
to design the experiments and optimize the transesterification process.
Optimization is based upon the experimental data obtained by experimen-
tal values of transesterification.
RSM is used to build an empirical model based on statistical and math-
ematical techniques. Output variables depend upon input variables vari-
ables. Design-Expert software is used to develop a mathematical relation
between the response and independent variables and its optimization.
Central composite design (CCD) is a technique to perform RSM. Various
studies have been done using CCD as the optimization technique for bio-
diesel production. C.H. Ali et al. used five-level, four-factor CCD for opti-
mizing biodiesel production using WCO and obtained a yield of 85.8% at
reaction temperature of 44.2°C, methanol oil ratio (3.5:1), with an amount
of lipase 0.782 g with an incubation period of 24 h [68]. P. Goyal et al. used
five levels of four-factor CCD to optimize the yield of high FFA jatropha
oil. The yield of 98.3% was obtained. Fifty-four experiments for both ester-
ification and transesterification each were performed with increases run-
ning cost and time consumption [69]. T. Mathimani et al. [70] used eight
assays, four center points, and six axial points to predict the lipid content of
algae on input parameters as incubation days and pH. M. Balaji et al. used
CCD to optimize the biodiesel yield of Ceiba pentandra oil using three
center points. A biodiesel yield of 98.73% was obtained by the optimum
parameters of 2.68 wt% catalyst, 11.46 methanol-to-oil ratio and 106-min
reaction time [71].

5.3.5.2 Box Behnken Technique


Box Behnken design technique is another optimization technique of RSM.
Unlike CCD, BBD has a smaller number of experiments for the same num-
ber of independent variables for biodiesel production. BBD always have
three levels per factor, whereas CCD can have up to 5. Thus, it is less expen-
sive to run BBD as compared to CCD. Gaurav et al. used the BBD tech-
nique to optimize biodiesel production from Pongamia oil. The reaction
conditions of (11.06:1) methanol-to-oil ratio, KOH catalyst 1.43 (w/w%),
and 81.43 min at a reaction temperature of 56.6°C was found optimum to
obtain FAME yield of 98.4% [27]. V. Narula et al. [72] optimized the bio-
diesel yield of algae-jatropha oil blend using BBD technique and obtained
180 Biodiesel Technology and Applications

a FAME yield of 81.98% with (3:5) methanol/oil ratio, 0.9 wt% KOH
catalyst in 180 min at a reaction temperature of 50°C. R. Chamola et al.
optimized the biodiesel yield of microalgae using BBD and obtained satis-
factory results [63].

5.4 Properties of Biodiesel


Biodiesel properties mainly depend upon properties of feedstock used,
production process, and the storage technique. The physico-chemical
properties of the fuel characterize the biodiesel quality. It has better lubri-
cating properties than conventional diesel [73]. It is biodegradable, renew-
able, and non-toxic. The calorific value of biodiesel is about 29,442 (kJ/
kg). The density at 25°C was 874 and a very high flash point of 156°C as of
diesel is 60°C, thus making it safer to handle [74]. There are various other
properties on the basis of which the characterization of the quality of bio-
diesel as performed as fuel for CI engine.
Oxidation Stability: The presence of unsaturated fatty acids in alkyl
esters play major role in the oxidation stability. With an increase in unsat-
uration, the tendency of biodiesel to degrade and form insoluble products
also increases. Amount of water content in biodiesel contributes to an
increase in their oxidation. Metal contamination in biodiesel also results
in acceleration of free radical oxidation process; thus, metals have catalytic
behavior in oxidation [75].
Pour Point and Cloud Point: The key cold flow properties of the fuel are
cloud point and pour point. Pour point is the minimum temperature at which
the flow tendency of the fuel decreases. The fuel with a lower degree of unsat-
uration has poor pour points. Cloud point is the minimum temperature at
which the wax starts to crystallize and the fuel starts to solidify. This is the
minimum temperature until which the fuel can be used in an engine [76].
Flash Point: The lowest temperature at which the vapors of fuel ignite
when an ignition source is applied to it. It is inversely proportional to the
volatility of fuel. A higher flash point temperature is necessary for storage
of the fuel. The flashpoint of biodiesel is much higher than conventional
diesel that making it safer to handle.
Cetane Number: The measurement of ignition delay for a fuel is denoted
as its cetane number. A fuel with a high cetane number is considered as a
good fuel because of short ignition delay [76].
Kinematic Viscosity: It is the measurement of the magnitude of the
internal friction of the fluids. During the storage of biodiesel, the precip-
itation occurs, which leads to gum formation and increases the viscosity.
Synthesis and Characterization of Biodiesel 181

High viscosity decreases tendency of atomization of the fuel spray and thus
affecting the proper combustion of the fuel [77].

5.5 Analytical Methods


All the above-mentioned properties in (Section 5.4) can be measured by
the analytical methods. These properties are the standards to characterize
biodiesel and its blends as fuel. The quality control of biodiesel is necessary
as the presence of contaminants may lead to severe operational and envi-
ronmental hazards. The majorly used methods for monitoring biodiesel
production process and determination of biodiesel quality are chromato-
graphic and spectroscopic methods. The various biodiesel properties and
respective measuring apparatus is shown in Table 5.5.

5.5.1 Titration
Biodiesel consists of methyl esters of fatty acids and always a portion of free
fatty acids. The acidity of FFAs is very low but they act as acids during the
titration method for determining acid value. Thus, two methods, potentio-
metric titration method and the titration method with two indicators, were
introduced to find the acid number of the fuel [78]. The potentiometric
titration method was found more reliable, giving separate concentrations
of strong acids and weak free fatty acids simultaneously. A new study pro-
posed the use of coulometric titration for determining the acid number. This
method was found as good alternative to the volumetric method because
of its higher precision and no requirement of standard solutions for analy-
sis [79]. Water and LiCl are the major solution in the coulometric titration.
The method was based on constant-current coulometric titration integrated
with potentiometric detection. The method contributes in better accuracy in
determining acid number with reducing waste generation from the analysis.

5.5.2 Chromatic Methods


Monitoring transesterification reaction is a necessary task as all the prop-
erties and contaminants change in this reaction. The first chromatographic
analysis of oils was done by means of thin-layer chromatography (TLC)
using a silica gel. It was observed that about 80% of allyl ester yield was
obtained in 2 min [80]. The presence of alcohol does not interfere with
the estimation of yields by TLC. In another report, TLC with flame ion-
ization detection (TLC/FID) was used through a latroscan TH-10 analyzer
[81]. The FID was operated with hydrogen and air flows. The formation of
182 Biodiesel Technology and Applications

Table 5.5 Biodiesel properties and measuring apparatus [63, 97, 99–109].
Properties Methods Specifications
Oxidation Rancimat • Metrohm 892 Professional Rancimat
Stability • Determine stability according to
national and international standards:
AOCS Cd 12b-92 fats and oils
ISO 6886 vegetable oils and fats
2.4.28.2-93 Fat stability, CDM Japan
• Simultaneous analysis of up to eight
samples.
• User-friendly StabNet software for
instrument control.
• It can measure the oxidation stability
of biodiesel, vegetable oils, fats, and
food.
Karl Fischer • A digital meter for measuring
apparatus conductivity)
(modified) • The reaction vessel of 24 mm in
diameter coupled with 150 mm long
of quartz glass
• Proper piping to allow the air to pass
through.
Water Near-Infrared • The wavelength is measured one at a
spectroscopy time.
• The grating and slit combination is
used to select the wavelength to be
measured.
• The detector produces a spectrum
of a plot of transmittance against
frequency for analysis.
Acid value Titration • Important materials for titration-
pH indicator
Titrant
Equivalence point indicator
UV-VIS Spectra • Absorption spectroscopy in the
ultraviolet-visible spectral region.
• Uses light near-UV and near-infrared
ranges
(Continued)
Synthesis and Characterization of Biodiesel 183

Table 5.5 Biodiesel properties and measuring apparatus [63, 97, 99–109]. (Continued)
Properties Methods Specifications
Ester Gas • GC is equipped with a FID.
Content chromatography • Quantitative analysis is performed in
the presence of nitrogen as a carrier
gas.
• European standard EN14103: 2003 is
used for quantitative analysis.
FTIR Spectroscopy • Relies upon various frequencies of
light to produce a spectrum.
• An assembly of a beam splitter and
two strategic mirrors that act as an
interferometer.
• A laser beam is superimposed for the
operation of the instrument.
• Components in IR spectrometer-
Radiation source
Monochromator
Detector
Sample
(Na+K) Flame Atomic • Used to determine the concentrations
Absorption of Ca2+, Mg2+, Na+, and K+.
Spectroscopy • The concentration of elements greater
than 1 µg ml−1 in biological fluids.
Flash Point Pensky-Martens • Works on the standards ASTM D93,
closed cup ISO 2719, IP 34.
test apparatus • Automatic high-precision flash point
(D-93) testing
• It is used in petroleum, chemical, and
fragrance industries.

mono-, di-, and triglycerides during the reaction was also analyzed along
with the ester yield. Other chromatographic methods are as follows.

5.5.2.1 Gas Chromatography


Capillary gas chromatography (CGC) discussed the analysis of transes-
terified soyabean oil with a Spectra-physics SP100 gas chromatograph
equipped with FID and a 1.8 × 0.32 mm bonded SE-30 fused silica capil-
lary column [82]. Helium was used as a carrier gas. The results were favor-
ble with the actual and measured values of the content of biodiesel. Gas
184 Biodiesel Technology and Applications

chromatography works on the basis of ASTM 6584. The presence of glyc-


erol and glycerides in a fuel determines fuel quality as their presence can
damage engines and cause harmful emissions [83]. After this, an import-
ant method was developed performing trimethylsilylation of glycerol, and
mono, di followed by GC and triglycerides in vegetable oil methyl esters
using CGC [84]. A method of determination of different fatty acid profiles
of three different oils was studied by GC using a fused silica capillary col-
umn at the split ratio of 1:5 [85].

5.5.2.2 High-Performance Liquid Chromatography


The first HPLC method was used for determining the free glycerol content
in esters [86]. High-pressure liquid chromatography combined with pulsed
amperometric detection was used in the method to determine free glycerol
in methyl or ethyl esters (HPLC-PAD). HPLC-PAD was found attractive
because it is simple and fewer operations for analysis. This method was
capable of determining the glycerol and alcohol traces simultaneously. The
limit of detection was 1 lg/g with high accuracy. In another study, HPLC
analysis was performed on soyabean oil biodiesel. HPLC chromatogram
integrated with refractive index (RI) and ultraviolet (UV) detectors in
series was used for determining fatty acid composition [87]. The RI detec-
tor was found universal for all FAMEs, whereas UV detects the unsaturated
FAMEs consisting carbon-carbon double bonds. The FAME composition
and FAME yield at the time of transesterification of soybean oil was esti-
mated. As HPLC requires low temperature during analysis, the risk of
isomerization of double bonds is reduced during analysis [88].

5.5.3 Spectroscopic Methods


Emission of light is used in this method for the analysis. It is discriminated
by the wavelength of light and regions of the electromagnetic spectrum
produced. Transesterification is monitored for the analysis of biodiesel.
Some most commonly used methods of spectroscopy are nuclear magnetic
resonance spectroscopy and infrared spectroscopy [89]. The peaks charac-
terize the conversion of the triglyceride to the FAME yield.

5.5.3.1 Nuclear Magnetic Resonance Spectroscopy


NMR spectroscopy technique is used for the characterization of chemical
structures. NMR method is integrated with hydrogen signals. NMR has
been found to be able to determine the blending level with high accuracy.
Synthesis and Characterization of Biodiesel 185

A study reports that 1H NMR was used to characterize the palm biodiesel
and its blends. The amount of unreacted vegetable oil was detected due to
incomplete transesterification reaction [90]. Another study was performed
for the investigation of results obtained on transesterification of rapeseed
oil. The rate of transesterification was monitored [91]. The catalytic activ-
ity and reaction kinetics were determined in a study for different reaction
parameters [92]. Peak fitting and partial least squares regression (PLS-R)
were integrated with high-field NMR spectroscopy was used to monitor
the FAME concentration with variation in time. The PLS-R method was
found to be better than the peak fitting method. NMR was found much
accurate to determine the reaction kinetics.

5.5.3.2 Infrared Spectroscopy


Infrared spectroscopy is used to study the chemical bonds and the func-
tional groups in the molecules. This method is faster than GC, easy to
use, and can also measure the flashpoint [93]. In this study, spectra were
recorded with the help of a fiber-optic probe. Transesterification was mon-
itored with the help of infrared spectroscopy. In another study, NIR mea-
sured the cetane number of WCO and its biodiesel [94]. NIR absorption
was assumed to be linear with the concentration of organic materials. The
standard error of prediction of CN from WCO was 1.2 with an RPD (ratio
of performance to deviation) of 3.83. Mid- and near-infrared spectroscopy
were used for the prediction of the dynamic and kinematic viscosities of
biodiesel-diesel blends [95]. MIR and NIR were integrated with an FT-IR
spectroscopy machine. The fingerprint spectral region of 550–1,500 cm−1
was observed suitable for MIR spectroscopy and NIR spectroscopy chose
multiple regions which were 1,100–1,500, 1,600–1,700, and 1,800–2,200
nm. Results for both MIR and NIR spectroscopy were found acceptable,
NIR was found more accurate.

5.5.4 Rancimat Method


Rancimat test is performed to check the oxidative stability of biodiesel. The
apparatus is based on the EN14214 standard. It works on the basis of the
aging process. The temperature of the samples is increased and a continu-
ous stream of air is passed through it [100]. The effect of five antioxidants
was studied using the rancimat method to evaluate the oxidation stabil-
ity of fuel produced [96]. The conductivity was continuously measured
by passing a stream of pure air 10 L/h through the samples at 110°C. The
stream swept the volatile oxidation products and collected in an electrical
186 Biodiesel Technology and Applications

conductivity measuring vessel containing deionized water. But the testing


of this parameter is very costly thus a new method was developed. A Karl
Fischer (KF) apparatus was modified to measure the oxidation stability [97].
The working principle was based on Rancimat, i.e., a conductivity-based
determination of volatile degradation products and automatic plotting of
the conductivity against time. The results showed that tests performed with
modified Karl Fischer apparatus were in a good agreement with the values
on Rancimat. Modified Karl Fischer apparatus was of less cost and can be a
replacement of the Rancimat apparatus with slight modification.

5.5.5 Viscometry
The oils are highly viscous to use as a fuel in diesel engines. This is the
main reason for performing transesterification and decrease the viscosity
of the oils and produce biofuels. The viscosity differences of biodiesel from
different feedstocks characterize them. A study was reported by using vis-
cometry for the analysis of the products of transesterification, i.e., FAME
[98]. The FAME yield was investigated with respect to the viscosity curves
of the products. The viscometric analysis was performed for the binary
mixtures of soybean biodiesel/soybean oil. All the samples were tested at
40°C according to ASTM D445. The study showed that viscometry is a
promising analytical method for determining the amount of ester levels in
biodiesel. This is a simple and inexpensive technique.

5.6 Conclusion
The non-edible sources are the most promising feedstocks to meet the gov-
ernment targets of biofuel in India. The study has reported various produc-
tion and analytical methods of biodiesel. The fuel properties were reported
according to ASTM and European standards which are to be considered for
the commercialization of biodiesel in India. More focus is to be concerned
on monitoring the transesterification process and the analytical methods
to characterize biodiesel according to their properties. The monitoring of
transesterification process is mainly performed by the spectroscopy meth-
ods and are very much efficient, whereas chromatography methods are
used to estimate the biodiesel yield. Both the methods work on standards
and can analyze the fuel quality for operation in CI engine. Some modified
methods were developed which can contribute in decreasing the cost of
biodiesel analysis. Thus, the development of alternate analytical methods
is necessary which can be used in analyzing the process as well as the yield.
Synthesis and Characterization of Biodiesel 187

References
1. N. D. Central statistics office, Ministry of statistics and programme imple-
mentation, Government of India, Energy Statistics 2019 (Twenty Sixth
Issue) Central Statistics Office Ministry of Statistics and Programme
Implementation Government of India New Delhi http://mospi.nic.in/
sites/default/files/publication_reports/Energy Statistics 2019-finall.pdf, pp.
1–123, 2019.
2. Garg, P., Energy Scenario and Vision 2020 in India, J. Sustain. Energy
Environ., 2012.
3. B. No and C. G. O. Complex, National policy on biofuels, Explor. Renew.
Altern. Energy Use India, no. 14, pp. 205–214, 2011.
4. Jain, S., and Sharma, M.P., Prospects of biodiesel from Jatropha in India: A
review, Renew. Sustain. Energy Rev., vol. 14, no. 2, pp. 763–771, 2010.
5. Lohan, S.K., Ram, T., Mukesh, S., Ali, M., and Arya, S., Sustainability of bio-
diesel production as vehicular fuel in Indian perspective, Renew. Sustain.
Energy Rev., vol. 25, pp. 251–259, 2013.
6. Thapa, S., Indrawan, N., and Bhoi, P.R., An overview on fuel properties and
prospects of Jatropha biodiesel as fuel for engines, Environ. Technol. Innov.,
vol. 9, pp. 210–219, 2018.
7. Mukta, N., and Sreevalli, Y., Propagation techniques, evaluation and
improvement of the biodiesel plant, Pongamia pinnata (L.) Pierre-A review,
Ind. Crops Prod., vol. 31, no. 1, pp. 1–12, 2010.
8. Dwivedi, G., and Sharma, M.P., Prospects of biodiesel from Pongamia in
India, Renew. Sustain. Energy Rev., vol. 32, pp. 114–122, 2014.
9. Dwivedi, G., Jain, S., and Sharma, M.P., Pongamia as a Source of Biodiesel in
India, Smart Grid Renew. Energy, vol. 02, no. 03, pp. 184–189, 2011.
10. Surana, A.P., Singh, Y., Rajubhai, V.H., Suthar, K., and Sharma, A.,
Development of mahua oil as a lubricant additive and its tribological charac-
teristics, Mater. Today Proc., no. xxxx, pp. 8–12, 2019.
11. Jena, P.C., Raheman, H., Kumar, G.P., and Machavaram, R., Biodiesel pro-
duction from mixture of mahua and simarouba oils with high free fatty acids,
Biomass and Bioenergy, vol. 34, no. 8, pp. 1108–1116, 2010.
12. Ghadge, S.V., and Raheman, H.,Biodiesel production from mahua (Madhuca
indica) oil having high free fatty acids, Biomass and Bioenergy, vol. 28, no. 6,
pp. 601–605, 2005.
13. Ogbuewu, M.U., Odoemenam, I.P., V.U., Obikaonu, H.O., Opara, M.N.,
Emenalom, O.O., Uchegbu, M.C., Okoli, I.C., Esonu, B.O. and Iloeje, The
growing importance of neem.pdf, Research Journal of Medicinal Plant, vol. 5,
no. 3. pp. 230–245, 2011.
14. Dixit, S., Kanakraj, S., and Rehman, A., Linseed oil as a potential resource
for biodiesel: A review, Renew. Sustain. Energy Rev., vol. 16, no. 7, pp. 4415–
4421, 2012.
188 Biodiesel Technology and Applications

15. Ahmad, J., Yusup, S., Bokhari, A., and Kamil, R.N.M., Study of fuel prop-
erties of rubber seed oil based biodiesel, Energy Convers. Manag., vol. 78,
pp. 266–275, 2014.
16. Atabani, A.N., et al., Non-edible vegetable oils: A critical evaluation of oil
extraction, fatty acid compositions, biodiesel production, characteristics,
engine performance and emissions production, Renew. Sustain. Energy Rev.,
vol. 18, pp. 211–245, 2013.
17. Azad, A.K., Rasul, M.G., Khan, M.M.K., Sharma, S.C., Mofijur, M., and
Bhuiya, M.M.K., Prospects, feedstocks and challenges of biodiesel produc-
tion from beauty leaf oil and castor oil: A nonedible oil sources in Australia,
Renew. Sustain. Energy Rev., vol. 61, pp. 302–318, 2016.
18. Jain, S., Sharma, M.P., and Rajvanshi, S., Acid base catalyzed transesteri-
fication kinetics of waste cooking oil, Fuel Process. Technol., vol. 92, no. 1,
pp. 32–38, 2011.
19. Elkady, M.F., Zaatout, A., and Balbaa, O., Production of Biodiesel from Waste
Vegetable Oil via KM Micromixer, J. Chem., vol. 2015, 2015.
20. Kumar, M., Sharma, M.P., and Dwivedi, G., Algae oil as future energy source
in indian perspective, Int. J. Renew. Energy Res., vol. 3, no. 4, pp. 913–921,
2013.
21. Narula, V., Khan, M.F., Negi, A., Kalra, S., Thakur, A., and Jain, S., Low tem-
perature optimization of biodiesel production from algal oil using CaO and
CaO/Al2O3as catalyst by the application of response surface methodology,
Energy, vol. 140, pp. 879–884, 2017.
22. Nabi, M.N., Hoque, S.M.N., and Akhter, M.S., Karanja (Pongamia Pinnata)
biodiesel production in Bangladesh, characterization of karanja biodiesel and
its effect on diesel emissions, Fuel Process. Technol., vol. 90, no. 9, pp. 1080–
1086, 2009.
23. Karmakar, A., Karmakar, S., and Mukherjee, S., Biodiesel production from
neem towards feedstock diversification: Indian perspective, Renew. Sustain.
Energy Rev., vol. 16, no. 1, pp. 1050–1060, 2012.
24. Prasad, L., Experimental Investigation of Performance of Diesel Engine
Working On Diesel and Neem Oil Blends, IOSR J. Mech. Civ. Eng., vol. 1,
no. 4, pp. 48–51, 2012.
25. Tiwari, A.K., Kumar, A., and Raheman, H., Biodiesel production from jatro-
pha oil (Jatropha curcas) with high free fatty acids: An optimized process,
Biomass and Bioenergy, vol. 31, no. 8, pp. 569–575, 2007.
26. Jain, S., and Sharma, M.P., Biodiesel production from Jatropha curcas oil,
Renew. Sustain. Energy Rev., vol. 14, no. 9, p. 3140, 2010.
27. Dwivedi, G., and Sharma, M.P., Application of Box – Behnken design in opti-
mization of biodiesel yield from Pongamia oil and its stability analysis, Fuel,
vol. 145, pp. 256–262, 2015.
28. Sahoo, P.K., and Das, L.M., Process optimization for biodiesel production
from Jatropha, Karanja and Polanga oils, Fuel, vol. 88, no. 9, pp. 1588–1594,
2009.
Synthesis and Characterization of Biodiesel 189

29. Acharya, N., Nanda, P., Panda, S., and Acharya, S., Analysis of properties and
estimation of optimum blending ratio of blended mahua biodiesel, Eng. Sci.
Technol. an Int. J., 2017.
30. Agarwal, A.K., Khurana, D., and Dhar, A., Improving oxidation stability of
biodiesels derived from Karanja, Neem and Jatropha: Step forward in the
direction of commercialisation, J. Clean. Prod., vol. 107, pp. 646–652, 2015.
31. Kumar, R., Tiwari, P., and Garg, S., Alkali transesterification of linseed oil for
biodiesel production, Fuel, vol. 104, pp. 553–560, 2013.
32. Godfrey, O.O., Ifijen, I.H., Mohammed, F.U., Aigbodion, A.I., and Ikhuoria,
E.U., Alkyd resin from rubber seed oil/linseed oil blend: A comparative study
of the physiochemical properties, Heliyon, vol. 5, no. 5, pp. 0–4, 2019.
33. Asuquo, J.E., Anusiem, A.C.I., Etim, E.E., and Chemistry, I., Extraction And
Characterization Of Castor Seed Oil, Internet J. Nutr. Wellness, vol. 8, no. 2,
pp. 109–115, 2012.
34. Giannelos, P.N., Zannikos, F., Stournas, S., Lois, E., and Anastopoulos, G.,
Tobacco seed oil as an alternative diesel fuel: Physical and chemical proper-
ties, Ind. Crops Prod., vol. 16, no. 1, pp. 1–9, 2002.
35. Çon, A.H., and Ug, E., Properties and quality verification of biodiesel pro-
duced from tobacco seed oil, vol. 52, pp. 2031–2039, 2011.
36. Keera, S.T., El Sabagh, S.M., and Taman, A.R., Castor oil biodiesel produc-
tion and optimization, Egypt. J. Pet., vol. 27, no. 4, pp. 979–984, 2018.
37. Khan, T.M.HY., Atabani, A.E., Badruddin, A.I., Badarudin, A., Khayoon,
M.S., and Triwahyono, S., Recent scenario and technologies to utilize
non-edible oils for biodiesel production, Renew. Sustain. Energy Rev., vol. 37,
pp. 840–851, 2014.
38. Rezania, S., et al., Review on transesterification of non-edible sources for
biodiesel production with a focus on economic aspects, fuel properties
and by-product applications, Energy Convers. Manag., vol. 201, no. July,
p. 112155, 2019.
39. Agarwal, D., and Agarwal, A.K., Performance and emissions characteristics
of Jatropha oil (preheated and blends) in a direct injection compression igni-
tion engine, Appl. Therm. Eng., vol. 27, no. 13, pp. 2314–2323, 2007.
40. Agarwal, A.K., and Dhar, A., Experimental investigations of performance,
emission and combustion characteristics of Karanja oil blends fuelled DICI
engine, Renew. Energy, vol. 52, pp. 283–291, 2013.
41. Jain, S., and Sharma, M.P., Biodiesel production from Jatropha oil Biodiesel
production from Jatropha curcas oil, Renew. Sustain. Energy Rev., vol. 14,
no. 9, pp. 3140–3147, 2010.
42. Tariq, M., Ali, S., and Khalid, N., Activity of homogeneous and heterogeneous
catalysts, spectroscopic and chromatographic characterization of biodiesel: A
review, Renew. Sustain. Energy Rev., vol. 16, no. 8, pp. 6303–6316, 2012.
43. Lee, J.S., and Saka, S., Biodiesel production by heterogeneous catalysts and
supercritical technologies, Bioresour. Technol., vol. 101, no. 19, pp. 7191–
7200, 2010.
190 Biodiesel Technology and Applications

44. Martínez, A., Mijangos, G.E., Romero-Ibarra, I.C., Hernández-Altamirano,


R., and Mena-Cervantes, V.Y., In-situ transesterification of Jatropha curcas L.
seeds using homogeneous and heterogeneous basic catalysts, Fuel, vol. 235,
no. January 2018, pp. 277–287, 2019.
45. Hawash, S., Kamal, N., Zaher, F., Kenawi, O., and Diwani, G.E1., Biodiesel
fuel from Jatropha oil via non-catalytic supercritical methanol transesterifi-
cation, Fuel, vol. 88, no. 3, pp. 579–582, 2009.
46. Joshi, G., Rawat, D.S., Sharma, A.K., and Pandey, J.K., Microwave enhanced
alcoholysis of non-edible (algal, jatropha and pongamia) oils using chem-
ically activated egg shell derived CaO as heterogeneous catalyst, Bioresour.
Technol., vol. 219, pp. 487–492, 2016.
47. Ortiz-Martínez, V.M., et al., In-depth study of the transesterification reac-
tion of Pongamia pinnata oil for biodiesel production using catalyst-free
supercritical methanol process, J. Supercrit. Fluids, vol. 113, pp. 23–30, 2016.
48. Ghadge, S.V., and Raheman, H., Process optimization for biodiesel produc-
tion from mahua (Madhuca indica) oil using response surface methodology,
Bioresour. Technol., vol. 97, no. 3, pp. 379–384, 2006.
49. Baskar, G., Gurugulladevi, A., Nishanthini, T., Aiswarya, R., and Tamilarasan,
K., Optimization and kinetics of biodiesel production from Mahua oil using
manganese doped zinc oxide nanocatalyst, Renew. Energy, vol. 103, pp. 641–
646, 2017.
50. Lamba, N., Modak, J.M., and Madras, G., Fatty acid methyl esters synthe-
sis from non-edible vegetable oils using supercritical methanol and methyl
tert-butyl ether, Energy Convers. Manag., vol. 138, pp. 77–83, 2017.
51. Thangaraj, B., Ramachandran, K.B., and Raj, S.P., Homogeneous Catalytic
Transesterification of Renewable Azadirachta indica (Neem) Oil and Its
Derivatives to Biodiesel Fuel via Acid/Alkaline Esterification Processes,
IBIMA Publ. Int. J. Renew. Energy Biofuels Int. J. Renew. Energy Biofuels, vol.
2014, 2014.
52. Oladipo, A.S., et al., Magnetic recyclable eggshell-based mesoporous catalyst
for biodiesel production from crude neem oil: Process optimization by cen-
tral composite design and artificial neural network, Comptes Rendus Chim.,
vol. 21, no. 7, pp. 684–695, 2018.
53. Gargari, M.H., and Sadrameli, S.M., A single-phase transesteri fi cation of
linseed oil using different co-solvents and hydrogel in the presence of cal-
cium oxide: An optimization study, Renew. Energy, vol. 139, pp. 426–434,
2019.
54. Demirbas, A., Production of biodiesel fuels from linseed oil using methanol
and ethanol in non-catalytic SCF conditions, Biomass and Bioenergy, vol. 33,
no. 1, pp. 113–118, 2009.
55. Widayat, Wibowo, A.D.K., and Hadiyanto, Study on production process of
biodiesel from rubber seed (hevea brasiliensis) by in situ (trans)esterification
method with acid catalyst, Energy Procedia, vol. 32, pp. 64–73, 2013.
Synthesis and Characterization of Biodiesel 191

56. Sai, B.A.V.S.L., Subramaniapillai, N., Khadhar Mohamed, M.S.B., and


Narayanan, A., Optimization of continuous biodiesel production from
rubber seed oil (RSO) using calcined eggshells as heterogeneous catalyst,
J. Environ. Chem. Eng., vol. 8, no. 1, p. 103603, 2020.
57. Shokib, A., Gumanti, P., and Rachimoellah, M.R., Biodiesel Production from
Rubber Seed Oil by Supercritical Methanol Method, IPTEK J. Technol. Sci.,
vol. 21, no. 2, pp. 21–24, 2010.
58. Veljković, V.B., Lakićević, S.H., Stamenković, O.S., Todorović, Z.B., and
Lazić, M.L., Biodiesel production from tobacco (Nicotiana tabacum L.) seed
oil with a high content of free fatty acids, Fuel, vol. 85, no. 17–18, pp. 2671–
2675, 2006.
59. García-Martínez, N., et al., Optimization of non-catalytic transesterification
of tobacco (Nicotiana tabacum) seed oil using supercritical methanol to bio-
diesel production, Energy Convers. Manag., vol. 131, pp. 99–108, 2017.
60. Román-Figueroa, C., Olivares-Carrillo, P., Paneque, M., Palacios-Nereo, F.J.,
and uesada-Medina, J., High-yield production of biodiesel by non-catalytic
supercritical methanol transesterification of crude castor oil (Ricinus com-
munis), Energy, vol. 107, pp. 165–171, 2016.
61. Mohamed, R.M., Kadry, G.A., Abdel-samad, H.A., and Awad, M.E., High
operative heterogeneous catalyst in biodiesel production from waste cooking
oil, Egypt. J. Pet., no. xxxx, 2019.
62. Patil, P., Deng, S., Rhodes, J.I., and Lammers, P.J., Conversion of waste cook-
ing oil to biodiesel using ferric sulfate and supercritical methanol processes,
Fuel, vol. 89, no. 2, pp. 360–364, 2010.
63. Chamola, R., Khan, M.F., Raj, A., Verma, M., and Jain, S., Response sur-
face methodology based optimization of in situ transesterification of dry
algae with methanol, H2SO4 and NaOH, Fuel, vol. 239, no. November 2018,
pp. 511–520, 2019.
64. Patil, P.D., et al., Optimization of direct conversion of wet algae to biodiesel
under supercritical methanol conditions, Bioresour. Technol., vol. 102, no. 1,
pp. 118–122, 2011.
65. Mohadesi, M., Aghel, B., Maleki, M., and Ansari, A., Production of bio-
diesel from waste cooking oil using a homogeneous catalyst: Study of semi-­
industrial pilot of microreactor, Renew. Energy, vol. 136, pp. 677–682, 2019.
66. Hsiao, M.C., Hou, S.S., Kuo, J.Y., and Hsieh, P.H., Optimized conversion of
waste cooking oil to biodiesel using calcium methoxide as catalyst under
homogenizer system conditions, Energies, vol. 11, no. 10, p. 117114, 2018.
67. Aboelazayem, O., Gadalla, M., and Saha, B., Valorisation of high acid
value waste cooking oil into biodiesel using supercritical methanolysis:
Experimental assessment and statistical optimisation on typical Egyptian
feedstock, Energy, vol. 162, pp. 408–420, 2018.
68. Ali, C.H., Qureshi, A.S., Mbadinga, S.M., Liu, J.F., Yang, S.Z., and Mu, B.Z.,
Biodiesel production from waste cooking oil using onsite produced purified
192 Biodiesel Technology and Applications

lipase from Pseudomonas aeruginosa FW_SH-1: Central composite design


approach, Renew. Energy, vol. 109, pp. 93–100, 2017.
69. Goyal, P., Sharma, M.P., and Jain, S., Optimization of Esterification and
Transesterification of High FFA Jatropha Curcas Oil Using Response Surface
Methodology, J. Pet. Sci. Res., vol. 1, no. 3, pp. 36–43, 2012.
70. Mathimani, T., Beena Nair, B., and kumar, R.R., Evaluation of microalga for
biodiesel using lipid and fatty acid as a marker – A central composite design
approach, J. Energy Inst., vol. 89, no. 3, pp. 436–446, 2016.
71. Balajii, M., and Niju, S., A novel biobased heterogeneous catalyst derived
from Musa acuminata peduncle for biodiesel production – Process optimi-
zation using central composite design, Energy Convers. Manag., vol. 189, no.
March, pp. 118–131, 2019.
72. Narula, V., Thakur, A., Uniyal, A., Kalra, S., and Jain, S., Process parameter
optimization of low temperature transesterification of algae-Jatropha Curcas
oil blend, Energy, vol. 119, pp. 983–988, 2017.
73. Li, F., Liu, Z., Ni, Z., and Wang, H., Effect of biodiesel components on its
lubrication performance, J. Mater. Res. Technol., vol. 8, no. 5, pp. 3681–3687,
2019.
74. Dwivedi, G., and Sharma, M.P., Performance evaluation of diesel engine
using biodiesel from pongamia oil, Int. J. Renew. Energy Res., vol. 3, no. 2,
pp. 325–330, 2013.
75. Jain, S., and Sharma, M.P., Oxidation, Thermal, and Storage Stability Studies
of Jatropha Curcas Biodiesel, ISRN Renew. Energy, vol. 2012, pp. 1–15,
2012.
76. Altitude, H., Research, E., and Refining, P., Combustion of Fat and Vegetable
Oil Derived Fuels in Diesel Engines Michael S. Graboski* and Robert L.
McCormick, Science (80-. )., vol. 24, no. 97, pp. 125–164, 1998.
77. Agarwal, A.K., Chaudhury, V., Agarwal, A., and Shukla, P.C., Comparative
study of macroscopic spray parameters and fuel atomization behaviour of
straight vegetable oils (jatropha), its biodiesel and blends, Therm. Sci., vol. 17,
no. 1, pp. 217–232, 2013.
78. Komers, K., Skopal, F., and Stloukal, R., Determination of the Neutralization
Number for Biodiesel Fuel Production, Fett/Lipid, vol. 99, no. 2, pp. 52–54,
1997.
79. Gonzaga, F.B., and Sobral, S.P., A new method for determining the acid num-
ber of biodiesel based on coulometric titration, Talanta, vol. 97, pp. 199–203,
2012.
80. Gauglitz, E.J., and Lehman, L.W., The preparation of alkyl esters from highly
unsaturated triglycerides, J. Am. Oil Chem. Soc., vol. 40, no. 5, pp. 197–198,
1963.
81. Freedman, B., Pryde, E.H., and Mounts, T.L., Hour Screening Test for
Alternate Fuels in Energy Notes for, Variables Affecting the Yields of Fatty
Esters from Transesterified Vegetable Oils 1, Am. Soc. Agric. Eng., vol. 2,
no. 10, pp. 385–390, 1981.
Synthesis and Characterization of Biodiesel 193

82. Freedman, B., Kwolek, W.F., and Pryde, E.H., Quantitation in the analysis
of transesterified soybean oil by capillary gas chromatography 1, J. Am. Oil
Chem. Soc., vol. 63, no. 10, pp. 1370–1375, 1986.
83. Mittelbach, M., Diesel fuel derived from vegetable oils, VI: Specifications and
quality control of biodiesel, Bioresour. Technol., vol. 56, no. 1, pp. 7–11, 1996.
84. Plank, C., and Lorbeer, E., Simultaneous determination of glycerol, and
mono-, di- and triglycerides in vegetable oil methyl esters by capillary gas
chromatography, J. Chromatogr. A, vol. 697, no. 1–2, pp. 461–468, 1995.
85. Refaat, A.A., Attia, N.K., Sibak, H.A., El Sheltawy, S.T., and ElDiwani, G.I.,
Production optimization and quality assessment of biodiesel from waste veg-
etable oil, Int. J. Environ. Sci. Technol., vol. 5, no. 1, pp. 75–82, 2008.
86. Lozano, P., Chirat, N., Graille, J., and Pioch, D., Measurement of free glyc-
erol in biofuels, Fresenius. J. Anal. Chem., vol. 354, no. 3, pp. 319–322, 1996.
87. Shang, N.C., Liu, R.Z., Chen, Y.H., Chang, C.Y., and Lin, R.H., Characterization
of fatty acid methyl esters in biodiesel using high-performance liquid chro-
matography, J. Taiwan Inst. Chem. Eng., vol. 43, no. 3, pp. 354–359, 2012.
88. Czauderna, M., and Kowalczyk, J., Separation of some mono-, di- and tri-­
unsaturated fatty acids containing 18 carbon atoms by high-performance
liquid chromatography and photodiode array detection, J. Chromatogr. B
Biomed. Sci. Appl., vol. 760, no. 1, pp. 165–178, 2001.
89. Monteiro, M.R., Ambrozin, A.R.P., Lião, L.M., and Ferreira, A.G., Critical
review on analytical methods for biodiesel characterization, Talanta, vol. 77,
no. 2, pp. 593–605, 2008.
90. Ng, M.H., and Yung, C.L., Nuclear magnetic resonance spectroscopic char-
acterisation of palm biodiesel and its blends, Fuel, vol. 257, no. August,
p. 116008, 2019.
91. Gelbard, G., Brès, O., Vargas, R.M., Vielfaure, F., and Schuchardt, U.F., 1H
nuclear magnetic resonance determination of the yield of the transesterifi-
cation of rapeseed oil with methanol, J. Am. Oil Chem. Soc., vol. 72, no. 10,
pp. 1239–1241, 1995.
92. Singh, K., Kumar, S.P., and Blümich, B., Monitoring the mechanism and
kinetics of a transesterification reaction for the biodiesel production with
low field 1H NMR spectroscopy, Fuel, vol. 243, no. January, pp. 192–201,
2019.
93. Knothe, G., Rapid monitoring of transesterification and assessing biodiesel
fuel quality by near-infrared spectroscopy using a fiber-optic probe, JAOCS,
J. Am. Oil Chem. Soc., vol. 76, no. 7, pp. 795–800, 1999.
94. García-Martín, J.F., Alés-Álvarez, F.J., López-Barrera, M.D.C., Martín-
Domínguez, I., and Álvarez-Mateos, P., Cetane number prediction of waste
cooking oil-derived biodiesel prior to transesterification reaction using near
infrared spectroscopy, Fuel, vol. 240, no. June 2018, pp. 10–15, 2019.
95. Zhang, W., Yuan, W., Zhang, X., and Coronado, M., Predicting the dynamic
and kinematic viscosities of biodiesel-diesel blends using mid- and near-­
infrared spectroscopy, Appl. Energy, vol. 98, pp. 122–127, 2012.
194 Biodiesel Technology and Applications

96. Zhou, J., Xiong, Y., and Liu, X., Evaluation of the oxidation stability of bio-
diesel stabilized with antioxidants using the Rancimat and PDSC methods,
Fuel, vol. 188, pp. 61–68, 2017.
97. Jain, S., and Sharma, M.P., Measurement of the oxidation stability of bio-
diesel using a modified Karl Fischer apparatus, JAOCS, J. Am. Oil Chem. Soc.,
vol. 88, no. 7, pp. 899–905, 2011.
98. Sousa, F.P., Luciano, M.A., and Pasa, V.M.D., Thermogravimetry and
Viscometry for Assessing the Ester Content (FAME and FAEE), Fuel Process.
Technol., vol. 109, pp. 133–140, 2013.
99. Bampi, M., Scheer, A.D.P., and Castilhos, F.D., Application of near infrared
spectroscopy to predict the average droplet size and water content in bio-
diesel emulsions, Fuel, vol. 113, pp. 546–552, 2013.
100. Delfino, J.R., et al., A simple and fast method to determine water content
in biodiesel by electrochemical impedance spectroscopy, Talanta, vol. 179,
pp. 753–759, 2018.
101. Verma, P., Sharma, M.P., and Dwivedi, G., Evaluation and enhancement
of cold flow properties of palm oil and its biodiesel, Energy Reports, vol. 2,
pp. 8–13, 2016.
102. Conceição, J.N., et al., Evaluation of molecular spectroscopy for predicting
oxidative degradation of biodiesel and vegetable oil: Correlation analysis
between acid value and UV–Vis absorbance and fluorescence, Fuel Process.
Technol., vol. 183, no. August 2018, pp. 1–7, 2019.
103. Putra, M.D., Nata, I.F., and Irawan, C., Biodiesel production from waste
cooking oil using heterogeneous catalyst: Biodiesel product data and its
characterization, Data Br., vol. 28, p. 104879, 2020.
104. Rosset, M., and Perez-Lopez, O.W., FTIR spectroscopy analysis for monitor-
ing biodiesel production by heterogeneous catalyst, Vib. Spectrosc., vol. 105,
no. June, p. 102990, 2019.
105. Campbell, N.R., and Ingram, J.C., Characterization of 234U/238U Activity
Ratios and Potential Inorganic Uranium Complexation Species in Unregulated
Water Sources in the Southwest Region of the Navajo Reservation. Elsevier
Inc., 2014.
106. https://www.metrohm.com/en-in/products/stability-measurement/
rancimat/
107. http://delloyd.50megs.com/MOBILE/infrared_spectroscopy.html
108. https://www.iitk.ac.in/dordold/index.php?option=com_content&view=­cate
gory&layout=blog&id=219&Itemid=238
109. https://www.anton-paar.com/corp-en/products/details/pma-500/
6
Commercial Technologies for
Biodiesel Production
Chikati Roick1, Leonard Okonye2, Nkazi Diankanua1 and Gorimbo Joshua2*

School of Chemical and Metallurgical Engineering, Faculty of Engineering and the


1

Built Environment, University of the Witwatersrand, Johannesburg, South Africa


2
Institute for the Development of Energy for African Sustainability (IDEAS),
University of South Africa’s College of Science, Engineering and Technology,
Florida, South Africa

Abstract
The commercial production of biodiesel and other biofuels is a pursuit central
to the modern fuel industry. The chapter reviews various established technolo-
gies that are available for producing of biodiesel. These include mainly thermo-
chemical conversion, biomechanical conversion, direct combustion, and chemical
reaction. These technologies are hardly comparable as they are used to process
different biomass feedstocks and developed to satisfy different aims. Extracting
fuels from renewable feedstocks is the only feasible option thus far to cut down
on fossil fuels utilization, which also has a positive effect on the global health. The
chapter looks at Africa as a case study, with its vast renewable energy sources, and
the continent is expectant to reap massive benefits, with the growing attractiveness
of biodiesel to replace conventional fuels. The anticipated benefits calls for eco-
nomically sound and innovative process technologies for commercial production
to meet the growing demand and compete with conventional petroleum products.
Currently, second-generation biodiesel production is not competitive in Africa
due to the associated high cost of production. To reduce the influence of biofuels
production on food security, more attention should be dedicated to the develop-
ment of modern second-generation biodiesel production technologies in Africa,
to their optimum potential.
Keywords: Biofuels, biomass to fuel, thermochemical conversion, gasification,
anaerobic digestion, fermentation, transesterification

*Corresponding author: joshuagorimbo@gmail.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (195–214) © 2021 Scrivener Publishing LLC

195
196 Biodiesel Technology and Applications

Abbreviation
Institute for the Development of Energy for African Sustainability (IDEAS)
greenhouse gases (GHG)
polycyclic aromatic hydrocarbons (PAHs)
Zimbabwe’s National Biodiesel Feedstock’s Production Program (ZNBFPP)
Biomass Energy Strategy (BEST)
International Energy Agency (IEA)
Fischer Tropsch Synthesis (FTS)

6.1 Introduction
Energy is a necessity for socio-economic growth, as it makes available vital
services for the enhancement of the quality of life, with a direct influence
on all aspects of development [1, 2]. In the wake of the continual surge
in the demand for energy, declining petroleum/fossil fuel production and
energy security issues, combined with adverse effects on the environment,
there is impetus to secure an unconventional and adequate, renewable
energy source, that is reasonably priced, dependable with lesser adverse
effect on the environment for energy services [3–9].
In addition to fossil fuels having a negative impact on our environment,
fossil fuels are reportedly depleting rapidly. As such, extracting fuels from
renewable feedstocks can be a better option to cut down on utilization of
fossil fuels, which also has a favorable effect on the global health. Several
research centers around the world viz a viz Institute for the Development
of Energy for African Sustainability (IDEAS) in South Africa, Exxonmobil,
BP USA continues to fund and conduct research on renewable energy
with emphasis on biofuels. This area of work is undoubtedly relevant with
a transformative potential to ease energy supplies and reduce emissions.
However, the use of food crop to produce the much-needed transportation
fuel may not be an ideal solution for the energy problem; hence, research
focuses on non-food-based biomass.
Alternative utilization of biomass as energy fountainhead to generate liq-
uid fuels (biodiesel) with vast environmental and socio-economic advan-
tages cannot be over emphasized [8, 10–12]. Biomass is defined as the plant
or animal material that can be used for energy production such as electric-
ity or heat. Biomass conversion to biodiesel is a favored choice for future
energy needs and sustainable socio-economic [1, 10]. Biodiesel is one of
the energy carrier derived from biomass, which can be utilized to pro-
vide energy for heating, electricity, as well as transportation. Undoubtedly,
Technologies for Biodiesel Production 197

sources of renewable energy can improve energy supply security [11, 12].
The conversion from biomass sources can be achieved through biochemi-
cal, physicochemical, and thermochemical processes. Biodiesel is unequiv-
ocally the favored pathway to supplant existing petroleum fuels, as it offers
a direct route to decarbonize the transport system and lessen the reliance
on fossil fuel and greenhouse gases (GHG) emission, with a positive con-
tribution to socio-economic development [13–15].
Africa, with vast renewable energy sources, is expectant to reap massive
benefits, with the growing attractiveness of biodiesel to replace conven-
tional fuels. There will be opportunities for rural development, employ-
ment, and economic growth [12, 13, 16]. The anticipated benefits calls for
technically innovative technologies for commercial production to meet the
growing demand and compete with conventional petroleum products [12].
Although development practitioners, resource owners, policy makers, and
other stakeholders want to cash in on the numerous advantages offered
by the technology, early projects did not hit the ground running due to
unforeseen challenges. Some of which include lack of proper regulatory
institutions, lack of appropriate agronomic knowledge, lack of “sustainable
agricultural systems”, investor caution, poor market development, improper
“natural resource management” ability, and the potential socio-economic,
food and social security, as well as environmental impacts [17–19].
Biodiesel offers the following advantages [6–8, 20–27]:

• biodegradability,
• sulfur-free,
• with lower level of polycyclic aromatic hydrocarbons (PAHs)
and n-PAHs emitted,
• enhanced performance, higher cetane number, and flash-
point than conventional diesel,
• reduced emmisions,

as well as improved energy security, some prospect of energy indepen-


dence, and improved socio-economic development [1–3, 15, 28].

6.2 Biodiesel Production


The production of biodiesel from renewable biomass resources is growing
rapidly in commercial significance. The biomass feedstock include but not
limited to vegetable oils, algae, water hyacinth, waste cooking oils, and animal
fats [2, 11, 12, 29–31]. Various conversion processes have been established
198 Biodiesel Technology and Applications

as technological routes for the commercialization of biodiesel [13]. The


starting material and the fuel production pathway determines the naming
of the biofuel produced, for example, one can have first-generation biofuel,
second-­generation, etc. Usually, first-generation biodiesel can be reaped from
oily crops such as sunflower and or derived from waste oil, animal waste via
bio-chemical conversion pathway, while second-generation biodiesel is synthe-
sized from non-food crops via thermo-chemical pathway [6, 7, 13, 32–40]. A
number of first-generation technologies for biofuel production applicable in
Africa exist. Currently, second-generation biodiesel production is not com-
petitive in Africa due to the associated unfavorable capital expenditure. To
reduce the influence of biofuels production on food security, more attention
should be dedicated to establishment of state of the art second-generation bio-
diesel production technologies in Africa, to their optimum potential [1, 41].

6.3 Technologies Used for Biodiesel Production


Various established biomass to biodiesel pathways are documented [2,
8–10, 42, 43]. They include thermochemical conversion (gasification,
pyrolysis, or thermal cracking), biomechanical conversion (fermentation,
anaerobic digestion, etc.), direct combustion, chemical reaction (transes-
terification), ultrasound technology, membrane technology, reactive dis-
tillation technology, micro-emulsions, and blending [2]. The mostly used
technologies are presented in Figure 6.1.

(>700 °C), without combustion, controlled Syngas (CO, H2, CO2)


amount of oxygen and/or steam.

Thermochemical Pyrolysis Syngas, Bio-oil,


conversion above> 430 °C and under pressure. Charcoal

Hydrothermal liquefaction
250-550 °C and high pressures of 5-25 MPa Bio-oil

Anaerobic digestion Methane (CH4),


utilization of microorganisms,
in anoxigenic conditions Hydrogen (H2)
Biomechanical Fermentation Bio-ethanol, Acetone,
Biomass conversion substrate such as sugar. Some yeasts,
25-35 °C. Butanol

Photobiological hydrogen
production Hydrogen (H2)
uses microorganisms and sunlight to turn
H2O, and sometimes organic matter, into H2

Power generation
Direct Bio-oil
combustion °

Chemical reaction alcohol, ester molecule react in either Bio-oil


the presence of an acid or base

Figure 6.1 Four level flow chat for the production of fuel energy from biomass.
Technologies for Biodiesel Production 199

6.3.1 Chemical Reaction (Transesterification)


Transesterification reaction entails reacting an alcohol molecule with an
ester molecule acidic, basic medium, or biological enzyme usually lipases.
Basically, the reaction happens by swapping the organic moiety of an ester
with the one of the alcohol to form a new ester. Catalyzed transesterifica-
tion of triglyceride is the main technique used (first-generation production
technique for biofuel production [12, 38–45]. The catalysts can either be
acids, alkalis (carbonates, alkali metal oxides, and alkaline earth metals),
or enzymes (lipases) [12, 32, 45], with alkali catalysts given more prefer-
ence due to higher reactivity and moderate process conditions [12]. The
key downside of this process catalyst depletion and difficulty in separation,
leading to high cost of production [46].
The general equation for transesterification is given as follows:

Ester + alcohol ↔ ester + alcohol (6.1)

For instance, the transesterication of triglyceride with methanol goes as


follows:

Triglyceride + methanol ↔ glycerol + methyl esters (6.2)

6.3.2 Thermochemical Conversion


Thermochemical conversion can be classified as gasification, pyrolysis, and
liquefaction. This second-generation technology uses high temperature to
gasify or pyrolyse biomass and then trailed by the fuel synthesis. The gas
produced is then processed different liquid and gaseous fuels, after ade-
quate treatment and conditioning to syngas.
Gasification is a thermal process which takes place at elevated tem-
peratures (600°C to 1,000°C) in an oxidizing atmosphere. Oxidizing agent
making up the oxidizing atmosphere typically air and steam and, in some
cases, N2, CO2, O2, or a mixture of these are used. During gasification under
ideal conditions large polymeric moieties or molecules of biomass disin-
tegrate to yield lighter molecules and useful gases. The emergence minute
contaminants such as sulfide (H2S), carbonyl sulfide (COS), and hydrogen
cyanide (HCN) during the process is inevitable. Incomplete conversion of
biomass yields char and tar which in most cases contains trace elements
of both organic and inorganic origin in the feed such heavy metals mainly
mercury and arsenic.
200 Biodiesel Technology and Applications

All inclusive, gasification reaction in different atmosphere is given by


Equation 6.3, and this proceeds via multiple pathways. Equations 6.4–6.10
are side reactions feasible during the gasification process. Equations 6.6–
6.9 are observed when steam is added during the process.
Biomass + air + steam = methane + carbon monoxide + carbon dioxide
+ hydrogen + unreated steam + char + tar

CHxOy+ O2+ H2O CH4+ CO + CO2+ H2+ H2O + C (6.3)

2C + O2 2CO (6.4)

C + O2 CO2 (6.5)

C + 2H2 CH4 (6.6)

CO + H2O CO2+ H2 (6.7)

CH4+ H2O CO + 3H2 (6.8)

C + H2O CO + H2 (6.9)

C + CO2 2CO (6.10)

The product gas from biomass gasification is further processed


downstream for instances via Fischer Tropsch Synthesis (FTS) to yield
hydrocarbons.

(2n+1) H2 + n CO → Cn H(2n+2) + n H2O

where n takes the values of positive whole numbers. The FTS can be opti-
mized to yield more valued products.
Another thermochemical conversion process of significance is pyrolysis.
Pyrolysis (conventional, fast pyrolysis, or flash pyrolysis) by using catalyst
or heat in the lack of air or oxygen is used to reduce the crude oil’s viscosity,
leading to lower pour and flash point, viscosity, and comparatively higher
cetane number to conventional diesel [12, 27, 32, 47–51]. Hydrothermal
liquefaction is mostly applied to wet biomass and it entails using super-
critical fluids (H2O and CO2) to extract bio-liquid. Reaction conditions for
these processes are summarized in Figure 6.1.
Technologies for Biodiesel Production 201

6.3.3 Biomechanical Conversion


Anaerobic digestion, fermentation, and photobiological hydrogen produc-
tion are mature technologies that are already in use for the transformation
of the biomass and other carbon containing waste to useful products.
Anaerobic digestion involves processes such as methanogenesis, hydro-
lysis, acetogenesis, and acidogenesis. The general equation can be repre-
sented as follows:

organic matter → Carbon dioxide + methane

For example, glucose gets digested by anaerobic microorganisms as


follows:

C6H12O6 → 3CO2 + 3CH4

In the case of fermentation, sugars in biomass are broken down to


yield alcohol and carbon dioxide. For example, glucose is broken down as
follows:

C6H12O6 → 2 C2H5OH + 2CO2

Photobiological hydrogen production entails the use of sunlight and


microorganisms to convert sugars in biomasssunlight to turn water and,
sometimes, organic matter into hydrogen.

6.3.4 Direct Combustion


The process involves directly combusting biomass feedstock in surplus air
to heat the boiler to produce steam. Direct combustion is usually executed
inside a furnace or steam turbine. The operation is mostly applicable to
biomass with low percentage moisture.

6.4 Other Technologies in Use for Biodiesel


Production
Ultrasound technology as applied to biofuel synthesis has been studied and
developed. Mixing intensity is a key parameter a progressive transesterifi-
cation reaction [48]. This homogenization method has been successfully
employed for both batch and continuous biodiesel production processes
202 Biodiesel Technology and Applications

[12, 49, 50, 52]. The ultrasound technology presents a swift, innovative, and
cost-effective process with the advantage of fast separation and a reduced
temperature and reaction time to occur, and higher yields compared to
conventional processes [11, 12, 44, 49, 51–53].
Membrane technology: application of membrane reactors for synthe-
sis of bio-derived diesel is being developed commercially in Sub-Saharan
Africa. There is ongoing design, and optimization studies are being car-
ried out to improve the membrane (organic or inorganic in nature) reactor
for operations. Transesterification of lipids may possibly be fused with the
membrane reactor hybrid. If reaction followed by separation is integrated
as a unit operation, there will be reduction in the cost of separation and
recycle requirements, leading to higher conversions [12, 54–56].
Reactive distillation technology: Due to the controllability of the chem-
ical reaction used in producing biodiesel by chemical equilibrium, reac-
tive distillation technology can be successfully applied for production, as
several features of this technique have been explored to improve the yield.
The addition of an extra flash evaporator and a decanter will lead to better-­
quality biodiesel from various feedstock types. The energy requirement of
the process lower (with roughly 45% savings) [12, 57–59].
The energy stored in biomass discussed in the above sections can be
extracted via various means as shown in Figure 6.1. There are many dif-
ferent types of process technologies used. These technologies are hardly
comparable as they are used to process different biomass feedstocks and
developed to satisfy different aims.
Figure 6.1 gives an overview of how energy is extracted from biomass
(especially lignocellulosic biomass). Biomass from lignocellulosic crops
comparatively need much effort to degrade than first-generation feed-
stocks such as sugarcane, corn, soy, palm, or grapeseed etc, so for these
second-­generation feedstocks gasification seems to be the option favored
by researchers though research keeps focusing on improving the yield opti-
mizing the production process. Gasification as one of the thermochemical
conversions of any carbon containing material to synthesis gas (syngas) and
char. Gasification essentially converts carbonaceous materials into mixture of
CO, H2, and CO2 in varying ratios depending of feed source and conditions.
This process involves heating the biomass at elevated temperatures, without
combusting the feed that is tailoring the amounts of O2 and steam amounts.
The gasification of biomass reportedly shows a methodical and effective
way of producing heat, power, and hydrogen production as well as second-
generation biofuels. Attention should therefore be given to state-of-the-art
equipment or setups for biomass gasification, assessing the downstream pro-
cess of converting the syngas to fuel and electricity via Fischer Tropsch.
Technologies for Biodiesel Production 203

6.5 Feedstock Requirement


To successfully guarantee the sustenance of industries in biofuel synthesis,
furnishing enough cheap raw materials (biomass) should not impact on
food production. In Africa, possible suitable feedstocks being considered
include the following [1, 2, 12, 32, 60, 61].
Waste oil: waste oil is the major raw material used for biodiesel synthe-
sis by Southern Africa biodiesel plants (both small and medium scale);
however, inadequate supply and availability pose as a hurdle large-scale
synthesis of biodiesel.
Canola: The exceptional cold-flow properties and low triglyceride con-
tent saturation of canola oil makes it efficient feedstock biodiesel for bio-
diesel production. When crushed, it is possible to extract approximately
44% oil from the seed.
Sunflower and soybean: The seeds of sunflower and soybean (which is
widely cultivated in the Southern African region) yield approximately 30%
and 18% oil.
Jatropha: Jatropha was considered a bio-energy feedstock as oils can be
extracted from the seed, and processed to biodiesel.
Algae: Algae is seen as an alternative and sustainable energy source, with
potentially higher oil yield per hectare compared to vegetable oils. It can
be grown with the aid of waste materials (like sewage) to harvest oil for
biodiesel production.
Other feedstocks being considered include cotton seed, sunflower,
cashew nut, castor oil, pumpkin, rapeseed, sesame seeds, avocados, coco-
nut, soyabean, Canola, rubber seed, oil palm, neem, animal fat/waste,
olives, moringa, karanja, linseed, moringa, oleifera, and tobacco plants.

6.6 Some Problems Facing Commercialization


of Biodiesel in Africa
Food in contrast to fuel factor: The shift from food crops to energy crops
is the main concern to local farmers and the rural communities in Africa,
with most of the countries still in developmental stage. A rapid decline in
food crop production will lead to increase in price and starvation, in par-
ticular among the rural populations [2, 62, 63].
Land: To sustain a beneficial commercialized biodiesel production, vast
land is required for the cultivation of energy crops. Traditionally, land
ownership in Africa is essentially by heritage, usually in small portions,
204 Biodiesel Technology and Applications

as such vast land for energy crop cultivation can be acquired from com-
munity lands, which, in some cases, leads to dispute. Taking land from
socio-economically vulnerable communities/families will have a negative
impact on their livelihood [2, 64].
Finance: The initial cost for the acquisition of resources and infrastruc-
ture for production of biodiesel is high. The inadequate financial arrange-
ments as well as the lack of low interest rate credit facilities is a main
obstruction for the commercialization of biodiesel technology in many
African countries, as small-scale investments is not supported by the cap-
ital markets [2].
Policy hurdles: Lack of coordination among government, energy and sci-
ence ministries, and technology, financial institutions, and research groups
have contributed to the slow growth in biodiesel process development and
commercialization [2].

6.7 Case Studies/Current Status and Future Potential


The commercialization of biodiesel synthesis in Africa is projected to
provide, among other benefits, the opportunity for Africa to have a vari-
ety agricultural activity, energy sources, and contribute immensely to
socio-economic growth in a more sustainable way [3, 65]. Although
biodiesel technology development is nascent in Africa, the large-scale
advanced projects investment plans, limited the commencing of com-
mercial production in many countries in African countries, e.g., Nigeria,
Ghana, Senegal, Liberia, Zambia, Ethiopia, Tanzania, Kenya, and even
South Africa [2, 66, 67]. Other investments include Madagascar, Niger,
Ethiopia, D1 oils in Swaziland, Malawi, Mali, Uganda, Nigerian National
Petroleum Corporation, Angola, International Biofuels Crops in Liberia,
and the Dutch biodiesel equipment manufacturer in Senegal [2].
Biodiesel industry development in most African countries is in the early
stages, with South Africa being an exception (implementations of small to
medium scale) and Zimbabwe (with first ever commercial biodiesel plant
in Africa) [13, 68, 69].
A number of countries in the Southern region of Africa came up with
policies to empower biodiesel production—achieved via the direct fiscal
motivation for biodiesel production or penalties for use of conventional
fuels. Although many biodiesel production routes are fully established, they
are comparatively expensive than the conventional diesel synthesis route,
which makes it difficult for commercialization and widespread adoption of
the technology [13, 69]. Even though South Africa is the most advanced
Technologies for Biodiesel Production 205

nation technologically, tax exemptions and favorable leasing contracts in


neighboring countries attract private investments to these countries.
South Africa: Presently, South Africa has over 200 small-scale biodiesel
production capacity, from waste vegetable oils, and by 2007, South Africa
has implemented a 10% biodiesel blend in petroleum fuels sales at its fill-
ing stations, However, the key concern has been sustained availability of
feedstock, and refining it to specifications of petrochemical industries [2].
Rainbow Nation Renewable Fuels Ltd. (Coega, near Port Elizabeth)
made the first serious commercial investment into biodiesel production by
acquiring a license and erecting a 1.1 Mt per annum use Australian technol-
ogies (Desmet Ballestra). The crushing facility for soybean, it is expected to
produce and distribute roughly 228 ML biodiesel fuel. This Investment of
US $250 M created 350 full‐time jobs, 1,500 in construction and associated
industries, and roughly 5,000–10,000 in agricultural sector [70].
Zimbabwe: Zimbabwe’s National Biodiesel Feedstock’s Production
Program (ZNBFPP) targeted a 10% substitution of imported fossil fuel
daily consumption with biodiesel by the end of 2018 [15]. Zimbabwe’s $6
million biodiesel plant is capable of producing 100 million liters (Ml) of
biodiesel annually, if operated at full capacity, and has the capability to pro-
cess different feedstock, such as sunflower and soya, cotton seed, jatropha,
among others. Roughly $80 million a year is the expected annual savings in
foreign exchange from diesel import by this plant [3, 15].
Zambia: Zambia, with the aid of Marli Investment set up an $8 million
plant, which will utilize jatropha oil as the major raw material for biodiesel
synthesis. Another biodiesel plant capable of producing 60 Ml/year of bio-
diesel started operations in August 2007 [2].
Ghana: Anuanom Industrial Bio-Products and Biodiesel was capable of
producing about 360,000 t/a of biodiesel, but the construction was stalled.
However, Anuanom currently have two production plants, with a com-
bined capacity of 140,000 metric tonnes of jatropha oil. Ghana Energy
Commission targeted a 20% national biodiesel production by 2015, and
in partnership with South Africa-based BD-1Group, Ghana planned a
12,000-hectare project focused on J. curcas production [2, 11, 71, 72].
Other African countries with initiated projects include in place for the
development include the following:
Botswana: Biomass Energy Strategy (BEST) was launched to ensure that
the biomass derived energy is carried out in a social-economic and envi-
ronmentally sustainable way. BEST has the backing of German-EU Energy
Initiative [32].
Tanzania: Ministry of Energy and Minerals established a Biofuels Task
Force in conjunction with other government ministries and institutions.
206 Biodiesel Technology and Applications

The aim is to ensure a sustainable socio-economic, environmental, and


bioenergy development. Sun Biofuels originating from UK contributed 20
million dollars toward the production of J. curcas and processing plant for
biodiesel in Tanzania, with panned expansion to Burkina Faso, Malawi,
and Madagascar [11, 33, 34].
Kenya: The National Biofuels Committee established 14 years ago, to
increase energy supply security and diversify local energy sources via bio-
diesel supplements and reduce imported fossil fuels dependency [32].
Mozambique: In the course of 2006–2007, three important projects
were announced, with Canadian-based Energem Resources investing $4
million on its first plantation. Duelco renewable energy based in South
Africa is in partnership with Mozambique for a plantation that measures
60,000 hectares, while ESV-Bio Africa has a plantation that measures
11,000 hectare and they project to achieve 100,000 hectares. As per the
International Energy Agency (IEA), Mozambique is capable of producing
roughly 3,000,000 barrels/day of biomass derived oil from non-food feed-
stock [11, 73].
Angola—Angola has the largest non-forest agricultural lands in the
world (with over 80,000,000 hectares reserved land for biodiesel crops
growing). If developed, the country’s biodiesel daily export is estimated at
2.7 million barrels of oil [11, 74].

Coal Nuclear energy


Renewables Natural gas
Hydroelectricity Oil
100

90
80

70
60
50

40
30
20

10

North S. & Cent. Europe CIS Middle East Africa Asia Pacific 0
America America

Figure 6.2 Primary energy regional consumption by fuel 2018 in Percentage. BP


Statistical Review of World Energy 2019 (Outlook, 2019) [75].
Technologies for Biodiesel Production 207

Angola and Mozambique are collaboration with Petrobras, Brazil, to


develop soybean-based biodiesel, to aid biofuels supply to industries in
both countries [74].
Renewables are mainly consumed in Asia Pacific and Europe. Together,
these regions account for 70% of global consumption. From Figure 6.2,
Africa and Middle East reap less fuel from renewables in the energy mix.
Oil and natural gas remain the dominant fuel in most parts of the globe.
Global coal consumption is heavily concentrated in Asia Pacific while more
than two thirds of nuclear consumption is concentrated in North America
and Europe. Asia Pacific and South and Central America account for almost
60% of hydro. More than 90% of renewables are consumed in Asia Pacific,
Europe, and North America. Biofuels from renewables are now common
fuels components blended in petroleum derived fuels in many countries.

6.8 Conclusions
The chapter reviews various established technologies that are available
for producing of biodiesel. Extracting fuels from renewable feedstocks is
the only feasible option thus far to cut down on the utilization of fossil
fuels, which also has a positive effect on the global health. They include
thermochemical conversion, biomechanical conversion, direct combus-
tion, chemical reaction, ultrasound technology, membrane technology,
reactive distillation technology, and micro-emulsions. These technologies
are hardly comparable as they are used to process different biomass feed-
stocks and developed to satisfy different aims. The commercialization of
biodiesel synthesis in Africa is projected to provide, among other benefits,
the opportunity for Africa to broaden energy supply and diversify agricul-
tural activities and contribute immensely to socio-economic growth in a
more sustainable way. To reduce the influence of biofuels production on
food security, more attention should be dedicated to the development of
modern second-generation biodiesel production technologies in Africa, to
their optimum potential. Several research centers around the world con-
tinues to fund and conduct research on renewable energy with emphasis
on biofuels. This area of work is undoubtedly relevant with a transforma-
tive potential to ease energy supplies and reduce emissions. However, the
use of food crop to produce the much-needed transportation fuel may not
be an ideal solution for the energy problem; hence, research focuses on
non-food-based biomass.
Biodiesel industry development in African is still in infancy, except for
South Africa and Zimbabwe with Africa’s first ever commercial biodiesel
208 Biodiesel Technology and Applications

plant. Although a number of biodiesel synthesis routes are well developed,


the lack of cost competitiveness with conventional diesel fuel is the fore-
most difficulty for the commercialization and the widespread adoption of
the technology especially in Africa.

Acknowledgments
The authors are grateful for the financial support provided by the
University of South Africa (UNISA), University of the Witwatersrand,
National Research Foundation (NRF) of South Africa, and the Institute for
the Development of Energy for African Sustainability (IDEAS) research
unit at UNISA.

References
1. Amigun, B., Musango, J.K. and Stafford, W. Biofuels and sustainability in
Africa. Renew. Sustain. Energy Rev., 15: 1360–1372, 2011.
2. Yang, L., Takase, M., Zhang, M., Zhao, T. and Wu, X. Potential non-edible
oil feedstock for biodiesel production in Africa: A survey. Renew. Sustain.
Energy Rev., 38: 461–477, 2014.
3. Amigun, B., Sigamoney, R. and Von Blottnitz, H. Commercialisation of bio-
fuel industry in Africa: a review. Renew. Sustain. Energy Rev., 12: 690–711,
2008.
4. Brittaine, R. and Lutaladio, N. Jatropha: a small holder bioenergy crop: the
potential for pro-poor development. Food and Agriculture Organization of
the United Nations (FAO); 2010.
5. Mangoyana, R.B. Bioenergy for sustainable development: an African con-
text. Phys. Chem. Earth, Parts A/B/C, 34: 59–64, 2009.
6. Kumar, S.S. and Purushothaman, K. High FFA rubber seed oil as an alterna-
tive fuel for diesel engine – an overview. Int. J. Eng. Sci., 10 (1):16–24, 2010.
7. Mofijur, M., Masjuki, H.H., Kalam, M.A., Atabani, A.E., Arbab, M.I., Cheng,
S.F., et al. Properties and use of Moringa oleifera biodiesel and diesel fuel
blends in a multi cylinder diesel engine. Energy Convers. Manage, 82:169–76,
2014.
8. Dwivedi, G., Jain, S. and Sharma, M.P. Diesel engine performance and emis-
sion analysis using biodiesel from various oil sources-reviews. J. Mater.
Environ. Sci., 4 (4): 434–47, 2013.
9. Sanjel, N., Gu, J.H. and Oh, S.C. Transesterification kinetics of waste vegeta-
ble oil in supercritical alcohols. Energies, 7: 2095–106, 2014.
Technologies for Biodiesel Production 209

10. Thanh, L.T., Okitsu, K., Boi, L.V. and Maeda, Y. Catalytic technologies for
biodiesel fuel production and utilization of glycerol: a review. Catalysts, 2:
191–222, 2012.
11. Babajide, O. Sustaining Biodiesel Production via Value-Added Applications
of Glycerol. Journal of Energy, Vol. 2013, Article ID 178356, 7.
12. Babajide, O., Petrik, L. and Ameer, F. Technologies for Biodiesel Production
in Sub-Saharan African Countries. In: Biofuels - Status and Perspective.
InTech E-Publishing Co.; p. 39–57, 2015.
13. Stafford, W.H.L., Lotter, G.A., von Maltitz, G.P. and Brent, A.C. Biofuels
technology development in Southern Africa. Development Southern Africa,
36:2, 155-174, 2019.
14. Fukuda, H., Kondo, A. and Noda, H. Biodiesel fuel production by transes-
terification of oils. Journal of Bioscience and Bioengineering, 92(5): 405-416,
2001.
15. Amigun, B., Müller-Langer, F. and von Blottnitz, H. Predicting the costs
of biodiesel production in Africa: learning from Germany. Energy for
Sustainable Development 2008; Vol. XII No. 1 March 2008.
16. Feygin, M. and Satkin, R. The oil reserves-to-production ratio and its proper
interpretation. Natural Resources Research, 13: 57–60, 2004.
17. Willis, K.J., Bennett, K.D., Burrough, S.L., Macias-Fauria, M. and Tovar,
C. Determining the response of African biota to climate change: using the
past to model the future. Philosophical Transactions of the Royal Society B:
Biological Sciences, 368 (1625): 201–204, 2013.
18. Tilman, D., Socolow, R., Foley, J.A., Hill, J., Larson, E., Lynd, L., Pacala, S.,
Reilly, J., Searchinger, T., Somerville, C. and Williams, R. Beneficial biofu-
els-The food, energy, and environment Trilemma. Science, 325 (5938), 270-1,
2009.
19. de Vries, S., van de Ven, G., van Ittersum, M. and Giller, K. Resource use
efficiency and environmental performance of nine major biofuel crops, pro-
cessed by first-generation conversion techniques. Biomass and Bioenergy, 34:
588–601, 2010.
20. Meng, X., Chen, G. and Wang, Y. Biodiesel production from waste cook-
ing oil via alkali catalyst and its engine test. Fuel Processing Technology, 89:
851–857, 2008.
21. Alcantara, R., Amores, J., Canoira, L., Fidalgo, E., Franco, M.J. and Navarro,
A. Catalytic production of biodiesel from soy-bean oil used frying oil and
tallow. Biomass and Bioenergy, 18: 515–527, 2000.
22. Jaya, N. and Ethirajulu, K. Kinetic studies of heterogeneously catalysed
transesterification of cotton seed oil to biodiesel. J. Environ. Res. Develop., 5
(3A):689–695, 2011.
23. Singh, S. and Singh, D. Biodiesel production through the use of different
sources and characterization of oils and their esters as the substitute of diesel:
a review. Renew. Sustain. Energy Rev., 14: 200–216, 2010.
210 Biodiesel Technology and Applications

24. Xue, J. Combustion characteristics, engine performances and emissions of


waste edible oil biodiesel in diesel engine. Renew. Sustain. Energy Rev. 23:
350–365, 2013.
25. Atadashi, I., Aroua, M. and Aziz, A.A. High quality biodiesel and its die-
sel engine application: a review. Renew. Sustain. Energy Rev., 14:1999–2008,
2010.
26. Encinar, J.M., González, J.F., Rodríguez, J.J. and Tejedor, A. Biodiesel fuels
from vegetable oils: transesterification of cynara cardunculus l. oils with eth-
anol. Energy Fuels, 16: 443–450, 2002.
27. Sivaprakasam, S. and Saravanan, C.G. Optimization of the transesterification
process for biodiesel production and use of biodiesel in a compression igni-
tion engine. Energy and Fuels, 21 (5): 2998–3003, 2007.
28. Peters, J. and Thielmann, S. Promoting biofuels: implications for developing
countries. Energy Policy, 36:1538–44, 2008.
29. Demirbas, A. Biodiesel from sunflower oil in supercritical methanol with
calcium oxide. Energy Convers. Manag., 48: 937–941, 2007.
30. Demirbas, A. Biodiesel from vegetable oils with MgO catalytic transester-
ification in supercritical methanol. Energy Sources, Part A, 30: 1645–1651,
2008.
31. Marchetti, J., Miguel, V. and Errazu, A. Possible methods for biodiesel pro-
duction. Renew. Sustain. Energy Rev., 11: 1300–1311, 2007.
32. Walimwipi, H., Yamba, F.D., Wörgetter, M., Rathbauer, J. and Bacovsky, D.
Biodiesel Production in Africa. In R. Janssen and D. Rutz (eds.), Bioenergy
for Sustainable Development in Africa, Springer Science + Business Media
B.V. p. 93–102, 2012.
33. Ahmad, M., Teong, L.K., Zafar, M., Sultana, S., Sadia, H. and Khan, M.A.
Prospects and potential of green fuel from some non-traditional seed oils
used as biodiesel. In: Fang Z, editor. Biodiesel-feedstocks, productions and
applications. InTech E-Publishing Co.; p. 103–26, 2013.
34. Asuquo, J.E., Anusiem, A.C.I. and Etim, E.E. Extraction and characterization
of rubber seed oil. Int. J. Modern Chem., 1 (3): 109–15, 2012.
35. Gimbun, J., Ali, S., Kanwal, C.C.S.C., Shah, L.A., Gazali, N.H.M., Cheng,
C.K., et al. Biodiesel production from rubber seed oil using a limestone
based catalyst. Adv. Mater. Phys. Chem., 2: 138–41, 2012.
36. Krishnakumar, U., Sivasubramanian, V. and Selvaraju, N. Physico-chemical
properties of the biodiesel extracted from rubber seed oil using solid metal
oxide catalysts. Int. J. Eng. Res. Applicat., 3 (4): 2206–2209, 2013.
37. Ashraful, A.M., Masjuki, H.H., Kalam, M.A., Rizwanul Fattah, I.M., Imtenan,
S., Shahir, S.A., et al. Production and comparison of fuel properties, engine
performance and emission characteristics of biodiesel from various non-
edible vegetable oils: a review. Energy Convers, Manage., 80: 202–228, 2014.
38. Ahmad, J., Bokhari, A. and Yusup, S. Optimization and parametric study of
free fatty acid (FFA) reduction from rubber seed oil (RSO) by using response
surface methodology (RSM). Aust. J. Basic Appl. Sci., 8 (5): 299–303, 2014.
Technologies for Biodiesel Production 211

39. Bhandare, P. and Naik, G.R. Biophysicochemical evaluation and micro-


propagation studies on neem for biodiesel production. Int. J. Appl. Biol.
Pharmaceut. Technol., 6 (1): 213–222, 2015.
40. Abdul Khalil, H.P.S., Sri Aprilia, N.A., Bhat, A.H., Jawaid, M., Paridah, M.T.,
Rudi, D. A jatropha biomass as renewable materials for biocomposites and its
applications. Renew. Sustain. Energy Rev., 22: 667–685, 2013.
41. Sorda, G., Banse, M. and Kemfert, C. An overview of biofuel policies across
the world. Energy Policy, 38: 6977–6988, 2010.
42. Yaakob, Z., Mohammad, M., Alherbawi, M., Alam, Z. and Sopian, K.
Overview of the production of biodiesel from Waste cooking oil. Renew.
Sustain. Energy Rev., 18: 184–93, 2013.
43. Oh, P.P., Lau, H.L.N., Chen, J., Chong, M.F. and Choo, Y.M. A review on
conventional technologies and emerging process intensification (PI) meth-
ods for biodiesel production. Renew. Sustain. Energy Rev., 16: 5131–5145,
2012.
44. Kumari, V., Shah, S. and Gupta, M.N. Preparation of biodiesel by lipase cata-
lyzed transesterification of high free fatty acid containing oil from Madhuca
indica. Energy and Fuels, 21: 368–372, 2007.
45. Qiu, Z., Zhao, L. and Weatherley, L. Process intensification technologies in
continuous biodiesel production. Chemical Engineering and Processing, 49:
323–330, 2010.
46. Gnanaprakasam, A., Sivakumar, V.M., Surendhar, A., Thirumarimurugan,
M. and Kannadasan, T. Recent Strategy of Biodiesel Production from Waste
Cooking Oil and Process Influencing Parameters: A Review. Journal of
Energy, Vol. 2013, Article ID 926392, 10 pages, 2013.
47. Babajide, O., Petrik, L., Musyoka, N. and Ameer, F. Novel zeolite Na-X syn-
thesized from fly ash as a heterogeneous catalyst in biodiesel production.
Catalysis Today, 190: 54–60, 2012.
48. Babajide, O., Petrik, L., Amigun, B. and Ameer, F. Low-cost feedstock con-
version to biodiesel via ultrasound technology. Energies, 3: 1691–1703, 2010.
49. Gogate, P.R. and Pandit, A.B. A review and assessment of hydrodynamic cav-
itations as a technology for the future. Ultrasonic Sonochemistry, 12: 21–27,
2005.
50. Stavarache, C., Vinatoru, M., Bandow, H. and Maeda, Y. Ultrasonically
driven continuous process for vegetable oil transesterification. Ultrasonic
Sonochemistry, 14: 413–417, 2007.
51. Refaat, A.A., Attia, N.K., Sibak, H.A., Sheltawy, S.T. and Eldiwani, G.I.
Production optimization and quality assessment of biodiesel from waste veg-
etable oil. Int. J. Environ. Sci. Tech., 5: 75–82, 2008.
52. Thanh, L.T., Okitsu, K., Sadanaga, Y., Takenaka, N., Maeda, Y. and Bandow,
H. Ultrasound assisted production of biodiesel fuel from vegetable oils in a
small scale circulation process. Bioresource Technology, 101: 639–645, 2010.
53. Ji, J., Wang, J., Li, Y., Yu, Y. and Xu, Z. Preparation of biodiesel with the help
of ultrasonic and hydrodynamic cavitation. Ultrasonics, 44: 411–414, 2006.
212 Biodiesel Technology and Applications

54. Cao, P., Tremblay, A.Y., Dube, M.A. and Morse, K. Effect of membrane pore
size on the performance of a membrane reactor for biodiesel production.
Ind. Eng. Chem. Res., 6 (1): 52–58, 2007.
55. Dube, M., Tremblay, A. and Liu, J. Biodiesel production using a membrane
reactor. Bioresour. Technol. 98 (3): 639–647, 2007.
56. Ranchod, N., Grewan, S., Nedambale, N., Matambo, T., Low, M. and Harding,
K. Biodiesel production using a membrane reactor. Chemical Technology,
21–25, 2013.
57. Kiss, A.A., Dimian, A.C. and Rothenberg, G. Biodiesel by catalytic reactive
distillation powered by metal oxides. Energy Fuel, 22 (1): 598–604, 2008.
58. Cossio-Vargas, E., Hernandez, S., Segovia-Hernandez, J.G. and Cano-
Rodriguez, M.I. Simulation study of the production of biodiesel using feed-
stock mixtures of fatty acids in complex reactive distillation columns. Energy,
36 (11): 6289–6297, 2011.
59. Kiss, A.A. Heat-integrated reactive distillation process for synthesis of fatty
esters. Fuel Process Technol., 92 (7): 1288–1296, 2011.
60. Borugadda, V.B. and Goud, V.V. Biodiesel production from renewable feed-
stocks: status and opportunities. Renew. Sustain. Energy Rev., 16: 4763–84,
2012.
61. Muller, A., Schmidhuber, J., Hoogeveen, J. and Steduto, P. Some insights in
the effect of growing bio-energy demand on global food security and natural
resources. Water Policy, 10: 83, 2008.
62. Abbott, P. and de Battisti, A.B. Recent global food price shocks: causes, con-
sequences and lessons for African governments and donors. J. Afr. Econ., 20:
i12–62, 2011.
63. Rosegrant, M.W. Biofuels and grain prices: impacts and policy responses.
Washington, DC: International Food Policy Research Institute; 2008.
64. Cotula, L., Vermeulen, S., Leonard, R. and Keeley, J. Land grab or develop-
ment opportunity? Agricultural investment and international land deals in
Africa, FAO, IIED and FAD, 2009.
65. Amigun, B. and von Blottnitz, H. Investigation of scale economies for
African biogas installations, Energy Conversion and Manag., 48: 3090-3094,
2008.
66. Pillay, D. and Da Silva, E.J. Sustainable development and bioeconomic pros-
perity in Africa: bio-fuels and the South African gateway. Afr. J. Biotechnol.,
8, 2009.
67. Winkler, H. Energy policies for sustainable development in South Africa.
Energy Sustain. Dev., 11: 26–34, 2007.
68. Amigun, B., Musango, J.K. and Brent, A.C. Community perspectives on the
introduction of biodiesel production in the Eastern Cape Province of South
Africa. Energy, 36: 2502-2508, 2011.
69. Chakauya, E., Beyene, G. and Chikwamba, R.K. Food production needs fuel
too: Perspectives on the impact of biofuels in Southern Africa. South African
Journal of Science, 105 (5–6): 174–181, 2009.
Technologies for Biodiesel Production 213

70. IEA 2009 South Africa Biofuels IEA Taskgroup 39 Progress Report, 2009.
71. Johnston, M. and Holloway, T. A global comparison of national biodiesel
production potentials, Environmental Science and Technology, 41 (23): 7967–
7973, 2007.
72. Antwi, E., Bensah, E.C., Quansah, D.A., Arthur, R. and Ahiekpor, J. Ghana’s
biofuels policy: challenges and the way forward. Int. J. Energy Environ., 1:
805–814, 2010.
73. Balat, M. and Balat, H. A critical review of biodiesel as a vehicular fuel,
Energy Conversion and Management, 49 (10): 2727–2741, 2008.
74. Kiss, A.A., Dimian, A.C. and Rothenberg, G. Solid acid catalysts for biodiesel
production-towards sustainable energy, Advanced Synthesis and Catalysis,
348 (1-2): 75–81, 2006.
75. Outlook, B. E. BP Statistical Review of World Energy Statistical Review of
World, 2019. Retrieved from https://www.bp.com/en/global/corporate/
energy-economics/statistical-review-of-world-energy.html
7
A Global Scenario of Sustainable
Technologies and Progress in
a Biodiesel Production
M. B. Kumbhar1, P. E. Lokhande1,2*, U. S. Chavan2 and V.G. Salunkhe3

Department of Mechanical Engineering, Sinhgad Institute of Technology,


1

Lonavala, India
2
Department of Mechanical Engineering, Vishwakarma Institute of Technology,
Pune, India
3
Department of Mechanical Engineering, Annasaheb Dange College of Engg.
and Technology, Ashta, India

Abstract
Biodiesel is an alternative fuel source to replace the non-renewable crude oil and
their possible environmental pollution.The production of a biodiesel has currently
many challenges such as performance-based efficiency, environmental friendly
cycle, next generation–based, flexible and ease production technique, sustainabil-
ity, and commercialization. Feedstock has played crucial role to run the biodiesel
production to sustain the global market. This chapter gives latest development
as regards the feedstock research review that gives remarkable progressive fea-
ture of an existing feedstock which survives a long-term consistent availability
for biodiesel commercialization. Feedstock sources and combustion parameter are
strongly believed to improve the production method that gives exact ways to apply
the knowledge to produce a biodiesel from a catalyst-assisted biodiesel plant or a
biorefinery. The current scenario of technology and research areas in this alterna-
tive fuel are highlighted.

Keywords: Feedstock sources, combustion performance, biodiesel


commercialization

*Corresponding author: prasadlokhande2007@gmail.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (215–240) © 2021 Scrivener Publishing LLC

215
216 Biodiesel Technology and Applications

7.1 Introduction
The natural renewable source as an alternative fuel among these some
well-known sources is a biodiesel instead of current crude oil sources. It
is available in unlimited large amount in the world. This sustainable and
strong potential source reduces petroleum holding and pollutant emission
contributes to global warming [1]. There is wide application such as auto-
mobile, industrial, heating oil, and aviation. The use of biofuels in the avia-
tion sector can eliminate greenhouse gas emissions between 50% and 90%
compared to existing kerosene. Advanced biofuels have environmental
footprints associated with greenhouse gas emission savings between 80%
and 90% compared with fossil references. Many challenges stand to zero
and reverse the harm done to Mother Nature by mankind, and still, society
cannot simply end its energy demand. Therefore, it must continue to utilize
its current energy sources more efficiently while investing in renewable
sources. Numerous replacements like nuclear, solar, biofuel hydro, wind,
biofuel, and biodiesel are recommended, but still, they are in the stage of
research and development stage [2]. Several researchers are focus to syn-
thesize with low cost and in emerging commercial technologies. Biodiesel
consumes oxygen of 11% by weight which satisfies and fulfills the com-
bustion and thus improves the performance characteristics of engine [3].
Although the biodiesel industry has sustainable with tremendous global
growth rate, raw material or feedstock supplies have strongly recom-
mended as a natural brake and created a strain on margins for biodiesel
producers. Therefore, commercial technologies for biodiesel production
from biodiesel feedstock receive tremendous interest all over the globe.
There have many challenges as regards biodiesel commercial technologies
and environmental concerns that are enlisted as follows.

• Researchers focus to produce high free fatty acid (FFA) for


selected level of biodiesel yield as well as higher flash point,
remarkable biodegradability, excellent lubricity, with high
rate of combustion efficiency associated with petro-diesel
properties [1].
• Biodiesel production from alternative resources poses many
limitations. Unsteady crude oil prices, depletion of fossil-fuel
resources, and ecological measure are the main limitations
for searching an innovative fuel which satisfies and fulfills
the parameters such as eco-friendly, inexpensive, extensively
accessible, and technically suitable [2].
Progress in Biodiesel Production 217

• To ensure stability of biodiesel, it is degraded by numerous


influences like (i) long-term storage conditions, (ii) auto-
oxidation, (iii) thermal decomposition, (iv) biodegradation
with microbial progress, (v) water absorption, (vi) contami-
nation of metal, and (vii) nonappearance and occurrence of
additives. It is significant to advance such additives, which
may provide all these benefits to modify and utilize the
application of biodiesel for commercialization [3].
• A major common feedstock currently utilized in biodiesel
production is vegetable oils. Although biofuel from vegeta-
ble oils is debatable to a straight utilization in a quantity, of
times, it has called criminal against humanity due to social
activities in biofuels, and biodiesel is foremost the cause
behind global growth rate in food market supply. This con-
troversial factor of “food vs. fuel” war does not fulfill circum-
vented if cherished cultivated land is consumed for making
of non-edible oil crops [4, 5].

The mandatory and revolutionized elements are energy, soil, water,


nutrient, agrochemical board and crop safety, landscape and biodiversity
safeguarding, and energy harvesting, which are good agreement to green
sustainable ecological practices that would report biodiesel invention and
commercialization. This methodology integrates technology with eco-
nomic, social, and environmental disputes that will upsurge biodiesel profit
and improve the biorefineries that are accomplished of sustainable bio-
diesel production and additional remarkable chemicals. As early it is seen,
government strategies will be the powerful energy for advance growth rate
in biodiesel production. These strategies must embrace the great signifi-
cance social attentions like employments and favorable working environ-
ment, creating awareness about environments among peoples, etc. Later,
strong support between the governments and stakeholders is required to
grow and spread over effectively sustainability criteria in worldwide [6].
Researchers collaborating with the investing a greater role of conducting
the initial waste, i.e., unwanted resources to the commercial equality that
creates the complete procedure toward fulfilling a green environment.
Continuously usage of waste feedstock is anticipated in centuries to emerge
as to withstand the production of biodiesel firm and additionally extend
the gap among biodiesel cut point and the trade price. The researcher is
always learning, sharing, and connecting to fulfill the gap between loca-
tions and production technology and commercializing the source of a bio-
fuel for biodiesel production technology and commercializing the source
218 Biodiesel Technology and Applications

of a biofuel. Also, the European Union communal has continuously in


research for oil crops additional for petro-based chemicals eliminate the
dependence on imports [7]. Hence, the former opportunities left out to
the corporate economically though evading its provocative structures
applicable waste-oriented or non-edible oil feedstock increase on non-
agricultural space and efficient biodiesel production from these diverse
feedstock categories.
For 2018, the world’s biodiesel production is overviewed by the United
States production of 1,190.2 thousand barrels/day, followed by Brazil out-
put of 693.2 thousand barrels/day. Third, Germany has produced 75.8
thousand barrels/day, fourth, Argentina production of 70.6 thousand
barrels/day, and fifth, China that reported a production of 68 thousand
barrels/day [5]. Biodiesel has been one part of the social discussion and
critically innovated by environmental organization for a biodiesel produc-
tion innovation and technology. It globally support policies that capture
biodiesel from residue and waste materials and the self-introduction to the
market, i.e., synthetic fuels from renewable power [8]. These synthetic fuels
run the automobiles, aviation, and industrial sectors.

7.2 Current Status of Feedstock for Biodiesel


Production Technology
The biodiesel market growth rate is built on the availability of feedstock
sources for biodiesel production technology. A long-term strategy (bio-
fuels after 2050) can only be based on “high yield per hectare”. A primary
source to run the biodiesel plant or refineries is biodiesel feedstock.
Potential biodiesel feedstock and technology resources of the future are
currently leading for higher effectiveness, commercialization, and reducing
economic impediments. Therefore, biodiesel feedstock has some research
circumstances to sustain the biodiesel production, reduces pressure on
cropland, portfolio of feedstocks, green house emission (GHG) reduc-
tions, integrated models across agricultural, forestry, and energy markets,
and profitable for farmers and forest managers to produce data for recent
­generation-wise feedstocks.
Also, conversion efficiency is an important parameter based on the cost
of biofuels production. Analysis of literature focused at modifying methods
used in feedstock production and conversion could modify feedstock avail-
ability and extract the quantity needed to achieve biofuels target, lowering
feedstock cost and making more utilization for other uses. A conceptual
Progress in Biodiesel Production 219

framework of literature describes the relationship between consciousness


feedstocks for biofuels and the substantial market for each and every feed-
stock, including how higher yields for respective feedstocks (e.g., resulting
from investments in research) affect feedstock and biofuel economy. The
following focused current status of feedstock analysis reviews emerge as
priorities for future investigating efforts to sustain the biodiesel from these
feedstock or complex integration of a feedstock.
The most significant principle for biodiesel expansion is to discover
low price feedstock with higher amount of oil content. Transesterification
method for production of biodiesel from goat waste with substantial FFAs
is investigate with infrared radiation assisted reactor [1]. For biodiesel pro-
duction, investigators have considered several categories of homogeneous
and heterogeneous catalyzed transesterification reaction. These methods
can boost the quality of Fatty Acid Methyl Ester (FAME) implementation
for diesel engines without any variation [2]. The integrated production
plant of first-generation soybean biofuel and sugarcane ethanol concludes
negligible influences to the GHG emission drop and greater prices owing
to the low sulfur diesel. A biorefinery design gives region-wise biodiesel
production that provides us a lot of local opportunities [9]. Next rich
source of oil content in rubber seed with combination of 40%–48% shell
and 52%–60% kernel is consumed as a feedstock for biodiesel as per their
sustainability. Particle size and extraction time optimization parameters
are analyzed to production of a higher yield biodiesel. The characteristic
physico-chemical oil proves that it can be sustained as a possible cause for
biodiesel production [10]. Second-generation source as a Moringa oleif-
era oil and leaf as a valuable contribution causes for the biodiesel produc-
tion as well as antioxidant biodiesels additives. This Moringa oleifera is
a sustainable and natural source as an antioxidant as compared to tert-­
butyl-hydroquinone (TBH) for biodiesel to provide low oxidation stability
and high induction period (IP) value [11]. The crude Moringa oleifera oil
considers as a higher FFA content and use for human utilization. As com-
pared to diesel fuel, cetane number of Moringa oil is approximately 67, and
among these resources, it is a maximum stated value for a biodiesel fuel.
In Moringa oleifera, oil contains antioxidant and low polyunsaturated fatty
acid which provides higher oxidation stability. The oil properties like oxi-
dative stability, kinematic viscosity, density, acid value, and iodine value of
synthesized Moringa oleifera methyl esters are in good settlement with the
EN and ASTM standard. Higher flash point marks Moringa oleifera methyl
esters to be intensely recommended as a non-flammable. Hence, Moringa
oleifera oil owns a major impending source to be used for production of
biodiesel as a feedstock. In India, it is stated that Moringa oleifera oil has
220 Biodiesel Technology and Applications

major impending for production of biodiesel and shows that high stability
to oxidative rancidity [12].
Sludge palm oil (SPO) is considered as a cheap substrate edible oil
(1 USD/Ton) for production of biodiesel. Conversion of SPO to biofuel is
an effective means of providing a cheap biodiesel production and waste
utilization. Generally, biodiesel is produced via a transesterification reac-
tion with the help of homogenous alkaline catalyst. This method has wide
range advantages such as high conversion, short reaction time, and low
catalyst use [15]. Palm oil mill effluent (POME) is a sustainable feedstock
for conversion biodiesel worthwhile of high volumes. Next criteria is its
huge quantities of freely available fatty acids (FFA), and water in POME
prevents trans/esterification reaction and harmfully disturbs present tech-
nology engaged in biodiesel production [16].
To boost the properties for sustaining the storage stability, research is
focused on availability of variable complex feedstock. The complex mix-
ing of Karanja and Coconut biofuel and their respective blends conserve
long-term storage. Also, good agreement with stability of biofuel upon
sustaining the storage, sunlight, and non-contact with air is found to be
the superior storage condition studied [17]. Rapeseed and soybean oils
have been considered as an alternative energy source in Europe and North
America, and the utilization of palm and jatropha oils for production of
biodiesel contributes to a new energy source in West Africa and South-East
Asia [18]. The second biggest traded commodity global is coffee, i.e., it is
a sustainable source for biodiesel. Instead of solvent extraction, esterifica-
tion, and transesterification (three-step process), direct transesterification
to biodiesel production from spent coffee grounds (SCGs) with favorable
reaction condition [19, 20].
Third-generation micro-algae are now promising the major resource of
production of biofuel in the global world. They are assessed as the non-
competitive, safer, and drastically growing organisms among those that
could be utilized for production of biodiesel. The potential to grow with-
out much care on waste nutrients and the better renewable source of bio-
diesel production as compared to other sources can affect food problems
as they are majorly considered as plants which are used for food [21]. A
primary standard process trained in microalgae culture, harvesting, and
processing, the existing study boons a summary of recent development
in area of algal biofuel. Impending resource of biofuels which is majorly
studied is algae. Algal fuels are striking to many investigators in the areas
of bioenergy and biodiesel. Investigation of dynamic algae species and its
revolution into alternative fuels and bioproducts is attractive in public and
private sector [22]. Lipid extracted from microorganisms, including algae,
Progress in Biodiesel Production 221

yeasts, fungi, and bacteria, are defined as single-cell oil (SCO) which is the
best capable feedstock for biodiesel production due to their homogene-
ity in structure of fatty acid which is good agreement to vegetable oils. A
microbial base for lipid production has numerous benefits such as small
growth duration with advanced lipid productivity and poor in climate
change. The reports on biodiesel were evaluated from different oleaginous
yeast integrated with low-cost raw material for large-scale commercial
production so as to progress a worksheet for production of biodiesel. The
concluded constraints help to analyze fuel blends which diminish the risk
of noncompliance with the technical standard [23]. Advanced biodiesel
and biogas can be extracted from a numerous of feedstock; Norwegian
companies majorly produced bioethanol and biodiesel from lignocellu-
lose biomass [24]. Edible mustard oil from Cruciferae family is consid-
ered as oil and spice plant which is in lengthwise 0.2–1.5 m and white or
yellow flowering. Entire world, approximately, 10 types of mustard were
obtained from white and black species. It has showed that 96.695% of bio-
diesel yield was found from yellow mustard [25]. Soybean biodiesel and
linseed-derived blends determined an inclination of minimal emission
that constitutes nitrogen oxides as related to maximum diesel engine loads
[26]. A sustainable city and society waste cooking oil (WCO) is comfort-
ably accessible from household kitchens, restaurants, and cafeterias. The
pretreated acid WCO was laid open to transesterification with base cata-
lyzed to production of biodiesel. In total, 94% of a higher yield was gained
with 1:3 methanol to oil ratio, reaction temperature 60°C, and 1% cata-
lyst dose [27]. Sunflower and rapeseed oils are majorly utilized in Europe,
canola oil in Canada, soybean oil in the United States, and palm oil avail-
able in tropical countries. It is significant to footprint that the price of bio-
diesel production from the first-generation feedstock is 30% greater than
of petroleum-based diesel. Feedstock for second-generation biodiesel such
as lowcost non-edible sources, e.g., from karanja, pongamia, jatropha, as
well as microalgae, alters sources of microorganisms. Waste frying oil and
waste from slaughterhouses and rendering industry can be classified under
this category. Microalgae are considered as “micro sunlight-operated bio-
chemical company”. They are obtained as a third-generation potential
source of biodiesel production [14]. The seeds of different plants indig-
enous of tropical scale could be used for production of biodiesel, and it
determines the suitable properties that they show. Such plants could be
cultured in several tropical ecoregions with marginal costs, to provide bio-
diesel instead of carbon-based fuels [28].
Nanoadditive is a modern nanocommercial application on microalgae
biofuel enrichment and it has been classified according to different phases
222 Biodiesel Technology and Applications

from waste material production to end product insinuations: nanoadditives


for microalgae cultivation for the resolution of microalgae biomass trans-
formation into biofuels and nanoadditives for microalgae-biodiesel appli-
cations. Scientists are still discovering an invention which will improve the
higher yield microalgae biofuel from the beginning step to end product
and prime concerns toward a cost-effective fuel results [29]. The ligno-
cellulose ethanol market has been slower progress rate due to concerns of
intensive capital costs, high technological risk, and the low oil prices that
highlighted in poor economics for the biorefineries. Food and Agriculture
Society of the United Nations (UN) (2015) has calculated a total capacity
of 1703 million liters for second-generation ethanol by 2024, majorly in
USA and Europe (UNCTAD, 2015). Second-generation biodiesel (2G) that
consists of herbaceous crops, lignocellulose materials, hardwood, and soft-
wood is the intensive feedstocks used for the production of liquid ethanol
[30].
Koelreuteria paniculata (KP) that belongs to a family of Sapindaceae
is a recent second-generation non-edible seed oil which will enhance
production of biodiesel. Approximately, 1-hectare zone will grow up to
115,000-kg seeds and the productivity of biodiesel is near about 30,000
kg per hectare zone. KP species firstly belong to countries such as Japan,
China, and Korea. The concluded optimization transesterification param-
eters are catalyst concentration (0.32 g wt.%), oil-to-methanol ratio (6:1),
temperature (65°C), stirring rate (700 rpm), and time (80 min), and the
highest biodiesel yield of 95.2% was determined, which is satisfactory for
marketable technology. The physical-chemistry properties of the biofuel,
i.e., flash point, cloud point, pour point, and density, are acceptable to the
ASTM (D6751) and (EN14214) standards; biofuel good values are closer
to mineral diesel, to improve the performance of an engine application.
KP seeds are analyzed to be substitute and alternative source of biofuel for
generation of potential technology, which might be a crucial auxiliary for
petroleum and diesel [31].

7.3 Scenario of Biodiesel in Combustion Engine


Biodiesel revolution is helping several countries to overcome their depen-
dence on diesel. The adoption of biodiesel worldwide economy within the
shipping sector has fetched with a reliable and smooth fuel supply. It can
be used without any variation or minute moderation in diesel engines. It is
biodegradable and nontoxic and it produces fewer pollutants when burnt
completely. Biodiesel having acceptable source such as 11% oxygen by
Progress in Biodiesel Production 223

weight can ensure total combustion and thereby results in upsurges in the
engine performance as well as it diminishes the exhausted gas emission as
associated to that of diesel.
Nevertheless, the certain limitations for marketable application of bio-
diesel such as higher viscosity cause injector coking, fuel filter plugging,
moving parts sticking, poor atomization, and fewer fuel injector precise
operations.
The real fact is that biodiesel is auto-oxidative and can easily affect its
other fuel properties such as flash point, cetane number, viscosity, and den-
sity [3, 32]. The analysis result shows that CO, HC, smoke, and CO2 reduc-
tion for biodiesel blend and NOx upsurges related to diesel are utilized in
the same engine. However, some analysis shows that NOx declines with
biodiesel. It exhibits feedstock quality, engine type, and running situations
that vary from one to another; later, the distinction of performance and
emission parameter is investigated [33].
Jatropha oil, Moringa oleifera, KP, and microalgae are foremost aspi-
rants for the commercialization of biodiesel production instead of diesel
fuel in combustion engines. The effective method to eliminate emissions
and refining engine performance is by supplementing nanoadditives with
emulsified fuels [34].

7.4 Biodiesel Production Technologies


There are also some little worth mentioning challenges such as environ-
mental parameters, feedstock sources, combustion-based parameter, and
convenient technology adopted with biodiesel production. Among these,
frequently stated are the rate of feedstock and the suitable optimization
technology for effective biodiesel production from several feedstock cat-
egories. The technologies, which enable us to utilize biofuel and fat feed-
stock types and to utilize fuel in diesel engines, are always defined as
straight utilizing or blending, microemulsion, pyrolysis, and transesteri-
fication. Transesterification is presently stated by several investigators as
utmost acceptable due to its improved quality of fuel produced [32]. All
these technologies are briefly explained with their applicability, benefits,
and limitation as follows.

7.4.1 Direct Blending


Direct utilization of vegetable oils is normally considered as non-
acceptable and unrealistic for direct and indirect diesel engines. The FFA
224 Biodiesel Technology and Applications

content, composition, high viscosity, and gum development due to oxida-


tion and polymerization at the time of storage and combustion, lubricat-
ing oil thickening, and carbon deposits are obvious problems. Inadequate
combustion and oil deterioration are the two simple limitation related to
the straight utilization of vegetable oils as biofuels [32].

7.4.2 Pyrolysis
Pyrolysis is considered as a chemical modification affected by the applica-
tion of heat in the occurrence of a catalyst and deficiency of air or oxygen,
in which result shows the splitting of bonds and establishment of several
tiny scale molecules. Three types of methodology for pyrolysis technique
are as follows: (1) conventional pyrolysis exhibits with temperature range of
550–900 K, (2) fast pyrolysis exhibits with temperature range of 850–1250
K, and (3) flash pyrolysis exhibits with temperature range of 1,050–1,300
K. Currently, acceptable number of pyrolysis reactors is utilized with 50%
to 70% yield range. These are categorized as a (1) circulating fluidized bed,
(2) bubbling fluidized bed, (3) ablative, (4) rotating cone, and (5) focused
solar reactor [35].
There are two in situ and ex situ onsite methods to produce a biodiesel
from pyrolysis technique (Figure 7.1). In situ catalytic pyrolysis, the ability
to splitting the reactions to cut downcast the substantial molecules exist-
ing in the pyrolysis goods foremost to the establishment of lighter and
min viscous bio-fuel. In situ catalytic process of biofuel vapors is also per-
formed in the same pyrolysis reactor, e.g., a fluidized bed reactor. Ex situ

In-situ catalytic pyrolysis Ex-situ catalytic pyrolysis

Biomass Gas Biomass Gas

Catalyst
Catalyst Reactor Reactor Packed
Bed

Char/Catalyst Bio-Oil Char Bio-Oil

Figure 7.1 Online two-way catalytic pyrolysis reaction process for conversion of biodiesel
with the help of zeolite catalysts [36].
Progress in Biodiesel Production 225

catalytic process of biofuel can be implemented at the downstream of the


pyrolysis reactor in a distinct reactor such as a packed bed reactor sys-
tem. Combination of pyrolysis with their appropriate catalyst is a com-
mercialized technique for cost competitiveness of production of biofuel
[36]. Innovating these pyrolysis techniques for commercialization has
many challenges such as feasibility, potential catalyst, reduction of unde-
sired compounds, imprecise product quality, uncontrolled reaction, and
post-treatment of end products [37].

7.4.3 Microemulsification
Microemulsification method is also identified as a viscosity reduction. It is
carried out by blending of two immiscible liquids. Microemulsions exhibit
colloidal distribution of optically isotropic fuel which occurs as micro-
structures under symmetry. Surfactants are used for processing of micro-
emulsions that have substantial role to aquatic and terrestrial ecotoxicity
[35]. Microemulsification technique formation, stability, and phase behav-
ior have been proposed over the years. With the help of these theories, the
biodiesel properties for sustainability are improved.
Microemulsions steady blend of oil are consist of, an anisotropic, trans-
parent and thermodynamically. This steady blend of oil exists in water

Water/ Vegetable
Ethanol Oil
(Polar) (Non-Polar)

W/O O/W W/O/W O/W/O

I II

Surfactant/Co-surfactant

W/O: Water in Oil Emulsion


O/W: Oil in Water Emulsion
W/O/W: Water in Oil in Water Emulsion
IV O/W/O: Oil in Water in Oil Emulsion
MHBF MHBF: Microemulsion based hybrid Biofuel

Figure 7.2 W/O/W and O/W/O emulsion P. A. Winsor I-II and IV type emulsion [37].
226 Biodiesel Technology and Applications

(O/W) or water in oil (W/O) stabilized with the assistance of surface-


active agents (surfactants and/or co-surfactants). Figure 7.2 shows the
Winsor I-II (W/O, O/W), and IV types of microemulsions and W/O/W
and O/W/O type microemulsions; the applications of microemulsions
are medicinal/food/paint industries and biodiesel. The sustainability
analysis of microemulsions, exergy analysis, and its lengthy occurrence
effects upon engine are also a major factor considered for forthcoming
research work [37].

7.4.4 Transesterification
Transesterification reactions are always catalyzed by the mixture of
an acid or base catalyst. This method is also widely use to convert fats (tri-
glycerides) into production of biodiesel. Transesterification extracted vege-
table oil (biodiesel) was utilized to power heavy-duty vehicles before World
War II in South Africa.
Therefore, transesterification is an ease process to production of bio-
diesel extract from fat and oil feedstock sources. Transesterification is a
reversible method, which basically exhibits principally by mixing the reac-
tants with the help of heat and/or pressure. Transesterification reaction for
effective production of biodiesel is designated by simple and effective part-
ing of an ester and glycerol layer after the reaction time. Production of bio-
diesel via a transesterification reaction by mixture of homogenous alkaline
catalyst. This method has several benefits such as short reaction time, high

Biodiesel Methods

Blending Pyrolysis Micro-emulsion Transesterification

• Double layer formation • Side product formation • Significant injector • Cost effective method and gives
• Gum Formation • Lack of an effective needle sticking high conversion efficiency
• Carbon Deposits catalyst • Incomplete combustion • Improve the fuel characteristics
• Increase smoke, CO and • Expensive setup • Carbon deposits • Glycerol produced as by-product
HC emission • Eliminates environmental • Increase the lubricating can be used for medicinal pupose
• Ring sticking benefits oil viscosity • Biodiesel produced by this method
is miscible in any proportional with
the mineral diesel
• Can be employed to a variety of
feedstocks

Figure 7.3 Biodiesel production methods.


Progress in Biodiesel Production 227

conversion, and low catalyst use [15]. Some of the advantages over other
biodiesel production methods are shown as Figure 7.3.

7.5 Microwave-Mediated Transesterification


Microwave irradiation process is a substitute to the traditional heating
system with minimum reaction time [13]. It has been widely acceptable for
a variation of applications such as organic synthesis reaction. The chemi-
cal reactions are augmented because of specific absorption of microwave
energy by polar molecules. Due to the blend of methanol, vegetable oil,
and potassium hydroxide that encloses both polar and ionic components,
rapid heating occurs upon microwave irradiation. The energy interrelates
regards to the reactant on a molecular level, actual effective heating is
achieved [38].
The uninterrupted microwave reactor constructed in this investigation
can generate about 19,000 L of biodiesel per day with approximate yield of
90% that is considered, as shown in Figure 7.4. Microwave irradiation pro-
cess and conducting reaction more capably, with less parting and reaction
times, eliminate the numbers of by-products and drop energy use. It shows
that microwave-mediated transesterification results are in close agreement

Line-Out

Product
reactor
Line•In
Biodiesel

Pump Pump

Glycerol

Jatropha•Oil CH3OH

Figure 7.4 Continuous microwave reactor [38].


228 Biodiesel Technology and Applications

Table 7.1 Biodiesel production from various feedstocks [13, 14].


Second- Fourth-
First-generation generation Third-generation generation
biofuel biofuel biofuel biofuel
Sources Sources Sources Sources
Edible oil: Non-edible oil: Oleaginous microbes Genetically
Soybean, Neem, Jatropha, microalgae, optimized
Coconut, Moringa, Algae, bacterium, yeast and feedstocks
Sunflower, Orange, Linseed, fungi algae to biofuels
Palestine, Mahua, Karanja
Corn, Rice or Pongamia
bran, Olive, Jojoba, Kusum,
Sesame seed, Cottonseed,
Palm, Pistachio, Rubber seed,
Rapeseed, Deccan hemp,
Peanut oil, Sea mango,
Safflower oil Waste cooking
oil, and Animal
source
Benefits Benefits Benefits Benefits
Renewable Renewable sources, Renewable sources, Renewable
sources, Environmental Environmental friendly, sources,
Environmental friendly, High lipid production, Capture large
friendly, Effective land Do not compete with amounts of
Easy conversion utilization, land utilization and carbon, with
into biofuel Do not compete food, genomically
with food crops High growth rate synthesized
tendencies microbes,
High yield
and lipid
containing
algae,
Economical
running
production
Limitations Limitations Limitations Limitations
Net energy losses, High cost, Insufficient biomass Advanced
Compete with Inefficiency, production to synthetic
food crops Unsustainability, commercialization, biology,
(food energy Land and water use Higher initial High initial
crime), competition investment cost for investment
Raising cost production
of food due
to food
competition,
Land scarcity
Progress in Biodiesel Production 229

with appropriate production process and cost competitive to make bio-


diesel in global market (Table 7.1) [38].

7.6 Ultrasound-Mediated Transesterification


Ultrasonication-mediated transesterification is the process to convert
mechanical energy for blending and to carry out a reaction. It consumes
minimum energy consumption and efficiency of cavitations which is
reliant on transformation to FAME (99%) within minute and irradiation
frequency [13]. The ultrasonication process has major benefits that are
reduced process time, positive effect on transesterification process, and
saved energy in the biodiesel production [39].
The benefit of microwave and ultrasonication technologies is to syn-
thesize the reaction, and it directly drops costs in the biodiesel production

Table 7.2 Difference between conventional, microwave, and ultrasonic heating [40].
Conventional heating Microwave heating Ultrasonic heating
Thermal gradient Inverse thermal Limited thermal
(outside to inside) gradient (inside to gradient due to
outside) mixing
Conduction and Molecular level hot Microbubble formation
convection currents spots and collapse
(compression and
rarefaction cycles)
Longer processing Very short and instant Very short reaction
times heating times
No or low solvent No or low solvent Solvent saving possible
saving reactions possible
Product quality and Higher product quality Same as conventional
quantity can be and quantity possible heating
affected
Separation times are Very short separation Less than conventional
longer times heating
High energy Moderate to low Moderate to low
consumption consumption consumption
Simple process Very simple process Moderate complexity
configuration
230 Biodiesel Technology and Applications

process due to clear FAME [13]. Ultrasonic and microwaves together


improve extraction of oils/lipids by diffusive and disruptive mechanisms
and simultaneously transesterify the lipids/oils due to thermal/specific
non-thermal effects at molecular levels and increased mass/heat trans-
fer phenomena. Microwave and ultrasound-mediated transesterification
methods have several power steps at several frequencies [39].
Among these two technologies, microwave and ultrasonication technol-
ogies are differentiated on the basis of chemical and physical parameters as
shown in Table 7.2. The biodiesel reaction proficiency can be modified by
integrating intensification process effects and the green chemistry princi-
ples. Green chemistry is defined as chemical products and reduces the use
and generation of hazardous substances methods. Microwave-assisted bio-
fuel intermixes the reaction efficiency due to maximum product recovery
and minimum by-product formation and eliminates energy use [40]. The
difference between these two processes over conventional method is given
in Table 7.2.

7.7 Catalysis in Biodiesel Production


Catalysis is the procedure of modifying the growth rate of a chemical reac-
tion by mixing a substance known as a catalyst, which is not expended in
the catalyzed reaction and can carry out to act repeatedly. That is why, only
very negligible amounts of catalyst are essential to change the reaction rate
in furthermost processes.
Catalytic reactions are selected in environmentally friendly green chem-
istry due to the eliminated amount of waste generated. Few of the reported
chemical catalysis utilized for biodiesel production are mentioned in the
following.

7.7.1 Homogeneous Catalysts


A homogeneous catalyst is defined as the molecules that are isolated in
the similar phase (usually gaseous or liquid) as the reactant’s molecules.
There are two types of homogenous catalysis that are acid and base catal-
ysis. The evergreen first method from homogenous catalyst is used as an
acid catalyzed transesterification to biodiesel production (ethyl ester) from
palm oil utilizing ethanol and sulfuric acid. Traditionally, homogeneous
chemical catalysts have several benefits: effortless optimization of activ-
ity including high selectivity, reaction rate, turnover frequency, and cheap
and readily available. NaOH, KOH, CH3ONa, and CH3OK (potassium
Progress in Biodiesel Production 231

methoxide): these are strong alkali catalysis that is utilized for production
of biodiesel. Therefore, alkali metal oxides are established to be extraener-
getic than hydroxides. Most regularly used organic sulfonic acid, sulfuric
acid, sulfonic acid, hydrochloric acid, and ferric sulfate are utilized for
homogenous catalytic transesterification process. There are some limita-
tions such as water content and minimum tolerable FFA, and the refine-
ment procedure is complex, used high energy which affects a considerable
upsurge in apparatus investment costs and safety and maintenance prob-
lem. Also, this method has FFA content in the feedstock and is highly sen-
sitive to water. Two-step transesterification homogenous catalysis method
is enhanced than one-step homogenous catalysis reaction since it needs a
minor alcohol, amount of catalyst, and reaction temperature but modified
an advanced translation rate [2, 14, 32, 42].

7.7.2 Heterogeneous Catalysts


A heterogeneous catalyst, where the molecules are not in the similar phase
as the reactants, which are naturally gases or liquids that are adsorbed onto
the surface of the solid catalyst. Heterogeneous catalysts are modifying
the transesterification method by reducing the large processing costs in
a homogeneous catalysis reaction and eliminate the growth of pollutants.
Heterogeneous catalysts enhance reusability, smooth recovery, and cost
reduction methods. These catalysts increase high FFA and moisture con-
tent. Low-cost and effective heterogeneous catalysts optimize the overall
production cost of biodiesel. Heterogeneous catalysts are acceptable inte-
gral in specific harsh situation like high pressure and temperature. These
catalysts are simple to regain from the reaction mixture, are enhance to
withstanding aqueous treatment methods, and modify the parameter
such as high activity, selectivity, and longer catalyst lifetimes [2, 14, 32,
42].
In various chemical reactions, the general catalysts are used such as
alkali metal compounds, alkali earth metal oxides, mixed metal oxides,
transition metal oxides, ion exchange resins, supported on alumina or zeo-
lite, waste material-based heterogeneous catalyst, carbon-based heteroge-
neous catalyst, and enzyme-based heterogeneous [2, 14, 32, 42].
Alkaline earth metal oxides catalysts include BeO, BaO MgO, CaO,
SrO, and RaO. Solid-base catalysts used in transesterification include CaO,
MgO, SrO, K2CO3/Al2O3 KNO3/Al2O3, KF/Al2O3, Li/CaO, KF/ZnO, basic
hydrotalcite of Mg/Al, Li/Al, anion exchange resins, and base zeolites. MgO,
CaO, and TiO2 catalysts at temperatures up to 200°C have higher poten-
tial biodiesel production, and it can achieved 95% yield. Heterogeneous
232 Biodiesel Technology and Applications

solid-base catalysts are widely used for industrial application due to its
oversimplification of purification processes and production [2, 14, 32, 42].

7.7.3 Heterogeneous Nanocatalysts


Nanocatalysis is one of the most promising subfields introduced in nano-
science. Higher growth rate of chemical reactions is achieved by altering
its size, chemical composition, dimensionality, and morphology of the
reaction center and by changing the kinetics using nanopatterning of the
reaction. Heterogeneous nanocatalysts to commercialize catalyst with
outstanding high stability, high activity, selectivity, and with a high yield
product is promising by altering the elemental composition, surface func-
tionality, and/or the number of atoms in the substance [2, 14, 32, 42].

7.7.4 Supercritical Fluids


Vegetable oil converted into biodiesel is via a transesterification reaction,
where the triglyceride is transformed to the methyl ester and glycerol. This
is often completed using methanol and acid or caustic catalysts, and it can
be attained by supercritical methanol without a catalyst. The supercriti-
cal fluid (SCF) method offers a several benefits that are fast reaction, easy
separation of the products, water content of feedstocks, allowing a greater
range, and it is also environmentally benign. A recent technology of pro-
duction of biodiesel was synthesized by a noncatalytic supercritical process
with higher yield at 96% within 10 min. It is simple fluids to continuous
biodiesel production [2, 14, 32, 42]. Table 7.4 describes the transesterifica-
tion process characteristics based on basic, acid, enzymatic, and supercrit-
ical transesterification.

7.7.5 Biocatalysts
Biocatalysts can play better role in the limitations of the chemical catalytic
system, free lipase, immobilized lipase, and lipase immobilized on MNPs.
Biocatalysts are produced by animals, microorganisms, and plants; these
are consist of biocatalysts for biodiesel production [13]. Biodegradability,
biocompatibility, and environmental applicable enzymes are playing bet-
ter role than homogeneous catalysts. Lipases have also a potential catalytic
method and stability in nonaqueous media. Conventional lipase consists
of immobilized lipase and lipase immobilized on MNPs. The immobilized
lipases should be available in batch and fixed bed processes with appropriate
Progress in Biodiesel Production 233

Table 7.3 Comparison between homogeneous, heterogeneous catalysis,


supercritical alcohol use, and basic enzyme catalysis in terms of advantages
and disadvantages.
Example of
catalyst Advantage Disadvantage
Homogenous
Basic (NaOH, Low cost, speeds Incompatible with free fatty
KOH) of high reaction, acid formation of soap,
advantageous polluted separation effects
operating condition,
high activity.
Acid (H2SO4) No soap formation, Corrosion, polluted separation
catalyze the effects, long reaction time,
esterification of high reaction temperature.
free fatty acid and
the transesfication
of glycerides
simultaneously.
Heterogeneous
Basic (CaO, Recyclable, non- Incompatible with free fatty
CaCO3, corrosive, easy acid formation of soaps,
KOH-Al2O3, separation, better incompatible with water-
Al2O3-KI, selectivity, long formation of free fatty
Zeolite catalyst life cycle. acids, high molar ratio
ETS-10) of alcohol/oil required,
high temperature and
high pressure, diffusion
limitation, higher costs.
Acid (ZnO-I2, No soap formation, Low acid site concentration,
ZrO2-SO4-2, catalyze the low catalytic activity,
TiO2-SO4-2, esterification of free diffusional limitations due
solid catalyst fatty acid, and the to low microposity, high
based on transesterification cost.
carbon or of glycerides
carbohydrates) simultaneously,
recyclable.
(Continued)
234 Biodiesel Technology and Applications

Table 7.3 Comparison between homogeneous, heterogeneous catalysis,


supercritical alcohol use, and basic enzyme catalysis in terms of advantages
and disadvantages. (Continued)
Example of
catalyst Advantage Disadvantage
Supercritical
Without catalyst Very low reaction Drastic operating condition,
time, no limitation (high temperature, high
to transfer due pressure, high molar ratios)
to uniqueness of high operating cost.
phase, no catalyst to
recycle.
Enzymatic
Candida No parasitic reaction, Risk of denaturation of the
antarctica, easy separation, no enzyme, very slow.
Pseudomonas pollutant.
fluorescens

supports (magnetic nanoparticles) as these stabilize immobilization with


regard to covalent linkage in the amino acid residue (Table 7.3) [42].

7.8 The Concept of Biorefinery


The novel concept of “multi feeds-multi products” in combined methods
will make production of biodiesel as feasible and practicable. The sustain-
able methods of biomass are scalable for economy-based bioproducts and
bioenergy for potential market. Biorefinery principles highlighted few
basic concepts to be fulfilled: minimal environmental impacts, non-critical
feedstock, regeneration of biomass and biodiversity, readiness variations in
feedstock quality and operational flexibility, quantity, in acceptable man-
ner to the biodiesel production, and the ability to utilize the multi-feed
multi-products method. The biorefinery method easily addresses the exu-
berance causes as well as eliminates the logistics cost and other savings in
term of energy use [7]. Biorefinery policies that boost the remarkable bio-
diesel regain as well as technology of spent biomass should be developed to
make the fulfilled, low-cost, and energy-positive process [39]. Biorefinery is
a system which integrates biomass conversion processes and equipment to
Table 7.4 Comparison of various transesterification processes [41].
Process Basic Acid Enzymatic Supercritical
characteristics transesterification transesterification transesterification transesterification
Yield alkyl’s esters High (>94%) High (>94%) High (>90%) High (>95%)
Reaction Moderate (50–70°C) High (60–120°C) Low (30–60°C) High (200–280°C)
temperature
Alcohol-oil ratio Excess alcohol molar High excess alcohol Alcohol High excess alcohol
ratio (6–12 : 1). molar ratio stiochiometrically molar ratio
(10–166 : 1). ratio (3–4 : 1). (methanol : brute
oil) (40 : 1).
Reaction time Fat (4 h) Short (12 h) Short (28 h) Very short (7–15 h)
Catalyst Cheaper Cheaper Expensive Without catalyst
Energy High High Moderate Low
consumption
Product recovery Difficult Easy Easy Very easier
Environmental Using high quantity of The process is Ecological, no step Using high quantity
impacts water to wash corrosive; the acid washing with of water to wash
Progress in Biodiesel Production

catalyst is polluting. necessary.


235
236 Biodiesel Technology and Applications

produce biofuels, power, and value-added chemicals from biomass. The


biorefinery approach is related to current petroleum refinery, and it will
exhibit products from petroleum and multiple fuels.

7.9 Summary and Outlook


In 2020, there is major controversy as regards electric vehicles and biofuel
sustainability. Yet, there are much more limitations to run the entire vehi-
cles and system on electric. Nature is always surviving the energy sources
with their optimum utilization. If this attention is right, then yes there is
huge demand for biofuel production in coming decades. There is need to
commercialize the biodiesel plant or refineries widely.
The 2018 biodiesel production in a global world report gives progressive
future to commercialize the plant for production of biodiesel in a com-
ing decade. The key challenges for biodiesel production are high FFA with
desired level of yield, stability, optimized and flexible production, commer-
cialization of feedstock, and environmental friendly cycle. The worldwide
region where huge sources of biodiesel feedstock. The only limitation is to
construct a biodiesel plant is to complex integration or combine the vari-
ous feedstock sources of their region-wise and availability for production
of a biodiesel. It will solve major obstacles to resolve the commercialization
of a biodiesel plant with effective production for biodiesel economy.
There is generation-wise report for biodiesel feedstock that has been
reviewed. Feedstock optimization with their location are sustainable for
biodiesel refinery. Among the all resources, the major productions are that
is going on to produce a biofuel are soybean, Moringa oleifera, jatropha,
karanja, KP, and genetically optimized microalgae for feedstock. Additives
plays major role to boost the efficiency of a biofuel production for a commer-
cialization. Moringa oleifera is a strong potential to produce complex bio-
diesel feedstock with others resources because of its antioxidant properties.
Transesterification by heterogeneous nanocatalysts or biocatalyst is a
potential technique to carry out biofuel from biodiesel plant as per the
requirements. Catalytic pyrolysis and microwave-mediated transesterifi-
cation have strong potential to commercialize the biodiesel plant for bio-
diesel production. Biorefinery is a commercial process for biodiesel and
biogas production systems in order to demonstrate higher valuable prod-
ucts that could additionally modify economically and environmentally
sustainability of these biomethods.
The utilization of renewable sources in few countries has an autho-
rized funding. For the transport sector, accomplished Renewable Energy
Progress in Biodiesel Production 237

Directive (2009/28/EC) is targeting 20% share of alternative energy con-


sumption in the year of 2020, European Policy Framework with a sub-
target of 10% renewables. The collective effort and commitment of research
survey as regards feedstocks and commercialization of technology around
the globe toward sustainable energy are expressed in terms of accelerating
the biofuel economy.

7.10 Conclusion
According to biodiesel production technology world, the latest develop-
ments in biodiesel are increasing day by day, and the current status of tech-
nology is discussed. The existing edible feedstocks have some limitation as
regards crime against humanity. Therefore, major role of feedstock sources
is to widely available in rich energy of non-edible sources for commer-
cializing production of biodiesel. Among these, well-known high FFA for
desired level of yield and stability of a biodiesel and sustainable advanced
generation-based feedstock fulfill the criteria for a biodiesel production.
As per availability of a number of multiple feedstock grow up the plant or
a biorefinery for biodiesel production. Catalytic pyrolysis and microwave-
mediated transesterification develops desirable characterized biodiesel for
revolutionize the commercialization market. The summarized extracted
results shows that the evolutionary future to sustain the biodiesel for future
of a fuel to run the global world.

References
1. Chakraborty, R., and Sahu, H., Intensification of Biodiesel Production from
Waste Goat Tallow Using Infrared Radiation : Process Evaluation through
Response Surface Methodology and Artificial Neural Network, Appl. Energy.,
2013.
2. Talebian-kiakalaieh, A., Aishah, N., Amin, S., and Mazaheri, H., A Review
on Novel Processes of Biodiesel Production from Waste Cooking Oil, 104,
pp. 683–710, 2013.
3. Jakeria, M. R., Fazal, M. A., and Haseeb, A. S. M. A., In Fl Uence of Different
Factors on the Stability of Biodiesel: A Review, 30, pp. 154–163, 2014.
4. Tabatabaei, M., Karimi, K., Horváth, I. S., and Kumar, R., Recent Trends in
Biodiesel Production, 7, pp. 258–267, 2015.
5. Rico, J. A. P., and Sauer, I. L., A Review of Brazilian Biodiesel Experiences
A Review of Brazilian Biodiesel Experiences, Renew. Sustain. Energy Rev.,
45(September 2018), pp. 513–529, 2015.
238 Biodiesel Technology and Applications

6. Ž, B., Veljkovi, M. V, Bankovi, I. B., Krsti, I. M., Konstantinovi, S., Ili, S. B.,
Avramovi, J. M., and Stamenkovi, O. S., Risk, Toxicological and Policy
Considerations of Biodiesel Production and Use, 79(July 2016), pp. 222–247,
2017.
7. Sha, I., Manaf, A., Embong, N. H., Norha, S., and Khazaai, M., A Review for
Key Challenges of the Development of Biodiesel Industry, 185(November
2018), pp. 508–517, 2019.
8. Dieter, B., European and Global View ☆. 2019.
9. Souza, S. P., and Seabra, J. E. A., Integrated Production of Sugarcane Ethanol
and Soybean Biodiesel : Environmental and Economic Implications of Fossil
Diesel Displacement, Energy Convers. Manag., 2014.
10. Reshad, A. S., Tiwari, P., and Goud, V. V, Extraction of Oil from Rubber Seeds
for Biodiesel Application : Optimization of Parameters, Fuel, (February),
2015.
11. Fernandes, D. M., Sousa, R. M. F., Oliveira, A. De, Morais, S. A. L., Richter, E.
M., and Muñoz, R. A. A., Moringa Oleifera : A Potential Source for Production
of Biodiesel and Antioxidant Additives, Fuel, 146, pp. 75–80, 2015.
12. Niju, S., Balajii, M., Anushya, C., and Niju, S., A Comprehensive Review on
Biodiesel Production Using Moringa Oleifera Oil, Int. J. Green Energy, 16(9),
pp. 702–715, 2019.
13. Selvaraj, R., Praveenkumar, R., and Moorthy, I. G., A Comprehensive Review
of Biodiesel Production Methods from Various Feedstocks, 7269(November),
2016.
14. Bušić, A., Morzak, G., Belskaya, H., and Ivančić, M., Recent Trends in
Biodiesel and Biogas Production, 56(2), 2018.
15. Muanruksa, P., Winterburn, J., and Kaewkannetra, P., Of, MethodsX, 2019.
16. Manuscript, A., Sustainable Energy & Fuels. 2020.
17. Jose, T. K., and Anand, K., Effects of Biodiesel Composition on Its Long Term
Storage Stability, Fuel, 177, pp. 190–196, 2016.
18. Nongbe, M. C., Ekou, T., Ekou, L., Benjamin, Y. K., Grognec, E. Le, and
Felpin, F., AC SC, Renew. Energy, 2017.
19. Liu, Y., Tu, Q., Knothe, G., and Lu, M., Direct Transesterification of Spent
Coffee Grounds for Biodiesel Production, Fuel, 199, pp. 157–161, 2017.
20. Mueanmas, C., Nikhom, R., Petchkaew, A., Iewkittayakorn, J., and Prasertsit,
K., Extraction and Esterification of Waste Coffee Grounds Oil as Non-Edible
Feedstock for Biodiesel Production, Renew. Energy, 2018.
21. Nabi, G., Sajjad, W., Siddique, R., and Hou, H., HAYATI Journal of Biosciences
Biodiesel Production From Algae to Overcome the Energy Crisis : A Review,
HAYATI J. Biosci., (November), pp. 1–5, 2017.
22. Jena, U., and Hoekman, S. K., Editorial : Recent Advancements in Algae-to-
Biofuels Research : Novel Growth Technologies, Conversion Methods, and
Assessments of Economic and Environmental Impacts, 5(March), pp. 1–2,
2017.
Progress in Biodiesel Production 239

23. Patel, A., Arora, N., Mehtani, J., Pruthi, V., and Pruthi, P. A., Assessment of
Fuel Properties on the Basis of Fatty Acid pro Fi Les of Oleaginous Yeast
for Potential Biodiesel Production, Renew. Sustain. Energy Rev., 77(April),
pp. 604–616, 2017.
24. Martin, A., Ph, F., Researcher, D. S., and Klitkou, A., Energy Research &
Social Science A Fuel Too Far ? Technology, Innovation, and Transition in
Failed Biofuel Development in Norway, Chem. Phys. Lett., 23, pp. 125–135,
2017.
25. Yesilyurt, M. K., Arslan, M., and Eryilmaz, T., Application of Response
Surface Methodology for the Optimization of Biodiesel Production from
Yellow Mustard (Sinapis Alba L.) Seed Oil, Int. J. Green Energy, 00(00),
pp. 1–12, 2018.
26. Veinblat, M., Baibikov, V., Katoshevski, D., Wiesman, Z., and Tartakovsky, L.,
Impact of Various Blends of Linseed Oil-Derived Biodiesel on Combustion
and Particle Emissions of a Compression Ignition Engine – A Comparison
with Diesel and Soybean Fuels, Energy Convers. Manag., 178(August),
pp. 178–189, 2018.
27. Sadaf, S., Iqbal, J., Ullah, I., Nawaz, H., and Nouren, S., Biodiesel Production
from Waste Cooking Oil : An e Ffi Cient Technique to Convert Waste into
Biodiesel, Sustain. Cities Soc., 41(May), pp. 220–226, 2018.
28. Guil-laynez, J. L., and Guil-guerrero, J. L., Industrial Crops & Products
Bioprospecting for Seed Oils in Tropical Areas for Biodiesel Production, Ind.
Crop. Prod., 128(June 2018), pp. 504–511, 2019.
29. Hossain, N., Mahlia, T. M. I., and Saidur, R., Biotechnology for Biofuels Latest
Development in Microalgae ‑ Biofuel Production with Nano ‑ Additives,
Biotechnol. Biofuels, pp. 1–16, 2019.
30. Valdivia, M., and Galan, J. L., Opinion Biofuels 2020: Biore Fi Neries Based
on Lignocellulosic Materials. 2020.
31. Khan, I. U., and Yan, Z., Production and Characterization of Biodiesel
Derived from a Novel Source Koelreuteria Paniculata Seed Oil. 2020.
32. Gebremariam, S. N., and Marchetti, J. M., Biodiesel Production Technologies:
Review, 2017.
33. Biodiesel, O., Singh, M., Ariani, F., Sitorus, T. B., Ginting, E., and Takano, Y.,
An Overview of Biodiesel Production and Its Utilization in Diesel Engines.
2018.
34. Ogunkunle, O., and Ahmed, N. A., A Review of Global Current Scenario
of Biodiesel Adoption and Combustion in Vehicular Diesel Engines, Energy
Reports, 5, pp. 1560–1580, 2019.
35. Karmakar, B., and Halder, G., Progress and Future of Biodiesel Synthesis:
Advancements in Oil Extraction and Conversion Technologies, Energy
Convers. Manag., 182(September 2018), pp. 307–339, 2019.
36. Eddy, A., Imran, A., Ea, B., Seshan, K., and An, B. G., An Overview of
Catalysts in Biomass Pyrolysis for Production of Biofuels. 2020.
240 Biodiesel Technology and Applications

37. Kumar, H., Sarma, A. K., and Kumar, P., A Comprehensive Review
on Preparation, Characterization, and Combustion Characteristics of
Microemulsion Based Hybrid Biofuels, Renew. Sustain. Energy Rev.,
117(February 2019), p. 109498, 2020.
38. Lin, J., and Chen, Y., Production of Biodiesel by Transesterification of
Jatropha Oil with Microwave Heating, J. Taiwan Inst. Chem. Eng., 0, pp. 1–8,
2017.
39. Martinez-guerra, E., and Gude, V. G., Energy Aspects of Microalgal Biodiesel
Production, 4(2), pp. 347–362, 2016.
40. Gude, V. G., and Martinez-guerra, E., Green Chemistry of Microwave-
Enhanced Biodiesel Production.
41. Montcho Papin, S., Konfo, T. R. Christian, Agbangnan, D. C. Pascal,
Sidouhounde Assou and Sohounhloue C. K., Comparative Study of
Transesterification Processes for Biodiesel Production. Elixir Appl. Chem.,
120, pp. 51235–51242, 2018.
42. Thangaraj, B., Solomon, P. R., Muniyandi, B., Ranganathan, S., and Lin, L.,
Catalysis in Biodiesel Production — a Review, pp. 1–22, 2018.
8
Biodiesel Production Technologies
Moina Athar1* and Sadaf Zaidi2
1
Department of Petroleum Studies, Z H College of Engineering and Technology,
Aligarh Muslim University, Aligarh, India
2
Department of Post Harvest Engineering and Technology, Faculty of
Agricultural Sciences, Aligarh Muslim University, Aligarh, India

Abstract
During the past few years, biodiesel has emerged as a potential renewable source
that could replace current petro-diesel. It is a less polluting, biodegradable, and
non-toxic that can be simply produced by a transesterification reaction. Vegetable
oils, animal fats, waste oil, and microbial oils are the main feedstocks for the pro-
duction of biodiesel. The current commercial use of edible and refined oils for
the production of biodiesel is not feasible and uneconomical due to the high cost
of feedstock and priority as a food source. However, non-edible oils which are
low-grade and low cost can be a better option but the presence of a high amount
of free fatty acids (FFAs) in these oils has been the main problem for their use as
potential feedstocks. This chapter is focused to give an overview of the various
feedstocks for the production of biodiesel especially non-edible oils. The benefits
and constraints of using homogeneous, heterogeneous, and enzymatic catalysts
for transesterification of vegetable oil containing a high amount of FFA are dis-
cussed in detail. Toward the end of this chapter, a few of the latest intensification
techniques for biodiesel production that are capable to manage the mass transfer
restrictions of oil and alcohol phases such as supercritical alcohol method, micro-
wave heating, ultrasonic irradiation, and co-solvent method are discussed. Some
other techniques that reduce the biodiesel production cost like in situ transesteri-
fication, membrane reactor, reactive distillation, static mixers, micro-channel, and
oscillatory flow have also been discussed briefly.
Keywords: Biodiesel feed stock, production methods, catalyst, microwave
heating, ultrasonic irradiation, co-solvent method, in situ method, static mixture

*Corresponding author: mathar.pe@amu.ac.in

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (241–266) © 2021 Scrivener Publishing LLC

241
242 Biodiesel Technology and Applications

8.1 Introduction
Present crises of fossil fuels, hike in crude oil prices, increased environmen-
tal pollution, and environmental awareness have changed the world’s focus
on the use of alternative fuels [1]. Biodiesel is one of the most eco-friendly,
renewable, and promising alternative for diesel engine [2, 3]. Chemically
biodiesel is defined as mono-alkyl esters of long-chain fatty acids (FAs)
present in vegetable oils or animal fats [4–9] which on combustion pro-
duces fewer pollutants like sulfur, carbon dioxide, particulate matters, car-
bon monoxide, smoke, and unburned hydrocarbons emissions. Complete
combustion takes place due to excess oxygen present in biodiesel which
also reduces emission [10, 11]. It can be mixed with petrodiesel in any
proportion and can be used in the present diesel engine with no or very
few engine modifications. The mixtures are represented by acronyms like
B20, which signifies a mixture of 20% biodiesel with petro diesel [12, 13].
Biodiesel provides fuel efficiency, torque, and horsepower similar to stan-
dard diesel fuel [14].

8.2 Biodiesel Feedstocks


Great varieties of feedstocks can be used for biodiesel production. More
than 350 oil containing seeds have the potential for biodiesel production
[15]. These feedstocks include both edible and non-edible type of vegetable
oils, fats of animals, different waste oils, and microbial oils. Table 8.1 shows

Table 8.1 Various types of oil and their sources for the production of biodiesel.
Type of oil Source of oil
Vegetable oils Edible Soya bean, mustard, rice bran, sunflower,
cottonseed, olive, peanut, sesame, palm, etc.
Non-edible Jatropha, mahua, jojoba, neem, karanja, milk
bush, rubber seed, nagchampa, castor,
petroleum nut, silk cotton tree, tung, etc.
Animal fat Chicken fat, beef tallow, fish oil, pork lard
Waste oil Waste frying oil, pit oil
Microbial oil Fungi, microalgae (chlorellavulg), algae
(cyanobacteria)
Biodiesel Production Technologies 243

3% 2% 1%
7%
Cost of oil
Cost of chemical
12%
Depreciation cost

Direct labour cost


Energy cost
75%
General overhead
cost

Figure 8.1 Cost breakup for biodiesel production [7, 21–23].

the various types of oils and their sources that can be used for the produc-
tion of biodiesel [16–19].
The availability of this vast range of feedstock is one of the major rea-
sons for producing biodiesel [5, 8, 20]. Low production cost and large pro-
duction scale are the two main demands which should be fulfilled by any
of the feedstocks. The feedstocks present in any region or area depend-
ing on the geographical locations, regional climate, conditions of soil, and
adopted agricultural methods of that country [15]. It was established from
various studies that biodiesel feedstock alone contributes to 75% of the
total production cost as depicted in Figure 8.1. Hence, the selection of low-
cost feedstock is essential for economical biodiesel production. Figure 8.2
shows the various feedstocks used for the production of biodiesel.

8.2.1 Selection of Feedstocks


Triglycerides are the basic component of all types of oils used for biodiesel
production. Chemically, they are esters of FAs with glycerol and contain
several different types of FAs. As the chemical and physical properties of
different FAs differ with each other, so they also affect the various prop-
erties of the associated oils and fats. Table 8.2 presents the quantities of
FA presents in selected fats and oils. At present, edible oils comprise more
than 95% of the main feedstocks for biodiesel production worldwide [16].
However, the utilization of edible oils for biodiesel production is uneco-
nomical and impractical because of its high cost and its priority use as food
sources. Thus, low quality [high free FA (FFA)] and low-cost non-edible
oil that is not being used in human nutrition could be an option. Plants
for non-edible oil seeds are easily grown in lands that are not appropriate
244 Biodiesel Technology and Applications

Soyabean
y Sunflower Mustard

Rice bran Cotton seed Palm seeds

Jatropha Jajoba j
Karanja

Mahua Neem Rubber seeds

Animal fat Waste frying oil Algal oil

Figure 8.2 Various types of feedstock for biodiesel production.

for edible oil crops with much lower costs [24]. Growing of these plants
can also reduce the amount of CO2 in the atmosphere [25]. Apart from all
the above-mentioned advantages, the presence of a high amount of FFA,
which leads to high production cost, is the major drawback associated with
these low-grade non-edible oils [26].
Jatropha is among the most prospective non-edible source for the pro-
duction of biodiesel in Central and South America, South-East Asia, Africa,
and India. Presently, it is the main feedstock of biodiesel in developing
countries like India, where its yearly production is around 15,000 t [27].
Jatropha can be cultivated anywhere like on saline and sandy soil, waste-
lands, in different climatic conditions, with low or high rainfall without
Table 8.2 Fatty acid profile of fats and oils [34].
C(12:0) C(14:0) C(16:0) C(16:1) C(18:0) C(18:1) C(18:2) C(18:3) C(20:1) C(22:1)
Lauric Myristic Palmitic Palmitoleic Stearic Oleic Linoleic Linolenic Gadoleic Erucic
Feedstock acid acid acid acid acid acid acid acid acid acid Unknowns
Algae – 0.6 6.9 0.2 3.0 75.2 12.4 1.2 – – Remaining
amount
Babassu 48.8 17.2 9.7 – 4.0 14.2 1.8 – – – Remaining
amount
Beef Tallow 0.2 2.9 24.3 2.1 22.8 40.2 3.3 0.7 0.6 – Remaining
amount
Borage – – 9.3 – 3.8 17.1 38.7 26.1 – 2.5 Remaining
amount
Camelina Oil – – 5.0 – 2.2 17.7 18.0 37.9 9.8 4.5 Remaining
amount
Canola Oil – – 3.8 0.3 1.9 63.9 19.0 9.7 – – Remaining
amount
Castor – – 0.9 – 1.1 93.4 4.0 0.6 – – Remaining
amount
Choice white – 1.3 21.6 2.8 9.0 50.4 12.2 1.0 0.5 0.3 Remaining
grease amount
Biodiesel Production Technologies

Coconut 49.2 18.5 9.1 – 2.7 6.5 1.7 – – – Remaining


amount
245

(Continued)
Table 8.2 Fatty acid profile of fats and oils [34]. (Continued)
246

C(12:0) C(14:0) C(16:0) C(16:1) C(18:0) C(18:1) C(18:2) C(18:3) C(20:1) C(22:1)
Lauric Myristic Palmitic Palmitoleic Stearic Oleic Linoleic Linolenic Gadoleic Erucic
Feedstock acid acid acid acid acid acid acid acid acid acid Unknowns
Coffee – – 11.0 0.5 3.4 70.0 12.7 0.8 0.1 – Remaining
amount
Corn – – 12.1 0.1 1.8 27.2 56.2 1.3 – – Remaining
amount
Cuphea – 4.7 18.2 – 3.5 46.9 22.8 2.3 – – Remaining
Viscosissima amount
Evening – – 6.0 – 1.8 6.6 76.3 9.0 – – Remaining
Primrose amount
Hemp – – 5.2 – 2.4 13.1 57.1 20.0 – – Remaining
amount
Jatropha – – 12.7 0.7 5.5 39.1 41.6 0.2 – – Remaining
Biodiesel Technology and Applications

amount
Linseed – – 4.4 – 3.8 20.7 15.9 54.6 – – Remaining
amount
Mustard – – 2.6 0.2 1.2 20.6 20.6 13.3 10.7 25.6 Remaining
amount
Neem – – 14.9 0.1 20.6 43.9 17.9 0.4 – – Remaining
amount
(Continued)
Table 8.2 Fatty acid profile of fats and oils [34]. (Continued)
C(12:0) C(14:0) C(16:0) C(16:1) C(18:0) C(18:1) C(18:2) C(18:3) C(20:1) C(22:1)
Lauric Myristic Palmitic Palmitoleic Stearic Oleic Linoleic Linolenic Gadoleic Erucic
Feedstock acid acid acid acid acid acid acid acid acid acid Unknowns
Palm 0.2 0.5 43.4 0.1 4.6 41.9 8.6 0.3 – – Remaining
amount
Perilla Seed – – 5.3 0.1 2.2 16.6 13.7 62.1 – – Remaining
amount
Poultry Fat 0.1 1.0 19.6 3.2 7.5 36.8 28.4 2.0 – 0.4 Remaining
amount
Rice Bran – 0.3 12.5 – 2.1 47.5 35.4 1.1 – – Remaining
amount
Soybean – – 9.4 – 4.1 22.0 55.3 8.9 – – Remaining
amount
Sunflower – – 4.2 – 3.3 63.6 27.6 0.2 – – Remaining
amount
Used cooking 0.1 0.1 11.8 0.4 4.4 25.3 49.5 7.1 – 0.3 Remaining
oil amount
Yellow Grease 0.1 0.5 14.3 1.1 8.0 35.6 35.0 4.0 – 0.2 Remaining
Biodiesel Production Technologies

amount
247
248 Biodiesel Technology and Applications

much care and with minimal efforts. The life cycle of the Jatropha plant is
30–40 years, which eliminates its yearly plantation. Although the oil con-
tent of Jatropha depends on the type of species as 40%–60% of oil is present
in its seeds and 46%–58% is present in kernels [28]. It is a potential alter-
native to petro-diesel as it has equivalent properties to diesel, like calorific
value, cetane number, and does not require any engine modification [27,
29]. While, its toxicity is a serious problem for the people and animals [30].
Among all the new generation non-edible feedstocks, microalgae are
being considered to be the most potential one. Microalgae produce oil in
the presence of water, sunlight, and carbon dioxide in a more efficient way
than crop plants. As different types of algae contain different amounts of oil
so among all the species only a few that contain higher amounts of oil are
appropriate for the synthesis of biodiesel. The oil content of some microal-
gae exceeded significantly the oil yield of some of the best oil-producing
crops [31, 32]. Moreover, it can be easily grown anywhere, like in sewage
or salty water, and no fertile land is needed and most of all its processing
needed less energy than it provides [33].

8.3 Biodiesel Production Technologies


Polyunsaturated character, viscous, and low volatile vegetable oils are not
suitable for diesel engines for direct use [35]. Refinement of these oils is
needed to transform them into quality fuel. The principal methods which
enable oils and fats to be used in diesel engines are usually direct use by
dilution, micro-emulsion, pyrolysis, and transesterification.

8.3.1 Pyrolysis
Pyrolysis is defined as chemical change after applying heat without air or
oxygen or by using catalyst resulted in chemical bond breakage and pro-
duction of different small molecules. The changes during pyrolysis take
place at a reaction temperature range from 400°C to 600°C using different
vegetable or animal oils, naturally accruing FAs or esters of FAs. Process
of thermal breakage of triglycerides yields aromatics, alkenes, alkanes,
alkadines, and carboxylic acids, which are fuel components for the die-
sel engine; however, some undesirable components are also produced
due to the absence of air during this process [35–37]. Low heating value,
incomplete volatility, and instability of the product are some unwanted
properties that often limit the use of biodiesel produced through pyrolysis
[38, 39].
Biodiesel Production Technologies 249

8.3.2 Dilution
The direct use of vegetable oil in an existing diesel engine is limited by its
high viscosity, carbon deposits, FFA contents, gum formation by oxidation
and polymerization processes, etc. In such situations, dilution of vegetable
oils with diesel fuel, solvents, or ethanol is recommended which reduces
its viscosity and density [40]. The use of pure oil in an existing petrodiesel
engine is not feasible; instead, a mixture of 20% oil and 80% petro-diesel is
recommended [41].

8.3.3 Micro-Emulsion
It is another effort toward the viscosity reduction of vegetable oils. The pro-
cess of biodiesel microemulsion is carried out by the use of petro-diesel,
alcohol, vegetable oil, surfactant, and cetane improvers in appropriate pro-
portions. Some low molecular weight alcohols like methanol and ethanol
are used as additives for viscosity reduction; high molecular weight alcohols
can be utilized as surfactants whereas alkyl nitrates are compounds that are
used for improving the cetane number [42]. Micro-emulsion reduces the
viscosity, improves cetane rating, and introduces excellent spray quality in
the biodiesel.

8.3.4 Transesterification
Transesterification is the reaction of oil and alcohol with or without a cat-
alyst to produce esters (biodiesel) and glycerol as a product (Figure 8.3).
Since the reaction involved is reversible, an extra volume of alcohol is
needed to support the forward reaction and to shift the equilibrium to the
product’s side [43]. The transesterification reaction is completed in three
steps in which the triglyceride conversion to diglycerides is the first step,
followed by the subsequent diglycerides conversion into monoglycerides
and then to glycerol, producing one ester molecule from each glyceride at
each step (Figure 8.4) [44].

CH2 OCOR1 CH2OH R1COOCH3


Catalyst
CH OCOR2 + 3CH3OH CHOH + R2COOCH3

CH2 OCOR3 CH2OH R3COOCH3

Triglyceride Methanol Glycerol Methyl esters

Figure 8.3 General transesterification reaction.


250 Biodiesel Technology and Applications

Triglyceride + R1OH Diglyceride + RCOOR1

Diglyceride + R1OH Monoglyceride + RCOOR1

Monoglyceride + R1OH Glyceride + RCOOR1

Figure 8.4 General mechanism of the transesterification reaction.

The transesterification reaction is carried out in the presence of acidic,


basic, or enzymatic catalysts. It can be homogeneously or heterogeneously
catalyzed, based upon the catalyst solubility in the reactants. The reaction
can be in a single step (by acid or base catalyst) or two steps (acid/base),
based on the amount of FFA content of the oil.

8.3.4.1 Homogeneously Catalyzed Transesterification Processes


Homogeneous catalysts can be basic or acidic type and their selection
mainly depends upon the FFA and water content of the oil.

8.3.4.1.1 Alkaline Catalyst


The alkaline catalysts are used for oils containing a low percentage of FFA
(less than 5 wt.%) [45]. Thus, the most common homogeneous alkaline
catalysts that have been used and successfully commercialized for pro-
duction of biodiesel are potassium hydroxide (KOH), sodium hydrox-
ide (NaOH), and sodium methoxide (CH3ONa) [3, 5, 43, 46–48]. Mostly
sodium hydroxide and potassium hydroxide are used by biodiesel cata-
lysts. Although the conversions in the presence of sodium and potassium
methoxides are better, they are not very frequently used as they are costly.
Transesterification by the alkaline catalyst is the fastest and economical
method compared to other catalytic methods. It is 4,000 times faster than
the acid catalyst of the same amount and achieves a high yield of biodiesel
with high purity in less reaction time (30–60 min) [15]. Therefore, it is
preferred over the other biodiesel production catalyst.

8.3.4.1.2 Acidic Catalyst


Despite many associated problems like the slower rate of reaction, a higher
volume of alcohol requirement, less catalyst activity, and requirement of
higher reaction temperature, the acid catalyst has some important advan-
tages for the transesterification reaction [49]. Low sensitivity for the high
FFAs present in the non-edible oils (>5%) and the chances for esterification
Biodiesel Production Technologies 251

O O
Acid catalyst
R” -C-OH + R’-OH R” -C-OR’ + H2O

Carboxylic acid Alcohol Biodiesel Water

Figure 8.5 Mechanism of the esterification reaction.

and transesterification reaction to occur simultaneously are some notable


advantages of the acidic catalyst [50]. Esterification is carried out of FFA
and alcohol (e.g., methanol) with the help of a strong acidic catalyst to form
at least one ester and water as a reaction product (Figure 8.5). The common
acid catalyst used for esterification includes sulfuric, hydrochloric, phos-
phoric, ferric sulfate, and organic sulfonic acid. The yield of biodiesel was
reported up to 90% in many studies of transesterification with an acid cata-
lyst, but a much longer reaction time (3–48 h) was required, in comparison
to the alkali-catalyst transesterification reaction, except for those carried
out with higher reaction temperatures [51–53]. The main variables that
influence the transesterification yield are the amount of catalyst, alcohol
type, and the reaction time [52]. Hence, the main benefit of using an acidic
catalyst is the low-cost production of biodiesel by using high FFA non-
edible oil in a single step [54]. However, these acidic catalysts are corrosive
and create a handling issue during the operation.

8.3.4.1.3 Two-Step Process


Two-step acid/base catalytic process has been developed to address the
drawbacks of single-step homogeneous acid and base catalytic processes
for biodiesel production using high FFA non-edible oils. In this two-
step method, the first step comprises of acid catalyst esterification of FFA
(pre-treatment) for minimizing the FFA below 1% followed by alkali-­
catalyzed transesterification [55]. The acid pretreatment process eliminates
the problem of slow reaction rate and soap formation. The higher cost of
production is the only problem associated with the two-step transesterifi-
cation process [49, 56]. Apart from the great advantages of homogeneous
catalysts for transesterification, numerous disadvantages like huge energy
requirement, the produced soap as a by-product, difficult and costly sep-
aration of catalyst, and a large amount of wastewater generation are also
associated with them [57, 58].
252 Biodiesel Technology and Applications

8.3.4.2 Heterogeneously Catalyzed Transesterification Processes


The catalyst which is not miscible with the reaction mixture is known as
heterogeneous. It may be a basic or acidic type and is more conveniently
separated from reaction products. Acidic heterogeneous catalysts could be
used for transesterification of high FFA oil as they avoid undesirable sapon-
ification reaction. Solid heterogeneous catalysts could lead to cheaper pro-
duction of biodiesel as it can be reused and enable both esterification and
transesterification to be carried out simultaneously [59]. However, a slow
reaction rate due to difficult diffusion in immiscible phases of oil, alcohol,
and catalyst mixture, complex catalyst production method, followed by
a considerable effect on the environment are the major disadvantages of
heterogeneous catalysts [60]. The basic heterogeneous catalyst commonly
used is alkaline earth metal oxides (SrO, BaO, CaO, MgO), alkaline earth
metal carbonates (CaCO3), alkaline metal carbonates (K2CO3, Na2CO3),
and transition metal oxides (ZnO) [44, 61–63]. However, acidic heteroge-
neous catalysts for transesterification such as Nafion-NR50 (perfluorinated
alkane sulfonic acid resin), Zirconium oxide (ZrO2), sulfated zirconia-­
alumina (SZA), sulfated oxides, cation-exchange resins, and tungstated
zirconia–alumina (WZA) are being utilized for the synthesis of biodiesel
[64, 65]. Besides, a solid acidic catalyst like sulfated zirconia and tungstated
zirconia can be used to support the transesterification as well as the esteri-
fication reactions [66, 67].

8.3.4.3 Enzymatic Catalyzed Transesterification Processes


Normally, enzymes are a bio-oriented catalyst, primarily natural lipases
that are separated from some bacterial species like Pseudomonas fluo-
rescens, Rhizopus oryzae, Thermomyces lanuginosus, Candida antarctica,
Pseudomonas cepacia, Candida rugosa, and Rhizomucor miehei [68].
Biodiesel production using enzymes has no problems of washing, purifica-
tion, neutralization, and saponification, so it is usually recommended from
these aspects. It can be used for those non-edible feedstocks having a high
amount of FFA and gives appreciable yield. Enzymatic catalysts are typi-
cally used in the immobilized form to ensure their stability and reusability.
The different ways by which immobilization can be provided are entrap-
ment, adsorption, encapsulation, covalent bonding, cross-linking, and by
the process of enzyme embedding on a solid support. These methods are
useful to give the stability and recycling of the enzyme and decrease the
total expenses on the catalyst that was the main challenge in commercial-
ization [69].
Biodiesel Production Technologies 253

8.4 Intensification Techniques for Biodiesel


Production
Numerous novel techniques are being used in the past few years for the
enhance the biodiesel yield especially from low-cost non-edible oils such
as supercritical alcohol method, microwave heating, ultrasonic irradiation,
co-solvent method, in situ transesterification processes, motionless mixer
technique, and membrane reactors.

8.4.1 Supercritical Alcohol Method


Biodiesel production by transesterification method with alcohol (metha-
nol, ethanol, propanol, and butanol) in its supercritical stage with or with-
out catalyst has been considered as the most promising technique. The
supercritical method takes very little time to complete the reaction, whereas
conventional catalytic transesterification method needs many hours to fin-
ish [70]. At conventional processing temperatures, oils (non-polar) and
alcohol (polar) form heterogeneous mixture because of the limited misci-
bility of the polar and non-polar reactants. In a supercritical environment,
the alcohol and triglyceride mixture turns into a single homogeneous
phase, which will enhance the reaction rate due to the non-existence of
any interphase mass transfer. The typical range of operating conditions
applied for the supercritical zone has been temperatures of 280°C–400°C
and pressures of 10–30 MPa [71]. The supercritical approach is found to be
more water and FFA resistant and more easily separated and purifies the
biodiesel more than the conventional catalytic transesterification method
[72, 73]. Alternatively, in the supercritical condition along with transes-
terification, esterification of FFA and hydrolysis of triglycerides due to the
presence of moisture takes place simultaneously and enhances the yield of
methyl ester [72, 74, 75]. The cost of high temperature and pressure reactor
is the main drawback that limits its use on a large scale in the industry [76].
Co-solvents, like carbon dioxide [77, 78], hexane [76, 79], propane [80],
and calcium oxide [81, 82] with a little quantity of catalyst added during
the reaction can reduce the temperature, pressure and the alcohol quantity.

8.4.2 Microwave Heating


The transesterification by microwave heating is an effective way for the
rapid production of biodiesel. During this process, heat is produced by
the friction of molecules due to the alignment of polar alcohol molecules
254 Biodiesel Technology and Applications

by the constantly varying magnetic field of microwaves. Due to very


effective heat transfer, reactions under microwave heating can take place
in a greatly reduced reaction time than those with conventional heating
[83, 84]. Microwaves which are of 0.3- to 300-GHz frequency range have
both thermal and non-thermal types of effects and generate heat due to
its molecular interaction in the sample without changing their structures
[85, 86]. Apart from the high yield of biodiesel, microwave heating also
reduces the separation and purification time [87, 88]. The microwave heat-
ing method has many advantages over the conventional heating method,
such as reduction of a thermal gradient, low overheating of surfaces due to
contactless heating, energy transfer rather than heat transfer, and heat gen-
eration from inside of the mass. In sum, microwave-induced transesteri-
fication offers a much-purified product with a high yield and reduces the
separation and purification time [87, 88]. Therefore, microwave heating
is a more environmentally compatible, more energy-efficient, and favor-
able technique as compared to conventional heating, where heat trans-
fer mainly takes place by conduction and convection by the walls of the
sample container to reach inner materials resulted in inefficient and slow
heating due to high thermal gradient because of non-uniform heating [89,
90]. Apart from the enormous advantages of this method, there are also a
few drawbacks. The scale-up of microwave technique from laboratory to
industrial multi-kilogram production is not very easy. Penetration depth
which is only a few centimeters is the major constraint for the scale-up of

Figure 8.6 Microwave reactor.


Biodiesel Production Technologies 255

this technology. The safety is another cause of rejecting microwave heating


in the industry [89]. Figure 8.6 shows an open view of a microwave reactor
used for laboratory purposes.

8.4.3 Ultrasonic Irradiation


The ultrasonic method is a newer mixing technique which is more efficient
for the production of biodiesel, compared to the conventional way of agi-
tation. Ultrasound is sound waves with frequencies higher than the upper
audible limit of human hearing. The typical range of hearing is between
16 Hz and about 18 kHz and ultrasound is normally known to be between
20 kHz and more than 100 MHz [91]. The ultrasonic mixing improves the
mass transfer among the immiscible oil and alcohol phase, accelerates the
chemical reaction and conversion of triglyceride, and reduces the reaction
time and energy consumption. One of the types of ultra-sonicator (probe
sonicator) that is normally used in laboratories is shown in Figure 8.7.
The molecular spacing of the medium gets compressed and stretched on
the application of high-frequency ultrasonic sound wave that continuously
vibrates the molecules and creates the cavities. The created micro-fine bub-
bles by sudden collapse and expansion and generate energy for chemical
and mechanical effects [92]. Additionally, the phase boundary is disrupted
by the collapsed bubbles, and micro-jets are created by the impinging of
the liquids, leading to extensive emulsification of the system [93].

Figure 8.7 Probe type ultra sonicator.


256 Biodiesel Technology and Applications

8.4.4 Co-Solvent Method


During the transesterification reaction, the oil and methanol phases are
immiscible with each other which affects the rate of mass transfer among
these phases. This mass transfer constraint could be overcome by the addi-
tion of a small amount of co-solvent which is soluble both in methanol as
well as in the triglycerides. The addition of the co-solvent increases the
transesterification rate because of the elimination of mass transfer resis-
tance between the phases. Many studies have revealed that the presence
of co-solvent significantly reduces the reaction temperature, pressure,
and quantity of alcohol and enhances the reaction rate [94–96]. Various
co-solvents like tetrahydrofuran (THF), hexane, and a mixture of THF
and hexane, acetone, propane, heptane, carbon dioxide (CO2), dimethyl
ether (DME), and diethyl ether (DEE) have been used in biodiesel produc-
tion [97, 98]. However, the use of co-solvent increases the cost of biodiesel
production due to the extra cost of separation of the co-solvent from the
methanol and product phase [99].

8.5 Other Techniques of Biodiesel Production


Some other techniques like in situ transesterification, membrane reactor,
reactive distillation, static mixers, micro-channel, oscillatory flow have
also been used to make the process economical by either improving sepa-
ration and purification method, reducing equipment cost, or making the
process simpler.
During the in situ transesterification method, extraction and transes-
terification of oil take place simultaneously. The used alcohol behaves as
a solvent for extraction as well as a reagent for transesterification reaction
[100]. After the oil is extracted from the seeds by the alcohol, it is simul-
taneously converted into ester due to already present alcohol. Hence, this
process not only reduces the processing time but also reduces the cost. The
only drawback associated is the requirement of a large volume of alcohol
as compared to other traditional processes [18].
Membrane reactors combine reaction and separation in a single step
which reduces recycling and separation costs and increases conversion due
to the enhancement of product inhibited reactions. Inorganic membranes
are much better than the organic (polymeric) ones as these are generally
more thermally stable [101, 102]. A higher amount of methanol and ultra-
low catalyst concentration is needed as compared to the conventional pro-
duction methods [103, 104].
Biodiesel Production Technologies 257

The distillation in which the reaction and the separation of the associated
products take place in the same column is known as reactive distillation.
Both packed and tray-type columns are suitable for reactive distillation,
even though tray columns are recommended for the homogenous type of
reactions due to the better liquid hold-up and prolonged retention time
[105, 106]. Higher conversion, reduced energy consumption, lower capital
cost, better selectivity, none or reduced amount of solvents throughout the
reaction, and avoidance of azeotrope formation are some benefits of reac-
tive distillation [107, 108]. The continuous removal of products from the
reaction system not only increases the conversion but also helps to reduce
capital and investment costs [109, 110].
Static mixers are made up of specially designed static helical mixing part
in a hollow cylinder that creates intense mixing and needs less energy in
comparison to conventionally used mechanical mixers [111, 112]. They are
usually used in continuous processes for biodiesel production. This has
no moving elements and therefore carry the benefit of low maintenance
and operating costs, along with low space requirements [113, 114]. By
static mixer, more fine and uniform droplets of methanol are generated,
which enhances the interfacial surface area between raw oil and metha-
nol. Thus, a higher reaction rate and a greater yield of FAME are obtained
than that with the mechanical mixer [115, 116]. Apart from the aforemen-
tioned technologies, micro-channel [117–120] and oscillatory flow [121,
122] techniques can also improve reaction rate and reduce the alcohol-
to-oil molar ratio and energy input by the intensification of mass transfer
and heat transfer, thus helping in attaining continuous production in the
reactor.

References
1. L. Reijnders, Conditions for the sustainability of biomass based fuel use,
Energy Policy, vol. 34, no. 7, pp. 863–876, 2006.
2. S. Fernando, C. Hall, and S. Jha, NO Reduction from Biodiesel Fuels, Energy
& Fuels, vol. 20, no. 1, pp. 376–382, 2006.
3. L. C. Meher, D. Vidya Sagar, and S. N. Naik, Technical aspects of biodiesel
production by transesterification - A review, Renew. Sustain. Energy Rev., vol.
10, no. 3, pp. 248–268, 2006.
4. A. C. Pinto et al., Biodiesel: An overview, J. Braz. Chem. Soc., vol. 16, no. 6 B,
pp. 1313–1330, 2005.
5. E. M. Shahid and Y. Jamal, Production of biodiesel: A technical review,
Renew. Sustain. Energy Rev., vol. 15, no. 9, pp. 4732–4745, 2011.
258 Biodiesel Technology and Applications

6. G. Kafuku and M. Mbarawa, Biodiesel production from Croton megalocar-


pus oil and its process optimization, Fuel, vol. 89, no. 9, pp. 2556–2560, 2010.
7. A. L. Ahmad, N. H. M. Yasin, C. J. C. Derek, and J. K. Lim, Microalgae as a
sustainable energy source for biodiesel production: A review, Renew. Sustain.
Energy Rev., vol. 15, no. 1, pp. 584–593, 2011.
8. J. Janaun and N. Ellis, Perspectives on biodiesel as a sustainable fuel, Renew.
Sustain. Energy Rev., vol. 14, no. 4, pp. 1312–1320, 2010.
9. M. Satyanarayana and C. Muraleedharan, A comparative study of vegetable
oil methyl esters (biodiesels), Energy, vol. 36, no. 4, pp. 2129–2137, 2011.
10. A. S. Silitonga, A. E. Atabani, T. M. I. Mahlia, H. H. Masjuki, I. A. Badruddin,
and S. Mekhilef, A review on prospect of Jatropha curcas for biodiesel in
Indonesia, Renew. Sustain. Energy Rev., vol. 15, no. 8, pp. 3733–3756, 2011.
11. M. A. Fazal, A. S. M. A. Haseeb, and H. H. Masjuki, Biodiesel feasibility
study: An evaluation of material compatibility; Performance; emission and
engine durability, Renew. Sustain. Energy Rev., vol. 15, no. 2, pp. 1314–1324,
2011.
12. D. N. Tziourtzioumis and A. M. Stamatelos, Experimental investigation of
the effect of biodiesel blends on a DI diesel engine’s injection and combus-
tion, Energies, vol. 10, no. 7, 2017.
13. S. Islam, A. S. Ahmed, A. Islam, S. A. Aziz, L. C. Xian, and M. Mridha, Study
on Emission and Performance of Diesel Engine Using Castor Biodiesel,
J. Chem., pp. 1–8, 2014.
14. J. Xue, T. E. Grift, and A. C. Hansen, Effect of biodiesel on engine perfor-
mances and emissions, Renew. Sustain. Energy Rev., vol. 15, no. 2, pp. 1098–
1116, 2011.
15. A. E. Atabani, A. S. Silitonga, I. A. Badruddin, T. M. I. Mahlia, H. H. Masjuki,
and S. Mekhilef, A comprehensive review on biodiesel as an alternative
energy resource and its characteristics, Renew. Sustain. Energy Rev., vol. 16,
no. 4, pp. 2070–2093, 2012.
16. M. M. Gui, K. T. Lee, and S. Bhatia, Feasibility of edible oil vs. non-­edible
oil vs. waste edible oil as biodiesel feedstock, Energy, vol. 33, no. 11,
pp. 1646–1653, 2008.
17. P. T. Vasudevan and M. Briggs, Biodiesel production—current state of the art
and challenges, J. Ind. Microbiol. Biotechnol., vol. 35, no. 5, p. 421, 2008.
18. I. B. Banković-Ilić, O. S. Stamenković, and V. B. Veljković, Biodiesel produc-
tion from non-edible plant oils, Renew. Sustain. Energy Rev., vol. 16, no. 6,
pp. 3621–3647, 2012.
19. A. Demirbas, Progress and recent trends in biodiesel fuels, Energy Convers.
Manag., vol. 50, no. 1, pp. 14–34, 2009.
20. I. M. Atadashi, M. K. Aroua, and A. A. Aziz, High quality biodiesel and its
diesel engine application: A review, Renew. Sustain. Energy Rev., vol. 14,
no. 7, pp. 1999–2008, 2010.
Biodiesel Production Technologies 259

21. S. S. Ragit, S. K. Mohapatra, K. Kundu, and P. Gill, Optimization of neem


methyl ester from transesterification process and fuel characterization as a
diesel substitute, Biomass and Bioenergy, vol. 35, no. 3, pp. 1138–1144, 2011.
22. S. Lim and L. K. Teong, Recent trends, opportunities and challenges of bio-
diesel in Malaysia: An overview, Renew. Sustain. Energy Rev., vol. 14, no. 3,
pp. 938–954, 2010.
23. L. Lin, Z. Cunshan, S. Vittayapadung, S. Xiangqian, and D. Mingdong,
Opportunities and challenges for biodiesel fuel, Appl. Energy, vol. 88, no. 4,
pp. 1020–1031, 2011.
24. S. A. Basha, K. R. Gopal, and S. Jebaraj, A review on biodiesel production,
combustion, emissions and performance, Renew. Sustain. Energy Rev.,
vol. 13, no. 6–7, pp. 1628–1634, 2009.
25. A. Karmakar, S. Karmakar, and S. Mukherjee, Properties of various plants
and animals feedstocks for biodiesel production, Bioresour. Technol.,
vol. 101, no. 19, pp. 7201–7210, 2010.
26. D. Y. C. Leung, X. Wu, and M. K. H. Leung, A review on biodiesel pro-
duction using catalyzed transesterification, Appl. Energy, vol. 87, no. 4,
pp. 1083–1095, 2010.
27. S. Jain and M. P. Sharma, Prospects of biodiesel from Jatropha in India: A
review, Renew. Sustain. Energy Rev., vol. 14, no. 2, pp. 763–771, 2010.
28. A. Kumar and S. Sharma, Potential non-edible oil resources as biodiesel
feedstock: An Indian perspective, Renew. Sustain. Energy Rev., vol. 15, no. 4,
pp. 1791–1800, 2011.
29. P. Sirisomboon, P. Kitchaiya, T. Pholpho, and W. Mahuttanyavanitch, Physical
and mechanical properties of Jatropha curcas L. fruits, nuts and kernels,
Biosyst. Eng., vol. 97, no. 2, pp. 201–207, 2007.
30. M. Balat and H. Balat, Progress in biodiesel processing, Appl. Energy,
vol. 87, no. 6, pp. 1815–1835, 2010.
31. B. J. Gallagher, The economics of producing biodiesel from algae, Renew.
Energy, vol. 36, no. 1, pp. 158–162, 2011.
32. A. E. F. Abomohra, W. Jin, R. Tu, S. F. Han, M. Eid, and H. Eladel, Microalgal
biomass production as a sustainable feedstock for biodiesel: Current status
and perspectives, Renew. Sustain. Energy Rev., vol. 64, pp. 596–606, 2016.
33. A. Demirbas and M. Fatih Demirbas, Importance of algae oil as a source of
biodiesel, Energy Convers. Manag., vol. 52, no. 1, pp. 163–170, 2011.
34. G. R. M. Shannon D. Sanford, James Matthew White, Parag S. Shah, Claudia
Wee, Marlen A. Valverde, Feedstock and Biodiesel Characteristics Report,
Renew. Energy Group, Inc, pp. 1–136, 2009.
35. A. Srivastava and R. Prasad, Triglycerides-based diesel fuels, Renew. Sustain.
energy Rev., vol. 4, no. 2, pp. 111–133, 2000.
36. A. W. Schwab, G. J. Dykstra, E. Selke, S. C. Sorenson, and E. H. Pryde, Diesel
fuel from thermal decomposition of soybean oil, J. Am. Oil Chem. Soc.,
vol. 65, no. 11, pp. 1781–1786, 1988.
260 Biodiesel Technology and Applications

37. R. Saidur, E. A. Abdelaziz, A. Demirbas, M. S. Hossain, and S. Mekhilef, A


review on biomass as a fuel for boilers, Renew. Sustain. Energy Rev., vol. 15,
no. 5, pp. 2262–2289, 2011.
38. A. Abbaszaadeh, B. Ghobadian, M. R. Omidkhah, and G. Najafi, Current
biodiesel production technologies: A comparative review, Energy Convers.
Manag., vol. 63, pp. 138–148, 2012.
39. R. French and S. Czernik, Catalytic pyrolysis of biomass for biofuels produc-
tion, Fuel Process. Technol., vol. 91, no. 1, pp. 25–32, 2010.
40. A. Bilgin, O. Durgun, and Z. Şahin, The effects of diesel-ethanol blends on
diesel engine performance, Energy Sources, vol. 24, no. 5, pp. 431–440, 2002.
41. S. P. Singh and D. Singh, Biodiesel production through the use of different
sources and characterization of oils and their esters as the substitute of diesel:
A review, Renew. Sustain. Energy Rev., vol. 14, no. 1, pp. 200–216, 2010.
42. D. Chiaramonti et al., Development of emulsions from biomass pyrolysis
liquid and diesel and their use in engines - Part 1: Emulsion production,
Biomass and Bioenergy, vol. 25, no. 1, pp. 85–99, 2003.
43. F. Ma and M. A. Hanna, Biodiesel production: A review, Bioresour. Technol.,
vol. 70, pp. 1–15, 1999.
44. H. Fukuda, A. Kondo, and H. Noda, Biodiesel fuel production by transester-
ification of oils, J. Biosci. Bioeng., vol. 92, no. 5, pp. 405–416, 2001.
45. Z. Ilham and S. Saka, Two-step supercritical dimethyl carbonate method for
biodiesel production from Jatropha curcas oil, Bioresour. Technol., vol. 101,
no. 8, pp. 2735–2740, 2010.
46. J. M. Ã. Marchetti, V. U. Miguel, and A. F. Errazu, Possible methods for bio-
diesel production, Renew. Sustain. Energy Rev., vol. 11, pp. 1300–1311, 2007.
47. G. Corro, N. Tellez, E. Ayala, and A. Marinez-ayala, Two-step biodiesel pro-
duction from Jatropha curcas crude oil using SiO 2 Á HF solid catalyst for
FFA esterification step, Fuel, vol. 89, no. 10, pp. 2815–2821, 2010.
48. S. Jain and M. P. Sharma, Biodiesel production from Jatropha curcas oil,
Renew. Sustain. Energy Rev., vol. 14, no. 9, pp. 3140–3147, Dec. 2010.
49. W. Parawira, Biodiesel production from Jatropha curcas: A review, Sci. Res.
Essays, vol. 5, no. 14, pp. 1796–1808, 2010.
50. A. P. Vyas, J. L. Verma, and N. Subrahmanyam, A review on FAME produc-
tion processes, Fuel, vol. 89, 2010.
51. S. M. P. Meneghetti et al., Biodiesel from castor oil: A comparison of etha-
nolysis versus methanolysis, Energy and Fuels, vol. 20, no. 5, pp. 2262–2265,
2006.
52. N. Saravanan, S. Puhan, G. Nagarajan, and N. Vedaraman, An experimental
comparison of transesterification process with different alcohols using acid
catalysts, Biomass and Bioenergy, vol. 34, no. 7, pp. 999–1005, 2010.
53. X. Deng, Z. Fang, and Y. H. Liu, Ultrasonic transesterification of Jatropha
curcas L. oil to biodiesel by a two-step process, Energy Convers. Manag.,
vol. 51, no. 12, pp. 2802–2807, 2010.
Biodiesel Production Technologies 261

54. E. Lotero, Y. Liu, D. E. Lopez, K. Suwannakarn, D. A. Bruce, and J. G.


Goodwin, Synthesis of biodiesel via acid catalysis, Ind. Eng. Chem. Res., vol.
44, no. 14, pp. 5353–5363, 2005.
55. M. Canakci and J. VanGerpen, A pilot plant to produce biodiesel from
high free fatty acid feedstocks, in Sacramento, CA July 29-August 1, 2001,
pp. 1–19.
56. C. Vijaya Lakshmi, K. Viswanath, S. Venkateshwar, and B. Satyavathi,
Mixing characteristics of the oil-methanol system in the production of bio-
diesel using edible and non-edible oils, Fuel Process. Technol., vol. 92, no. 8,
pp. 1411–1417, 2011.
57. N. S. Babu, R. Sree, P. S. S. Prasad, and N. Lingaiah, Room-Temperature
Transesterification of Edible and Nonedible Oils Using a Heterogeneous
Strong Basic Mg/La Catalyst, Energy & Fuels, vol. 22, no. 3, pp. 1965–1971,
May 2008.
58. W. Xie and H. Li, Alumina-supported potassium iodide as a heterogeneous
catalyst for biodiesel production from soybean oil, J. Mol. Catal. A Chem.,
vol. 255, no. 1–2, pp. 1–9, 2006.
59. D. E. López, J. G. Goodwin, D. A. Bruce, and E. Lotero, Transesterification of
triacetin with methanol on solid acid and base catalysts, Appl. Catal. A Gen.,
vol. 295, no. 2, pp. 97–105, 2005.
60. I. K. Mbaraka and B. H. Shanks, Advanced Mesoporous Catalytic Materials,
JAOCS, vol. 83, no. 2, pp. 79–91, 2006.
61. F. Ma and M. A. Hanna, Biodiesel production: A review, Bioresource
Technology. 1999.
62. X. Liu, H. He, Y. Wang, and S. Zhu, Transesterification of soybean oil to
biodiesel using SrO as a solid base catalyst, Catal. Commun., vol. 8, no. 7,
pp. 1107–1111, 2007.
63. M. Verziu et al., Sunflower and rapeseed oil transesterification to biodiesel
over different nanocrystalline MgO catalysts, Green Chem., vol. 10, no. 4,
pp. 373–38, 2008.
64. F. Chai, F. Cao, F. Zhai, Y. Chen, X. Wang, and Z. Su, Transesterification of
vegetable oil to biodiesel using a heteropolyacid solid catalyst, Adv. Synth.
Catal., vol. 349, no. 7, pp. 1057–1065, 2007.
65. D. E. López, J. G. Goodwin, and D. A. Bruce, Transesterification of triacetin
with methanol on Nafion®acid resins, J. Catal., vol. 245, no. 2, pp. 381–391,
2007.
66. V. Rathore, B. L. Newalkar, and R. P. Badoni, Processing of vegetable oil
for biofuel production through conventional and non-conventional routes,
Energy Sustain. Dev., vol. 31, pp. 24–49, 2016.
67. S. Furuta, H. Matsuhashi, and K. Arata, Biodiesel fuel production with solid
superacid catalysis in fixed bed reactor under atmospheric pressure, Catal.
Commun., vol. 5, no. 12, pp. 721–723, 2004.
68. A. B. R. Moreira, V. H. Perez, G. M. Zanin, and H. F. de Castro, Biodiesel
synthesis by enzymatic transesterification of palm oil with ethanol using
262 Biodiesel Technology and Applications

lipases from several sources immobilized on silica-PVA composite, Energy


and Fuels, vol. 21, no. 6, pp. 3689–3694, 2007.
69. M. K. H. Leung, A review on biodiesel production using catalyzed transes-
terification A review on biodiesel production using catalyzed transesterifica-
tion, Appl. Energy, vol. 87, no. 4, pp. 1083–1095, 2009.
70. H. He, S. Sun, T. Wang, and S. Zhu, Transesterification kinetics of soybean
oil for production of biodiesel in supercritical methanol, JAOCS, J. Am. Oil
Chem. Soc., vol. 84, no. 4, pp. 399–404, 2007.
71. V. F. Marulanda, G. Anitescu, and L. L. Tavlarides, Investigations on super-
critical transesterification of chicken fat for biodiesel production from low-
cost lipid feedstocks, J. Supercrit. Fluids, vol. 54, no. 1, pp. 53–60, 2010.
72. S. Saka and D. Kusdiana, Biodiesel fuel from rapeseed oil as prepared in
supercritical methanol, Fuel, vol. 80, no. 2, pp. 225–231, 2001.
73. K. Bunyakiat, S. Makmee, R. Sawangkeaw, and S. Ngamprasertsith,
Continuous Production of Biodiesel via Transesterification from Vegetable
Oils in Supercritical Methanol, Energy & Fuels, vol. 20, no. 2, pp. 812–817,
2006.
74. D. Kusdiana and S. Saka, Kinetics of transesterification in rapeseed oil to bio-
diesel fuel as treated in supercritical methanol, Fuel, vol. 80, no. 5, pp. 693–
698, 2001.
75. D. Kusdiana and S. Saka, Effects of water on biodiesel fuel production by
supercritical methanol treatment, Bioresour. Technol., vol. 91, no. 3, pp. 289–
295, 2004.
76. J. Z. Yin, M. Xiao, and J. Bin Song, Biodiesel from soybean oil in supercritical
methanol with co-solvent, Energy Convers. Manag., vol. 49, no. 5, pp. 908–
912, 2008.
77. Y. Warabi, D. Kusdiana, and S. Saka, Reactivity of triglycerides and fatty acids
of rapeseed oil in supercritical alcohols, Bioresour. Technol., vol. 91, no. 3, pp.
283–287, 2004.
78. Y. Warabi, D. Kusdiana, and S. Saka, Biodiesel Fuel from Vegetable Oil by
Various Supercritical Alcohols, Appl. Biochem. Biotechnol., vol. 115, no. 1–3,
pp. 0793–0802, 2004.
79. J. Z. Yin, M. Xiao, A. Q. Wang, and Z. L. Xiu, Synthesis of biodiesel from
soybean oil by coupling catalysis with subcritical methanol, Energy Convers.
Manag., vol. 49, no. 12, pp. 3512–3516, 2008.
80. W. Cao, H. Han, and J. Zhang, Preparation of biodiesel from soybean oil
using supercritical methanol and co-sol vent, Fuel, vol. 84, no. 4, pp. 347–
351, 2005.
81. A. Demirbas, Biodiesel from sunflower oil in supercritical methanol with
calcium oxide, Energy Convers. Manag., vol. 48, no. 3, pp. 937–941, 2007.
82. Z. Tang, L. Wang, and J. Yang, Transesterification of the crude Jatropha cur-
cas L. oil catalyzed by micro-NaOH in supercritical and subcritical metha-
nol, Eur. J. Lipid Sci. Technol., vol. 109, no. 6, pp. 585–590, 2007.
Biodiesel Production Technologies 263

83. N. Azcan and A. Danisman, Alkali catalyzed transesterification of cotton-


seed oil by microwave irradiation, Fuel, vol. 86, no. 17–18, pp. 2639–2644,
2007.
84. H. Venkatesh Kamath, I. Regupathi, and M. B. Saidutta, Optimization of two
step karanja biodiesel synthesis under microwave irradiation, Fuel Process.
Technol., vol. 92, no. 1, pp. 100–105, 2011.
85. R. S. Varma, Solvent-free accelerated organic syntheses using microwaves,
Pure Appl. Chem., vol. 73, no. 1, pp. 193–198, 2001.
86. A. A. Refaat, Different techniques for the production of biodiesel from waste
vegetable oil, Int. J. Environ. Sci. Tech., vol. 7, no. 1, pp. 183–213, 2010.
87. M. Nüchter, B. Ondruschka, A. Jungnickel, and U. Müller, Organic processes
initiated by non-classical energy sources, J. Phys. Org. Chem., vol. 13, no. 10,
pp. 579–586, 2000.
88. J. Hernando, P. Leton, M. P. Matia, J. L. Novella, and J. Alvarez-Builla,
Biodiesel and FAME synthesis assisted by microwaves: Homogeneous batch
and flow processes, Fuel, vol. 86, no. 10–11, pp. 1641–1644, 2007.
89. Y. Groisman and A. Gedanken, Continuous flow, circulating microwave sys-
tem and its application in nanoparticle fabrication and biodiesel synthesis,
J. Phys. Chem. C, vol. 112, no. 24, pp. 8802–8808, 2008.
90. A. A. Refaat, S. T. Sheltawy, and K. U. Sadek, Optimum reaction time, per-
formance and exhaust emissions of biodiesel produced by microwave irradi-
ation, Int. J. Environ. Sci. Technol., vol. 5, no. 3, pp. 315–322, 2008.
91. T. J. Mason, Ultrasound in synthetic organic chemistry, Chem. Soc. Rev., vol.
26, no. 6, pp. 443–451, 1997.
92. J. A. Colucci, E. E. Borrero, and F. Alape, Biodiesel from an alkaline trans-
esterification reaction of soybean oil using ultrasonic mixing, JAOCS, J. Am.
Oil Chem. Soc., vol. 82, no. 7, pp. 525–530, 2005.
93. J. Ji, J. Wang, Y. Li, Y. Yu, and Z. Xu, Preparation of biodiesel with the help
of ultrasonic and hydrodynamic cavitation, Ultrasonics, vol. 44, no. SUPPL.,
pp. e411–e414, Dec. 2006.
94. E. Çağlar, Biodiesel Production Using Co-solvent, in Book of Abstracts
European Congress of Chemical Engineering (ECCE-6) Copenhagen,16-20
Septembe, 2007.
95. G. Guan, K. Kusakabe, N. Sakurai, and K. Moriyama, Transesterification of
vegetable oil to biodiesel fuel using acid catalysts in the presence of dimethyl
ether, Fuel, vol. 88, no. 1, pp. 81–86, 2009.
96. S. Sakthivel, S. Halder, and P. D. Gupta, Influence of Co-Solvent on the
Production of Biodiesel in Batch and Continuous Process, Int. J. Green
Energy, vol. 10, no. 8, pp. 876–884, Sep. 2013.
97. C. Calgaroto, S. Calgaroto, M. A. Mazutti, D. de Oliveira, S. Pergher, and J. de
Oliveira, Production of biodiesel from soybean and Jatropha Curcas oils with
KSF and amberlyst 15 catalysts in the presence of co-solvents, Sustain. Chem.
Process., vol. 1, no. 1, p. 17, 2013.
264 Biodiesel Technology and Applications

98. L. T. Thanh, K. Okitsu, Y. Sadanaga, N. Takenaka, Y. Maeda, and H. Bandow,


A new co-solvent method for the green production of biodiesel fuel -
Optimization and practical application, Fuel, vol. 103, pp. 742–748, Jan.
2013.
99. G. R. Kumar, R. Ravi, and A. Chadha, Kinetic studies of base-catalyzed trans-
esterification reactions of non-edible oils to prepare biodiesel: The effect of
Co-solvent and temperature, Energy and Fuels, vol. 25, no. 7, pp. 2826–2832,
2011.
100. G. Knothe, J. Van Gerpen, and J. Krahl, The Biodiesel Handbook, 1st ed.
(AOCS Press), 2005.
101. P. Cao, A. Y. Tremblay, M. A. Dubé, and K. Morse, Effect of Membrane Pore
Size on the Performance of a Membrane Reactor for Biodiesel Production,
Ind. Eng. Chem. Res., vol. 46, no. 1, pp. 52–58, 2007.
102. J. Saleh, A. Y. Tremblay, and M. A. Dubé, Glycerol removal from biodiesel
using membrane separation technology, Fuel, vol. 89, no. 9, pp. 2260–2266,
2010.
103. P. Cao, M. A. Dubé, and A. Y. Tremblay, High-purity fatty acid methyl ester
production from canola, soybean, palm, and yellow grease lipids by means of
a membrane reactor, Biomass and Bioenergy, vol. 32, no. 11, pp. 1028–1036,
2008.
104. L. H. Cheng, S. Y. Yen, L. S. Su, and J. Chen, Study on membrane reactors for
biodiesel production by phase behaviors of canola oil methanolysis in batch
reactors, Bioresour. Technol., vol. 101, no. 17, pp. 6663–6668, 2010.
105. A. C. Dimian, C. S. Bildea, F. Omota, and A. A. Kiss, Innovative process for
fatty acid esters by dual reactive distillation, Comput. Chem. Eng., vol. 33,
no. 3, pp. 743–750, 2009.
106. B. B. He, A. P. Singh, and J. C. Thompson, A Novel Continuous-Flow
Reactor Using Reactive Distillation For Biodiesel Production, Trans. ASABE
(American Soc. Agric. Biol. Eng., vol. 49, no. 1, pp. 107–112, 2006.
107. C. Mueanmas, K. Prasertsit, and C. Tongurai, Feasibility Study of Reactive
Distillation Column for Transesterification of Palm Oils, Int. J. Chem. Eng.
Appl., vol. 1, no. 1, pp. 77–83, 2010.
108. N. M. Eleftheriades and H. von Blottnitz, Thermodynamic and kinetic con-
siderations for biodiesel production by reactive distillation, Environ. Prog.
Sustain. Energy, vol. 32, no. 2, pp. 373–376, Jul. 2013.
109. A. A. Kiss, Novel process for biodiesel by reactive absorption, Sep. Purif.
Technol., vol. 69, pp. 280–287, 2009.
110. G. B. Shinde, V. S. Sapkal, R. S. Sapkal, and N. B. Raut, Transesterification
by Reactive Distillation for Synthesis and Characterization of Biodiesel, in
Biodiesel - Feedstocks and Processing Technologies, Dr. Margarita Stoytcheva,
Ed. InTech, 2011, p. 29.
111. H. Noureddini, D. W. Harkey, and M. R. Gutsman, A Continuous Process for
the Glycerolysis of Soybean Oil, JAOCS, J. Am. Oil Chem. Soc., vol. 81, no. 2,
pp. 203–207, 2004.
Biodiesel Production Technologies 265

112. C. L. Peterson, J. L. Cook, J. C. Thompson, and J. S. Taberski, Continuous


flow biodiesel production, Appl. Eng. Agric., vol. 18, no. 1, pp. 5–11, 2002.
113. J. C. Thompson and B. B. He, Biodiesel Production Using Static Mixers,
Trans. Asabe, vol. 50, no. 1, pp. 161–166, 2007.
114. M. B. Boucher, C. Weed, N. E. Leadbeater, B. a Wilhite, J. D. Stuart, and R. S.
Parnas, Pilot scale two-phase continuous flow biodiesel production via novel
laminar flow reactor- separator, Energy & Fuels, vol. 23, no. 14, pp. 2750–
2756, 2009.
115. R. Alamsyah, A. H. Tambunan, Y. A. Purwanto, and D. Kusdiana, Comparison
of Static-Mixer and Blade Agitator Reactor in Biodiesel Production, Agric.
Eng. Int. CIGR Ejournal, vol. 12, no. 4, pp. 1–12, 2010.
116. P. Sungwornpatansakul, J. Hiroi, Y. Nigahara, T. K. Jayasinghe, and K.
Yoshikawa, Enhancement of biodiesel production reaction employing the
static mixing, Fuel Process. Technol., vol. 116, pp. 1–8, 2013.
117. N. Chueluecha, A. Kaewchada, and A. Jaree, Biodiesel synthesis using het-
erogeneous catalyst in a packed- microchannel, Energy Convers. Manag.,
2016.
118. Y. Natarajan et al., An overview on the process intensification of microchan-
nel reactors for biodiesel production, Chem. Eng. Process. - Process Intensif.,
vol. 136, pp. 163–176, 2019.
119. A. Mohd Laziz, K. Z. KuShaari, B. Azeem, S. Yusup, J. Chin, and J. Denecke,
Rapid production of biodiesel in a microchannel reactor at room tempera-
ture by enhancement of mixing behaviour in methanol phase using volume
of fluid model, Chem. Eng. Sci., vol. 219, p. 115532, 2020.
120. H. S. Santana, J. L. S. Júnior, and O. P. Taranto, Numerical simulations of
biodiesel synthesis in microchannels with circular obstructions, Chem. Eng.
Process. Process Intensif., vol. 98, pp. 137–146, 2015.
121. A. P. Harvey, M. R. Mackley, and T. Seliger, Process intensification of bio-
diesel production using a continuous oscillatory flow reactor, J. Chem.
Technol. Biotechnol., vol. 78, no. 2–3, pp. 338–341, 2003.
122. N. Masngut, A. P. Harvey, and J. Ikwebe, Potential uses of oscillatory baffled
reactors for biofuel production, Biofuels, vol. 1, no. 4, pp. 605–619, 2010.
9
Methods for Biodiesel Production
M.Gul1,2*, M.A. Mujtaba1,3, H.H. Masjuki1,4, M.A. Kalam1 and N.W.M. Zulkifli1
1
Center for Energy Science, Department of Mechanical Engineering,
University of Malaya, Kuala Lumpur, Malaysia
2
Department of Mechanical Engineering, Faculty of Engineering and
Technology, Bahauddin Zakariya University, Multan, Pakistan
3
Department of Mechanical Engineering, University of Engineering and
Technology, City Campus Lahore, Pakistan
4
Department of Mechanical Engineering, Faculty of Engineering, IIUM,
Kuala Lumpur, Malaysia

Abstract
This chapter discusses all types of feedstocks used for synthesizing biodiesel
and feedstock’s selection criteria. Moreover, all biodiesel production meth-
ods (i.e Dilution with hydrocarbons blending, Micro-emulsion, Pyrolysis, and
Transesterification) are also described in detail with their advantages and dis-
advantages. The major focus is given to the various Transesterification methods
because of their simplicity, easy handling, less time consumption, and maximum
biodiesel yield. Production methods also include experimental setup layouts, all
process parameters, reaction conditions, the latest advancement in reaction pro-
cesses, and their effects on biodiesel yield.
Keywords: Biodiesel, pyrolysis, transesterification, supercritical methanol

9.1 Selection of Feedstock for Biodiesel


There are different types of biodiesel feedstock that mainly comprised of
fatty acids (triglycerides) or lipids depending on local soil conditions in dif-
ferent regions and climate. Genetic engineering is also playing important
role in growing biodiesel feedstock crops to achieve desired composition

*Corresponding author: mustabshirha@bzu.edu.pk

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (267–284) © 2021 Scrivener Publishing LLC

267
268 Biodiesel Technology and Applications

according to application and to overcome their drawbacks. So, selection of


feedstock includes fatty acids (lipid) content/oil percentage, free fatty acid
(FFA) contents, moisture content, and its availability and productivity per
hectare [1].
The major biodiesel producers all over the world are the European Union
(EU) (rapeseed, soybeans), the United States (Soybean), Brazil (Soybean),
Indonesia and Malaysia (palm oil), and China (rapeseed) that use edible
vegetable oil (VO) for synthesizing biodiesel. The following are the kinds
of biodiesel feedstock.

9.1.1 First-Generation Feedstock


Edible vegetable oils like palm oil, soybeans, rapeseed, coconut, sunflower,
safflower, sesame, corn, peanut, groundnut, almond, barley, poppy seed,
okra seed, oat oil, camelina (camelina sativa), fish oil, and cottonseed oils
are considered as first-generation feedstock for producing biodiesel. Most
of the biodiesel are synthesized from edible oils. But, in future, world will
face severe malnutrition and serious basic food shortage issues. Growing
of crops for getting biodiesel feedstock will utilize land, water, and other
energy resources that are also required significantly to fulfil basic food pro-
duction for human beings.

9.1.2 Second-Generation Feedstock


Non-edible oils are defined as second-generation feedstock of biodiesel
production that obtained from jatropha tree/ratanjyote (jatropha curcas),
jojoba (simmondsia chinensis), neem (azadirachta indica), karanja/honge
(pongamia pinnata), desert date (balanites aegyptiaca), koroch seed oil, silk
cotton tree (ceiba pentandra), mahua (madhuca indica or madhuca longi-
folia), pachira glabra nagchampa (calophyllum inophyllum), tobacco seed
(nicotiana tabacum L.), sea mango seed (cerbera odollam), rubber seed tree
(hevea brasiliensis), aleurites moluccana, croton megalocarpus, terminalia
belerica, babassu tree, euphorbia tirucalli, and rice bran.
Significant non-edible oils are getting attention in developing countries
as these have no competition and dispute with food. Non-edible oil crops
are also environment friendly, efficient, and very economical, can easily
grow on wastelands, and reduce deforestation rate. Moreover, animal fats
like tallow from beef, sheep and poultry, pork lard, Chinese tallow, waste
cooking oils, and grease are also categorized as second-generation feed-
stocks for biodiesel synthesis. Utilization of these feedstock for biodiesel
production also eliminates their disposal problems but collection of waste
Methods for Biodiesel Production 269

oils from scattered sources required logistics and proper infrastructure.


Transesterification of some animal fats is not so easy because those have
high concentration of saturated fatty acids and exist in solidified form.

9.1.3 Third-Generation Feedstock


Recently, microalgae emerged as third-generation promising feedstock
for biodiesel production. Microalgae are photosynthetic microorganisms
having different varieties that convert CO2 and water into algal biomass
by photosynthesis process. Algal biomass has high productivity, faster
growth rates, and high yield of oil as compared to VO sources. Microalgae
can be grown in farm or in bioreactors that needs high production costs.
Moreover, its commercialization requires large-scale bioreactors so more
researches are required to explore its production methods in future [2].

9.2 Methods for Biodiesel Production


VOs in pure form are not good quality fuels for IC engines because of their
polyunsaturated fatty acids characteristics, high viscosity, and low vola-
tility in comparison to diesel fuels. These drawbacks can be resolved by
conventional methods like dilution by blending hydrocarbons and micro-­
emulsion. Both of these are physical methods so do not involve any chem-
ical reaction.

9.2.1 Dilution With Hydrocarbons Blending


VOs cannot replace diesel fuels completely, so high viscosities of VO can be
decreased by blending it with diesel. In addition, 20% VO and 80% diesel
blend was successfully implemented in IC engines as described in earlier
literatures [3–5]. High percentage blends are not recommended because
these do not fulfill ASTM standard limits.

9.2.2 Micro-Emulsion
In this method, oils are mixed and blended with immiscible solvents
including ionic/non-ionic amphiphiles and methanol, ethanol, butanol,
and hexanol to reduce their viscosities. This method ensures the thermo-
dynamically stable colloidal dispersions of micro-phase particles having as
small diameter as ¼ part of visible light’s wavelength. As micro-emulsions
contain alcohol so give low volumetric heating values in comparison to
270 Biodiesel Technology and Applications

diesel fuels, but useful cooling effect developed in combustion chamber


because of high latent heat of vaporization of alcohol.
These above two methods have improved the viscosity issues and rec-
ommended as fuel for short term applications. However, these fuels gave
incomplete combustion in diesel engines, premature injection‐nozzle dete-
rioration occurred, and high carbon deposits are observed on the piston
lands, intake ports and exhaust valves, and piston ring grooves. These car-
bon deposits also caused degradation of lubricating oil. So, at present, fuels
prepared from pure VOs by dilution and micro-emulsion are not suggested
suitable anymore for long term utilization in diesel engines [6].
Some new techniques/methods are adopted to solve all these problems
associated with dilution and micro-emulsion. These methods involves
pyrolysis, transesterification, and supercritical methanol [7].

9.2.3 Pyrolysis (Thermal Cracking)


In pyrolysis heat is used to decompose VOs, fatty acids, and animal fats
by using different types of catalysts without oxygen to synthesized suitable
potential fuel for DI-diesel engines. Pyrolysis is a chemical reaction tak-
ing place by thermal energy that convert triglycerides of VOs into alkanes,
alkenes, alkadines, linear and cyclic paraffins and olefins, aromatics, alde-
hydes, ketones, and carboxylic acids [8], as described in Figure 9.1.
O

CH3 (CH2)5 CH2 –CH2 CH=CHCH2 –CH2 (CH2)5C–O–CH2R

CH9 (CH2)5 CH2 –CH2 CH=CHCH2 –CH2 (CH2)5C–OH

CH3 (CH2)5CH2• CH2 =CHCH=CH2 • CH2 (CH2)5C–OH

H
O
Diels-
CH9 (CH2)3CH2 • + CH2=CH2 Alder CH3(CH2)5C–H

–CO2
–H2

CH3(CH2)3CH3 CH3(CH2)4CH3

Figure 9.1 Pyrolysis of vegetable oils into alkanes, alkenes, and aromatics [4].
Methods for Biodiesel Production 271

Saponification and pyrolysis of VOs directly produce the hydrocarbon-­


rich products [9] that are very good alternative of gasoline (a competitive
biodiesel fuel) as presented by Equations (9.1) and (9.2)

Saponification:
Vegetable oil/fats + NaOH RCOONa (sodium soap) + Glycerin
(9.1)

Pyrolysis of sodium soap:


4 RCOONa + (1/2) O2 R—R (biodiesel) + Na2CO3 + CO2
(9.2)

The products obtained after pyrolysis have good cetane no and low vis-
cosity with acceptable range of copper corrosion value, sulfur, and water
content. But, their poor pour point, low higher heating values (HHVs),
more carbon residues, and ash content have limited their utilization in spe-
cific applications [10].

9.2.4 Transesterification (Alcoholysis)


Transesterification is most preferable chemical reaction and consid-
ered best method among all other processes because of its simplicity
and less cost [4, 11]. Synthesis of biodiesel by transesterification pro-
cess can be performed by conventional transesterification method or by
microwave/­ultrasound-assisted method. Microwave/ultrasound-assisted
method takes very less reaction time and synthesizes biodiesel more
quickly at laboratory scale.

9.2.4.1 In Situ Transesterification (Reactive Extraction)


In in situ transesterification process, crop seeds containing oil are crushed
and directly reacted with methanol/ethanol (alcohol) and catalyst (acid
or alkali based) to produce biodiesel, thereby excluding the need for oil
extraction and reducing the intensive running and capital cost of biodiesel
synthesis methods [12]. Some co-solvents are also used to speed up the
reaction rate [13].
Some significant parameters of in situ transesterification process are
moisture content of seeds, oil-to-alcohol molar ratios, type of alcohol, type
272 Biodiesel Technology and Applications

and concentration of catalyst, reaction temperature, agitation speed, and


size of crushed seeds [12].

9.2.4.2 Conventional Transesterification


In conventional transesterification, oils are extracted from seed by crush-
ing, pressing, and solvent extraction (mostly by n-hexane) and then used
as raw material.
Figure 9.2 shows the difference between in situ and conventional trans-
esterification processes. Conventional process requires crushing and sol-
vent extraction steps so having more cost than in situ.
VO is pre-heated at 100°C to evaporate water; then, this VO is trans-
ferred into double jacket glass reactor and heated up to desired tempera-
ture. Optimized amount of catalyst is dissolved in alcohol and poured into
reactor and temperature of mixture is maintained below the boiling point
of alcohol (65°C) by using hot water bath. Chemical reaction between oil
and alcohols proceeds for 2 hours under continuous stirring. Then, this
mixture transferred into separating funnel for 12–20 hours at room tem-
perature; after that, two layers are formed by sedimentation. Conventional
transesterification experimental setup for biodiesel production is described
in Figure 9.3. Bottom layer of glycerol is removed and top layer containing
biodiesel is heated at 80°C to remove excess methanol. At the end, this
biodiesel is washed with 40°C warm water to eliminate remaining alcohol,
glycerol, and catalyst. Water is absorbed anhydrous Na2SO4 from biodiesel
and after filtration purified biodiesel is obtained.
There are various types of transesterification process [2] based on differ-
ent types of catalysts and without catalyst as classified in Figure 9.4.

Steps Conventional method In-situ method

1. Raw Material Crop seeds Crop seeds

2. Maceration Crushed into small size Crushed into small size

3. Crushing & solvent Oil extraction


extraction
4. Transesterification Biodiesel + Glycerol Biodiesel + Glycerol
(oil+alcohol+catalyst)
5. Purification Pure Biodiesel Pure Biodiesel

Figure 9.2 Comparison between conventional and in situ transesterification.


Methods for Biodiesel Production 273

Stirrer

Separating
funnel
Condenser
Thermometer

Biodiesel

Glycerol

Double jacket Circulated Refrigerated


glass reactor hot-water bath cooling solvent

Figure 9.3 Conventional transesterification process experimental setup.

Acidic: H2SO4, HCL, H3PO4


ferric sulfate, and organic
sulfonic acid
Homogeneous
Alkaline: KOH, NaOH,
Catalysts NaOCH3, KOCH3, and
K2CO3 etc
Catalytic-based Enzymes e.g immobilized
process lipase, Ryzopus oryzae
Heterogeneous lipase, CH2N2,
Catalysts
Titanium silicates, Sulfated
Titanium, Alkaline earth
metal (MgO, CaO, SrO),
Amorphous zirconia,
Transesterification Potassium zirconia

Supercritical
Methanol (SCM)
Ethanol/Propanol
Butanol
Non-catalytic
based process

BIOX co-solvent
Process

Figure 9.4 Classification of transesterification process.

9.2.4.2.1 Catalytic-Based Process


9.2.4.2.1.1 Homogeneous-Acid-Catalyzed-Esterification
(Pre-Transesterification Treatment)
If oils have FFA > 3 wt%, then esterification process (pre-­transesterification
treatment) is carried out to reduced FFAs less than 3 wt%. FFA can be
determined from the acidic values of oils by using Equation (9.3).
274 Biodiesel Technology and Applications

FFA = 0.5% * AV (9.3)

In pre-transesterification treatment, VO having high FFA is reacted


chemically with alcohol (methanol) in the presence of homogeneous acidic
catalyst like sulfuric acid/H2SO4 or ferric sulfate which convert the FFA
into biodiesel [14] as shown by Equation (9.4)

Free Fatty Acids (FFA) + ROH(Alcohol) Catalyst


 → FAME (Biodiesel) + H 2O
Acid
(9.4)

Then, this mixture is transferred into separating funnel along with cold
water and allowed to settle for appropriate time until two phases appeared.
Upper layer contains excess of water-methanol fraction that is removed,
and bottom layer containing fatty acid methyl ester, VO, and methanol
was transferred into evaporator to remove existing methanol, and then,
its acid value is determined prior to transesterification process. Then, this
obtained biodiesel + VO is purified and neutralized by different kind of
methods [15]; most preferable is transesterification.

9.2.4.2.1.2 Homogeneous Alkaline-Catalyzed


Transesterification
Most efficient alkaline-catalyzed transesterification is carried out on those
oils that have less than or equal to 1–3 wt% range of FFA ≤ 1–3 wt% [16,
17] because presence of more FFAs produce soaps by reacting with alkaline
catalyst that makes difficulty in separating the biodiesel and glycerol [18].
It is necessary that FFA content should be less than 3% to achieve greater
biodiesel conversion efficiency/yield [19].
Transesterification is the most popular method for producing biodiesel
[11, 20]; in this process, short-chain alcohols (methanol, ethanol, propa-
nol, butanol) are reacted chemically with VO (crop seed oils) or fats in the
presence of alkali catalysts (like KOH, NaOH, and sodium methoxide) as
shown by Equation (9.5). Mostly, methanol is preferred among all other
alcohols due to its low cost. VOs have long-chain triglyceride structure of
fatty acids. Mostly, methanol is preferable alcohol because of its low cost
among other alcohols.
CH2 COOR1 CH2 OH CH3 COOR1
Alkaline
CH COOR2 + 3 CH3OH
Catalyst
CH OH + CH3 COOR2 (9.5)
CH2 COOR3 CH2 OH CH3 COOR3
Triglyceride of oil Alcohol (Methanol) Glycerol Biodiesel (FAME)
Methods for Biodiesel Production 275

The conversion of triglyceride (VO) into diglyceride, monoglycer-


ide, and glycerol along with biodiesel is a reversible three-step process as
described by Equation (9.6) [4].
End products of alkaline-catalyzed transesterification process are glyc-
erol and biodiesel (FAME). When this mixture is transferred to separating
funnel, then glycerol sinks to the bottom of funnel and biodiesel remains
at the top of glycerol. After removing glycerol from bottom, biodiesel is
heated and filtered to remove alcohol and catalyst than biodiesel is washed
and purified with warm water.
Catalyst
1. Triglyceride (Vegetable oil) + ROH (Alcohol) → Diglyceride +FAME (Biodiesel)
Catalyst
2. Diglyceride + ROH → Monoglyceride +FAME
Catalyst
(9.6)
3. Monoglyceride + ROH → Glycerol +FAME
∴ R is alkyl group of hydro- carbons

9.2.4.2.1.3 Heterogeneous-Catalyzed Transesterification


Transesterification process also used heterogeneous catalysts like metal
oxides, oxides, mixed metal oxides, hydrotalcites (like aluminum and
magnesium hydroxycarbonates, and zeolites), and supported hydroxides.
Heterogeneous catalysts increase the biodiesel production, reduce sapon-
ification, and easy to recover from biodiesel than homogeneous catalyst.
But, these required high temperature and pressure [21].
Alkaline earth metal (CaO, MgO, SrO), titanium silicates, sulfates, tita-
nium, amorphous zirconia, potassium zirconia, and unsupported mixed
metal oxides including CaO-La2O3 CaMnOx, BaMnOx, CaFeOx, BaFeOx,
CaZrOx, and CaCeOx are the examples of heterogeneous catalyst. Table 9.1
showed the Comparison b/w heterogeneous and homogeneous catalytic
transesterification process.

9.2.4.2.1.4 Enzymatic-Catalyzed Transesterification


There are various types of eco-friendly immobilized or soluble enzymatic
catalysts that are also suitable for feedstock having more FFA and water
under mild reaction conditions [22]. Enzymes include immobilized lipase,
ryzopus oryzae lipase, liquid lipase eversa transform, callera trans L lipase,
liquid lipolase, lipozyme, novozyme, ZIF-67, novozym 435, lipase immobi-
lized, and CH2N2. Lipases enzymes are synthesized from plant, animal, and
microorganism sources. These are relatively costly but require low energy
consumption. Some companies are using enzymatic catalysts at industrial
276 Biodiesel Technology and Applications

Table 9.1 Comparison between heterogeneous and homogeneous catalytic


transesterification process.
Heterogeneous Homogeneous
Factors catalytic process catalytic process
Existence of water/free Not sensitive Sensitive
fatty acids
Cost Potentially cheaper Comparatively costly
Reaction rate Moderate conversion Fast and high
conversion
Catalyst reuse Possible Not possible
Post-treatment Catalyst can be Acidic catalyst are used
recovered in neutralization so
cannot be recovered,
and produce
chemical waste

levels including Piedmont Biofuel (USA), Sunho Biodiesel Corporation


(Taiwan), Lvming Co. Ltd, and Hainabaichuan Co. Ltd (China). Cost-
effective catalysts can also be synthesized by combining multiple types of
enzymes and lipases but alkali or acidic catalyzed transesterification pro-
cesses are more economical and practical than enzymatic catalyzed.

9.2.4.2.2 Non-Catalytic-Based Process


Non-catalytic transesterification process does not use catalyst, so there is
no need of catalyst removal and no saponification occurs, so glycerol is
removed easily. It takes very less reaction time because of high solubility
between alcohol and triglycerides of oils. Mostly, non-catalytic transesteri-
fication is performed by following two routes: (a) supercritical alcohol pro-
cess and (b) BIOX co-solvent process [4, 23].

9.2.4.2.2.1 Supercritical Alcohol Process


The non-catalytic transesterification of VO can be performed with super-
critical alcohols (like methanol/ethanol/propanol/butanol) which requires
high temperatures (170°C–350°C), high pressure (10–60 MPa), and high
alcohol consumption (molar ratio as 40–42), thereby increasing opera-
tional costs. In this process, fats or oils are treated with subcritical water
then glycerol is removed from fatty acids (oil) phase by decantation. Fatty
Methods for Biodiesel Production 277

Vegetable oil (Triglyceride + free fatty acids)

Hydrolysis by subcritical water

Phase separation

Oil phase (fatty acids) Water phase

Transesterification with Phase separation


supercritical alcohol

Phase separation
Waste water Glycerol

Alcohol recovery by Biodiesel


rotary evaporator

Biodiesel ready
Purification
for marketing

Figure 9.5 Supercritical transesterification process.

acid phase reacts with methanol under supercritical conditions to synthe-


size biodiesel (methyl ester). At the end, unreacted methanol and water
is removed to get purified biodiesel as described in Figure 9.5. Schematic
experimental setup is also described in Figure 9.6.
Supercritical method can also be performed with appropriate co-solvent
(CO2, etc.) to lower critical point of alcohol and decrease severity of super-
critical reaction condition [24].

9.2.4.2.2.2 BIOX Co-Solvent Process


BIOX co-solvent transesterification is a commercialized approach that
decreases the reaction time by enhancing the solubility of alcohol in co-
solvent that quickly reacts with triglyceride phase of VO. This process
278 Biodiesel Technology and Applications

4 1. Electric furnace
3 2. Autoclave
3. Temperature control monitor
4. Pressure control monitor
5. Biodiesel exit valve
6. Condenser
5 7. Biodiesel collector vessel
6

2
7
1

Figure 9.6 Supercritical transesterification experimental setup.

requires low temperature. Methyl tert-butyl ether (MTBE) and tetrahydro-


furan (THF) are utilized as recycled inert co-solvent to make a one-phase
system by solubilizing alcohol in it, and biodiesel is synthesized in very
short time of 5–10 min. Mostly, THF are preferred because its boiling point
is closer to methanol boiling point. As no catalyst was used, so there is no
need of catalyst removal from biodiesel or glycerol.

9.2.4.3 Microwave/Ultrasound-Assisted Transesterification


Transesterification can also be performed by 300-MHz to 300-GHz micro-
waves containing high-frequency radio and infrared waves. These waves
accelerate the chemical reaction between alcohol and VO to produce bio-
diesel [25], thus decreasing the reaction time from hours to few minutes.
Similarly, ultrasound waves generate expansion (−ive pressure) and com-
pression (+ive pressure) waves giving eddies/currents that increase the heat
and mass transfer in the VO and alcohol mixture. Ultrasound waves vary
between low frequency range of 20–100 kHz to high-frequency range of 2–10
MHz [26]. Both of these processes are efficient, economical, and time saving
because these require less amount of catalyst. Figures 9.7a and b represent
the experimental setup of microwave and ultrasound transesterification.
Table 9.2 shows the comparison between different biodiesel production
methods.

9.2.4.4 Variables Affecting Transesterification Reaction


There are various variables that influence the biodiesel yield and quality
but critical variables are moisture and FFA contents, oil-to-alcohol molar
ratios, type of alcohol and catalysts, reaction time, and reaction tempera-
ture [27]. All these variables have great influence on reaction rate, biodiesel
yield, and its conversion efficiency.
Methods for Biodiesel Production 279

Transducer
Water out

Microwave Water in
Oven
Ultrasonic generator
Ultrasonic
Power setting prob
Reactor Al-foil
Time setting
Magnetic Glass
stir bar beaker
Magnetic stirrer
(a) (b)

Figure 9.7 (a) Microwave assisted transesterification. (b) Ultrasonic assisted


transesterification process experimental setup.

Table 9.2 Comparison between biodiesel production methods [7].


Biodiesel
production
methods Advantage Disadvantage
Dilution Simple process, high Unsaturated hydrocarbon
viscosity, low chains reactivity, gum
production, and formation, high free fatty
capital costs acid (FFA), not good to use
in diesel engines directly,
poor atomization, high
viscosity, bad volatility, bad
stability, incomplete fuel
combustion, solidification
of blend occurs at cold
temperatures, injector
nozzles plugging, engine
durability of engine
reduced, higher engine
wear, higher engine
running and maintenance
costs, higher air pollution
emission, lubricating
oil thickening and
deterioration
(Continued)
280 Biodiesel Technology and Applications

Table 9.2 Comparison between biodiesel production methods [7]. (Continued)


Biodiesel
production
methods Advantage Disadvantage
Micro-emulsion Low viscosity biodiesel Heavy carbon deposition,
formation with incomplete combustion,
no by-product, lubricating oil degradation
single phase, lower and thickening, sticking of
nitrogen oxide injector needle
emissions, good
atomization of
biodiesel
Pyrolysis Simple process, good Require complex and
for hydro-processing expensive equipment,
industry, useful high temperature, high
by-products like production cost, short-
syngas is produced, chain molecules are
No-pollution, produced in biodiesel
biodiesel has that are much similar to
satisfactory gasoline than diesel fuel,
physiochemical low purity of biodiesel
properties
Transesterification The most preferable Vegetable oils with low
method for biodiesel free fatty acid and very
production, low less water content are
cost, Simple required (with base
equipment, suitable catalyst), reversible side
for industrialized reactions also occurred,
production, high reaction time,
high conversion obtained biodiesel must
yield, glycerol as be neutralized and
by-product can washed. Non-catalytic
be transformed transesterification required
into value- complex equipment due
added products, to high temperatures and
satisfactory pressure.
viscosities, biodiesel
properties are
similar to diesel fuel
(Continued)
Methods for Biodiesel Production 281

Table 9.2 Comparison between biodiesel production methods [7]. (Continued)


Biodiesel
production
methods Advantage Disadvantage
Supercritical Good adaptability, Equipment cost is high; high
methanol no catalyst, high energy consumption, high
conversion yield, temperature, and pressure
very less reaction are required
time

For transesterification process, moisture and FFA content should be


very low; otherwise, alkali-type catalyst will be consumed in neutralizing
FFAs and moisture will produce soap and frothing that will make the bio-
diesel and glycerol separation difficult. Transesterification is a reversible
process so higher alcohol-to-oil ratio is recommended than stoichiometric
ratio so that high yield of biodiesel can be achieved in less time. Mostly,
in catalytic transesterification, 1:6 oil-to-alcohol ratio is preferred, and in
non-catalytic transesterification process, 1:40 oil-to-alcohol is preferred,
but it can vary according to type of VO.
Among different type of alcohols, methanol is preferred due to its low
cost and low boiling point.
Many different types of catalysts (acidic, alkali, enzyme, homogeneous,
heterogeneous) are available but these are selected according to cost, avail-
ability, reaction condition, and type of oil. Concentrations of catalysts are
also important but their optimum range should be determined because
sometime higher concentrations have no or negative impact on biodiesel
synthesis.
Conventional transesterification takes more time (120 min), but time
can be minimized by using advanced equipment as used in microwave,
ultrasound or supercritical methods, etc. As it is reversible process, so it is
necessary to suitable biodiesel production time.
Similarly, transesterification occurred on different temperatures that
influenced the biodiesel yield. In case of conventional transesterification
temperature should be sustained below boiling point of alcohol. But, increas-
ing of temperatures in supercritical process also increases the biodiesel yield.
Optimum values and quantity of different variables should be determined
to get maximum yield of biodiesel form various feedstock having different
composition so behave differently. Production optimization for biodiesel
synthesis form different VOs should be carried out by varying types/values
and quantities of all variables involve in transesterification process.
282 Biodiesel Technology and Applications

References
1. Go, A.W., et al., Developments in in-situ (trans) esterification for biodiesel
production: A critical review. Renewable and Sustainable Energy Reviews, 60:
p. 284–305, 2016.
2. Atabani, A.E., et al., A comprehensive review on biodiesel as an alternative
energy resource and its characteristics. Renewable and Sustainable Energy
Reviews, 16(4): p. 2070–2093, 2012.
3. Pramanik, K., Properties and use of jatropha curcas oil and diesel fuel blends
in compression ignition engine. Renewable Energy, 28(2): p. 239–248, 2003.
4. Balat, M. and H. Balat, Progress in biodiesel processing. Applied Energy,
87(6): p. 1815–1835, 2010.
5. Ma, F. and M.A. Hanna, Biodiesel production: a review. Bioresource
Technology, 70(1): p. 1–15, 1999.
6. Ziejewski, M., et al., Diesel engine evaluation of a nonionic sunflower
oil-aqueous ethanol microemulsion. Journal of the American Oil Chemists’
Society, 61(10): p. 1620–1626, 1984.
7. Lin, L., et al., Opportunities and challenges for biodiesel fuel. Applied Energy,
88(4): p. 1020–1031, 2011.
8. Lima, D.G., et al., Diesel-like fuel obtained by pyrolysis of vegetable oils.
Journal of Analytical and Applied Pyrolysis, 71(2): p. 987–996, 2004.
9. Demirbaş, A. and H. Kara, New Options for Conversion of Vegetable Oils
to Alternative Fuels. Energy Sources, Part A: Recovery, Utilization, and
Environmental Effects, 28(7): p. 619–626, 2006.
10. Sharma, Y.C., B. Singh, and S.N. Upadhyay, Advancements in development
and characterization of biodiesel: A review. Fuel, 87(12): p. 2355–2373, 2008.
11. Mujtaba, M.A., et al., Critical review on sesame seed oil and its methyl ester
on cold flow and oxidation stability. Energy Reports, 6: p. 40–54, 2020.
12. Kasim, F.H., A.P. Harvey, and R. Zakaria, Biodiesel production by in situ
transesterification. Biofuels, 1(2): p. 355–365, 2010.
13. Park, J., et al., Wet in situ transesterification of microalgae using ethyl acetate
as a co-solvent and reactant. Bioresource Technology, 230: p. 8–14, 2017.
14. Lee, A.F., et al., Heterogeneous catalysis for sustainable biodiesel production
via esterification and transesterification. Chemical Society Reviews, 43(22): p.
7887–7916, 2014.
15. Chongkhong, S., et al., Biodiesel production by esterification of palm fatty
acid distillate. Biomass and Bioenergy, 31(8): p. 563–568, 2007.
16. Karmakar, A., S. Karmakar, and S. Mukherjee, Properties of various plants
and animals feedstocks for biodiesel production. Bioresource Technology,
101(19): p. 7201–7210, 2010.
17. Kombe, G.G., et al., Pre-Treatment of High Free Fatty Acids Oils by Chemical
Re-Esterification for Biodiesel Production: A Review. Advances in Chemical
Engineering and Science, Vol.03No.04: p. 6, 2013.
Methods for Biodiesel Production 283

18. Ramadhas, A.S., S. Jayaraj, and C. Muraleedharan, Biodiesel production


from high FFA rubber seed oil. Fuel, 84(4): p. 335–340, 2005.
19. Dorado, M.P., An Alkali-Catalyzed Transesterification Process for High Free
Fatty Acid Waste Oils. Transactions of the ASAE, v. 45(no. 3): p. 525–529,
2002.
20. M. Gul, et al., A review: Role of fatty acids composition in characterizing
potential feedstock for sustainable green-lubricant by advance transesteri-
fication process and it’s Global as well as Pakistani prospective. BioEnergy
Research, 2019.
21. Tabatabaei, M., et al., Reactor technologies for biodiesel production and pro-
cessing: A review. Progress in Energy and Combustion Science, 74: p. 239–303,
2019.
22. Tan, T., et al., Biodiesel production with immobilized lipase: A review.
Biotechnology Advances, 28(5): p. 628–634, 2010.
23. Demirbas, A., Progress and recent trends in biodiesel fuels. Energy Conversion
and Management, 50(1): p. 14–34, 2009.
24. Han, H., W. Cao, and J. Zhang, Preparation of biodiesel from soybean oil
using supercritical methanol and CO2 as co-solvent. Process Biochemistry,
40(9): p. 3148–3151, 2005.
25. Lidström, P., et al., Microwave assisted organic synthesis—a review.
Tetrahedron, 57(45): p. 9225–9283, 2001.
26. Veljković, V.B., J.M. Avramović, and O.S. Stamenković, Biodiesel production
by ultrasound-assisted transesterification: State of the art and the perspec-
tives. Renewable and Sustainable Energy Reviews, 16(2): p. 1193–1209, 2012.
27. Meher, L.C., D. Vidya Sagar, and S.N. Naik, Technical aspects of biodiesel
production by transesterification—a review. Renewable and Sustainable
Energy Reviews, 10(3): p. 248–268, 2006.
10
Non-Edible Feedstock for
Biodiesel Production
Chikati Roick1, Kabir Opeyemi Otun2, Nkazi Diankanua1
and Gorimbo Joshua2*

School of Chemical and Metallurgical Engineering, Faculty of Engineering and the


1

Built Environment, University of the Witwatersrand, Johannesburg, South Africa


2
Institute for the Development of Energy for African Sustainability (IDEAS),
University of South Africa’s College of Science, Engineering and Technology,
Florida, South Africa

Abstract
Continuous increase in the world’s population and high energy demand together
with environmental concerns have called for a sustainable and renewable substi-
tute for fossil fuels. Of all the existing solutions, biodiesel which can be gener-
ated from either edible and non-edible sources and these resources have proven
to be an appropriate alternative due to its renewable, low toxic, and environmen-
tally friendly nature. However, the use of non-edible feedstocks can be certain
to be a sustainable source of biodiesel production because they can be grown on
­abandoned/wasteland, where they do not have competition with food crops, they
are relatively cost-effective and produce a similar and sometimes higher yield and
fuel properties as the edible feedstocks. This chapter is, therefore, on the poten-
tials of non-edible feedstocks such as non-edible vegetable oils, waste cooking oil,
waste animal fats, and microalgae for biodiesel synthesis. Among the highlights of
this chapter are the reports relevant to global warming and climate change, mod-
ern technology for biodiesel synthesis from non-edible sources including transes-
terification, fuel properties of the biodiesel produced from non-edible feedstocks,
the economic benefits, as well as the environmental concerns. It can be concluded
from this chapter that non-edible feedstocks are promising for biodiesel industri-
alization as they meet the worldwide internationally recognized standards.

*Corresponding author: joshuagorimbo@gmail.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (285–310) © 2021 Scrivener Publishing LLC

285
286 Biodiesel Technology and Applications

Keywords: Biodiesel, renewable energy, non-edible feedstocks, global warming,


transesterification, climate change

List of Abbreviations
IMF International Monetary Fund
MDG Millennium Development Goals
AREI Africa Renewable Energy Initiative
SADCC Southern African Development Community
GHG Greenhouse gases
FAME Fatty acid methyl esters
WCO Waste cooking oil
ASTM American Standards for Testing Materials

10.1 Introduction
The 21st century is plagued with series of challenges that cause degradation
of our soils, freshwater, oceans, forests, and biodiversity. The scourge of cli-
mate change is also not helping the matter as it is even adding more pres-
sure on the resources we rely upon by aggravating the risks connected with
disasters, such as droughts and floods. Although, intensive efforts at all lev-
els of governance have contributed greatly in the campaign against climate
change. Yet, global warming continues to threaten the ecosystem and the
standard of living around the world, and the solutions to mitigate climate
change to barest minimum seem not adequate. The major contributor to
this is the greenhouse gas emission which contain majorly carbon dioxide.
While mitigation policies can help in no small way to mitigate the risks
associated global warming caused by human (anthropogenic), increasing
the capacity of carbon sinks via reforestation is another great measure. For
instance, several approaches have already been adopted to mitigate climate
change these include energy use reduction by increasing energy efficiency,
using alternative energy source to substitute petroleum-based fuels and
removal of carbon dioxide (CO2) from the atmosphere.
Biofuel, such as biodiesel and bioethanol, has no significant CO2 con-
tribution to the accumulation of greenhouse gas emission and, hence, can
serve as a viable solution to climate change. They are defined as liquid or
gaseous fuels which can be sourced from biomass and used chiefly in the
transportation sector. As a key replacement to petroleum, biodiesel is con-
sidered sustainable and environmentally benign which has elicited a surge
Non-Edible Feedstock for Biodiesel Production 287

of interest from researchers across the globe in the last few years. Biodiesels
are esters of long chain fatty acids that range in size from C12 to C24 [1, 2]. By
using biofuel produced from biomass for transportation, we can help pro-
mote the solution to meeting ever-increasing global energy demands and
restoring natural balance of CO2 in the atmosphere. Apart from replacing
fossil fuels, the feedstocks employed in producing biofuels require CO2 to
germinate and they absorb their food from the atmosphere. So, most, if not
all the CO2 released to the atmosphere when biomass is burnt in engines is
captured again when new biomass is cultivated to produce more biofuels.
Therefore, biofuel is a very effective way of reducing these emissions. The
last few years has witnessed a rapid growth in biodiesel production to curb
the adverse effects of CO2 emissions. This chapter, therefore, focuses on the
reports relevant to global warming and climate change and modern tech-
nology for biodiesel synthesis from non-edible sources including trans-
esterification with a view of mitigating the effects of climate change. We
also discuss the fuel properties of the biodiesel produced from non-edible
sources, their economic benefits, as well as the environmental concerns.

10.2 Reports Relevant to Global Warming


and Renewable Energy
Renewables may hold the key to a clean energy future. The main goal of
climate change mitigation as contained in the Paris Agreement is to limit
the world-wide average temperature increase to below 1.5°C above pre-
industrial levels. To meet this goal, the final energy consumption must
have a share of renewable energy that will increase from 19% in 2017 to
65% by 2050. The last few years has witnessed a surge of interests from
different nations implementing national or regional plans to mitigate the
adverse impact of carbon dioxide (CO2) emissions. Fossil fuels are now
considered as an unsustainable source of energy due to contamination of
the environment and causing climate changes. Some repercussions of cli-
mate change include new weather patterns that continue to exist for con-
siderable amount of time. Renewable energy as a source got attention due
to strict environmental regulations. Technologies for renewable-energy
production are comparatively cheap and sustainable than the existing coal
generation [3]. According to International Monetary Fund (IMF), some
governments, local and regional policymakers must bring to a resolution
whether to close down or reduce energy production from existing coal
plants en route toward decarbonization [4].
288 Biodiesel Technology and Applications

Global warming is now a pursuit central to many research laborato-


ries all over the world. Its effects are now visible and a present threat. Our
actions and commitments thus far are not good enough. International and
regional organizations have been pivotal in awareness and developing pol-
icies to curb climate change. For instance, UN is suggesting tax structures
and fiscal blueprint to put a stop emission of carbon from fossil fuels.
Climate related monitoring reports and agreements such as the IMF
[4], Paris Agreement [5], Copenhagen accord [6], United Nations, General
Assembly [7], Africa Renewable Energy Initiative (AREI) [8], Renewable
Energy and Energy Efficiency Strategy and Action Plan for the Southern
African Development Community [9], Millennium Development Goals
(MDG) [10], Kyoto Protocol [11] all aim at stabilizing GHG concentra-
tions in the atmosphere. Some of the objectives of the selected agreement
and reports, and in response to these, some countries have put through
comprehensive governmental climate action schemes to control climate
change. Up to the present time, these commitments are still not sufficient
to achieve the agreed temperature objectives as per the Paris Agreement;
nevertheless, the mentioned agreement draws up the way to further action.
Climate researcher and technologist proclaim that carbon emissions must
be rapidly reduced with the objective of achieving a net zero by 2050, or
earlier [12].

10.3 Biofuels as an Alternative Energy Source


Biomass is defined as the plant or animal matter that can be utilized for
energy production such as electricity or heat. Biofuels are categorized as
natural, primary and secondary biofuels. Generally, natural biofuels are
extracted from organic materials, in particular vegetables, waste from ani-
mals, and landfill gas. Primary biofuels are fuels from wood usually used
for cooking and heating in most rural areas and, in some case, electricity
production. Secondary biofuels refers to bioethanol and biodiesel that are
sourced from processing biomass. The secondary biofuels are additionally
grouped into first-, second-, and third-generation biofuels based on their
source, development levels, and the processing technology as shown in
Figure 10.1.

10.3.1 First-Generation Biofuels


This generation of biofuels is those fuels originating from what human
beings use as food for example sugar, starch, vegetable oil, animal fat, etc.
Non-Edible Feedstock for Biodiesel Production 289

1st Generation 2nd Generation 3rd Generation 4th Generation


• Edible biomass • Non-edible • Algal biomass Breakthrough
• E.g. sugar beet, biomass • Microalgae
corn, wheat, etc. • E.g. Wood, straw, Genetically
• Macroalgae
grass, waste, etc. engineered
crops and
algae

Figure 10.1 Biofuels production sources adapted from [13].

Some examples of the most common types of first-generation biofuels are


as follows:

• vegetable oil,
• biodiesel,
• syngas,
• biogas, and
• bio-alcohols.

Despite of these first-generation fuels possessing potential to be used


in generating carbon neutral biofuels, this approach could have notables
economic and political concerns as grain and food prices will go up [13].
Martin (2010) looked at the conflicts of first-generation biofuels and the
implications it has on developed and developing economies [14].
The major downside of using the first-generation fuels is the use of ara-
ble agricultural lands and as a consequence shortage of land for animal
and human food production rendering the food-vs-fuel quandary a moot
point.

10.3.2 Second-Generation Biofuels


Biofuels under this category are also referred to as advanced biofu-
els, and these fuels originate from various types of non-food biomass.
Lignocellulosic biomasses (such as agricultural waste, and woody crops)
are the major feedstocks in the production of second-generation bio-
fuels. Second-generation biofuel feedstock can also be the nonedible
290 Biodiesel Technology and Applications

by-product of food crops such as the corn husks or Panicum virgatum


(wheat straw) and other variety of plants that are not edible, for instance,
switch grass. The technology used allows the plant cellulose and lignin
to be separated in order for cellulose to be fermented into alcohol. These
biofuels can be sourced from a variety of biomass. A considerable number
of companies and research groups are venturing into cost effective pro-
cessing for producing second-generation biofuels such as ExxonMobil
which is pursuing research to determine how these biofuels may best fit
into future energy.

10.3.3 Third-Generation Biofuels


Beyond second-generation biofuels, engineers and scientists have also
looked at the so-called third-generation biofuels, largely derived from
algae (microalgae and cyanobacteria) [15, 16]. Different from the first-
and second-generation derived biofuels, algae poses no threat to food
security. Algae can be grown in brackish water and in some cases sea-
waters making it human friendly source of biofuel, as it does not put an
enormous strain on the freshwater. More so, no agricultural land is used
to grow the algae; therefore, little or no competition exist land for food
production. Algae are defined as photosynthetic organisms that convert
atmospheric CO2 and H2O to oil using light from the sun, and it is the oils
that gets converted to biodiesel portion via a series of processes [17]. Also,
of interest is the water hyacinth (scientific name Eichhornia crassipes)
which has been reported to have a high biomass yield and can grow expo-
nentially. These too can be a potential feedstock for biofuel production
and related bioactive ingredients [18]. The advantages of both algae and
the water hyacinth is the fact that they are aquaculture, that they repro-
duce exponentially, and can be cultivated in salty water giving biomass
of higher energy content. The amount of oil produced in both systems,
however, is a function of parameters such as sunlight exposure times and
atmospheric CO2 levels.

10.4 Benefits of Using Biodiesel


The advantages of utilizing biodiesel in internal combustion engines are
overwhelming and relate to its broader applications based on the following
highlights:
Non-Edible Feedstock for Biodiesel Production 291

a. Biodiesel is less toxic and safer to handle than petroleum-


based diesel due to its relative high flash point.
b. Biodiesel is a sustainable and renewable source of energy
with zero harmful emission.
c. The simplicity and low capital cost of biodiesel production
enhances its wider distribution and use commercially.
d. The starting materials for biodiesel synthesis are organic and
can be sourced locally. In fact, more biodiesel can be reaped
from non-edible and waste oils nowadays.
e. In terms of energy conversion and power outputs, generally,
diesels engines are more efficient than petrol engines.
f. Most countries have adopted biodiesel as a means of foreign
exchange. The availability of raw materials for biodiesel pro-
duction has encouraged local farmers to embark on mass
plantation which in turn has created more job opportuni-
ties and serves and increasing the production of foreign
exchange.
g. Biodiesels are clean burning fuels with low emission of
greenhouse gases [14, 19].

10.5 Technologies of Biodiesel Production


From Non-Edible Feedstock
Generally, four main methods have been developed to produce biodiesel to
fit in into our conventional engines without any amendments [20]. These
are as follows.
1. Direct use or blending (Dilution): This method is the oldest and
easiest method of biodiesel production. Vegetable/animal oils are not
advisable to use as is in engines because of its high viscosity which
may trigger other challenges. To circumvent these challenges, vege-
table oils can be blended with diesel oils in some limited ratio. Blend
ratios which results in optimum performances in diesel driven engines
are generally of less viscosity which is achieved by mixing considerable
proportion of petrodiesel (preferably 80%) with lower percentages of
renewable vegetable/animal oils (preferably 20%). Carbon deposition in
engine cylinder and improper combustion are some of the demerits of
this process.
292 Biodiesel Technology and Applications

H2C OH H3C OCOR1


H2C OCOR1
base OCOR2
+ HC OH + H3C
3 CH3OH
HC OCOR2
H3C OCOR3
H2C OCOR3 H2C OH

Triglyceride Methanol Glycerol Fatty acid methyl esters (Biodiesel)

Figure 10.2 Transesterification reaction for biodiesel production [24].

2. Micro emulsification: This is basically prepared by mixing alcohol,


surfactants, vegetable/animal oil, diesel fuel, and cetane improver in dif-
ferent fractions. This process is easy but produces a less volatile, less stable,
and highly viscous mixture, which is the main drawback of the process.
The challenges associated with the viscosity of vegetable oils can be cur-
tailed by making of microemulsion. It is the solubilization of oils with the
aid alcoholic solvents and surfactants [21].
3. Thermal cracking or pyrolysis: This process involves heating of veg-
etable or animal oils at high temperature (usually 300 to 1,300°C) in the
presence or absence of a catalyst and absence of oxygen. It is simple and
produces less waste and less emission. It is very effective when compared
with other cracking processes. High installation cost and oxygen removal
during pyrolysis are some of the challenges encountered by this process
[22].
4. Transesterification: This is considered as the reaction in which
triglycerides (oils/organic fats) are reacted with suitable alcohol in the
presence of acidic/alkaline catalyst to yield fatty acid esters (biodiesel)
and glycerol. This process produces biodiesels that have comparable
features with diesel fuels and is the most promising route for commer-
cial production of biodiesel from the economic point of view [23]. The
transesterification process is the most convenient method because it is
cheap and simple while the choice of feedstock is key since it contrib-
utes a significant seventy percent of the total production cost from edible
and non-edible sources [20]. The transesterification reaction is shown in
Figure 10.2.

10.6 Biodiesel Production by Transesterification


Among the methods previously discussed, transesterification is the com-
monest and the most preferred method to produce biodiesel because of
Non-Edible Feedstock for Biodiesel Production 293

its simplicity and it has been widely employed in industry to convert oils
into biodiesel. It also affords the use of a wide range of feedstocks to
produce a biodiesel that greatly resembles conventional diesel in terms
of quality.
First-generation biodiesel is produced via transesterification of tri-
glycerides from vegetable oils, while second-generation biodiesels are
obtained from feedstocks like lignocellulose and non-edible triglycerides.
Transesterification involves chemically converting various types of oils
into fatty acid methyl esters (FAME) and glycerol in the presence of cata-
lysts [25]. FAME, commonly known as biodiesel, is the monoalkyl esters of
long chain fatty acids. Table 10.2 shows the major free fatty acids present
in biodiesel production. Type of catalyst, nature of raw material, reaction
conditions, type of solvent, reactor type, and solvent to oil ratio are some
of the key variables required to obtain optimum yield for biodiesel produc-
tion via transesterification process [26]. Of all these factors, catalyst type
and type of feedstocks are the key parameters for effective biodiesel pro-
duction because they both determine the price of biodiesel production to
a greater extent. These catalysts are divided into two major types, homog-
enous or heterogeneous:

a) Homogeneous catalyst: In homogenous catalysis for trans-


esterification, alkaline or basic catalysts are more preferred
to the acid catalysts choice because of their high conver-
sion rate in relatively short time. In addition, basic catalysts
do not corrode easily with industrial equipment. The most
common examples of homogeneous catalysts for this reac-
tion are KOH, and NaOH (alkaline) and H2SO4 and H3PO4
(acidic). The major demerit of the basic catalysts is their
high sensitivity to fatty acids contained in the feedstock.
Acid catalysts are also well suited for the transesterification
reaction of highly fatty materials present in fats and oil.
However, they are slower than base catalyzed reaction and
require high temperature and pressure and high amount of
alcohols for the esterification reaction. In general, homoge-
nous catalysts is majorly at disadvantage because of the sep-
aration of the catalysts from the products and inability to
reuse. Removal of catalysts involves several washing steps
and activation which increases the total cost of production
[27].
294 Biodiesel Technology and Applications

b) Heterogeneous catalysts: Heterogeneous catalysts which


offers to replace homogeneous catalysts are usually solid
materials that can be obtained from renewable materials
following pyrolysis at high temperature between 500°C and
900°C [28]. They are better than homogeneous catalysts in
the sense that they are easy to separate from the products can
be reused, and no side reaction. Moreover, the biodiesel pro-
duced from these catalysts yielded about 96% FAME content
[28]. The most frequently used heterogeneous basic catalysts
are oxides and carbonates of alkali and alkaline-earth metals.
c) Enzymatic transesterification: Diesel production via this
enzyme-catalyzed transesterification pathway experimented
by Dhawane et al. (2018) [29] and Andrade et al. (2017) [30]
with the use of lipase enzyme and methanol [31]. It involves
hydrolysis and esterification of triglycerides (substrate) to
produce FAMEs (diesel). This process is highly promising
because of their high conversion rate, ease of separation
from the product, good glycerol recovery and high yield
when compared with the homogenous and heterogeneous
catalysts. Nonetheless, biodiesel synthesis via enzymatic
transesterification is expensive because of lipase manufac-
turing and other intricate stages involved in cellular enzyme
synthesis and separation. Table 10.1 gives some of the com-
mon free fatty acids in biodiesel production with Figure
10.3 showing a flow chart of biodiesel production process
through transesterification process.

Table 10.1 The common free fatty acids in biodiesel production [32].
Entry Name of fatty acid Chemical name Structure
1 Palmitic Hexadecanoic acid C16:0
2 Stearic Octadecanoic acid C18:0
3 Oleic Octadecenoic acid C18:1
4 Linoleic Octadecadienoic acid C18:2
5 Linolenic Cis-9, cis-12, cis-15- C18:3
octadecatrienoic acid
Non-Edible Feedstock for Biodiesel Production 295

Alcohol
Biodiesel
+
catalyst

Drying
Washing 350-400ºC,
Transesterification
(oil/alcohol ratio) To remove 20 h
unreacting raw To remove
t
en materials contaminants
atm
e
etr l,
Pr ysica ical,
Feedstock Ph olog Glycerol
bi

Figure 10.3 Flow chart of biodiesel production process through transesterification


process.

10.7 Non-Edible Feedstocks for Biodiesel Production


Mostly, biodiesels can be prepared from two major sources (feedstocks),
namely, edible and non-edible feedstocks as shown in Figure 10.4. In 2017,
the worldwide biodiesel production was credited to palm oil (31%), soy-
bean oil (27%), rapeseed oil (20%), waste cooking oil (WCO) (10%), ani-
mal fats (7%), and others (5%) [33]. Although, the use edible feedstocks
may be the cheapest feedstock for biodiesel production for the future, but it
is not sustainable to meet the increasing demand for biodiesel [34]. Hence,
the use of non-edible feedstocks is a sustainable alternative in the sense

• Corn, soybean,
Edible sunflower, etc

Biodiesel
Feedstocks

• Vegetable oil
Non • Waste animal oil
• Waste cooking oil
edible
• Algae, lignocellulose,
etc

Figure 10.4 Biodiesel production from edible and non-edible feedstocks.


296 Biodiesel Technology and Applications

that it does not compete with the food crops for limited plantation region
and has the potentials of reclaiming the wastelands [35]. Based on the data
available from the literature, WCO, animal oil, non-edible vegetable oil,
waste animal oil, and alga oil are the most promising alternatives for edible
oils and are elucidated as follows.

10.7.1 Non-Edible Vegetable Oils


The toxic chemicals present in non-edible vegetable oils make them unus-
able for human consumption. In addition, the threat posed by both the
food security and economic issues also make non-edible vegetable oils
more preferred than edible vegetable oils in the sense that it does not
encourage serious competition over land for food production and for the
food supply chain.
i) Jatropha is a tropical and drought-resistant plant that can thrive in an
abandoned and fallowed farmland [20]. It has been recognized as one of
the suitable non-edible sources for biodiesel production due to its phys-
iochemical properties. It is a rich source of hydrocarbons and has picked
the interest of researchers all over the world because of the use of its seed
oils as alternative feedstock for biodiesel production. There are about
20%–60% oil in the seeds of jatropha plant and contain mainly unsaturated
fatty acids likes oleic (42%) as the major component, followed by linoleic
(35%) and smaller percentage of palmitic (14%) and stearic acid (6%) [36].
Recently, Salar-Garcia et al. (2016) obtained 99.5% biodiesel production
yield from Jatropha oil and 100% conversion rate of triglycerides at 325°C
in 90 min [37].
ii) Karanja (botanical name Pongamia pinnata) is another potential
non-edible vegetable oil feedstock that can be used to produce biodiesel.
Countries like Southeast Asia, US, Australia, New Zealand, India, and
China are the main producers of this plant [38]. Interestingly, these trees
can grow in many places including roadsides, canals, and boundaries of
farmlands. The oil content of its seed ranges between 30% and 40%, and
can yield 97% FAME (biodiesel) via transesterification process at a tem-
perature of 65°C using 1 wt% of KOH and alcohol to molar ratio of 6:1
in 2 h. Karanja oil contains stearic acid (up to 8.9%), linoleic acid (up to
18.3%), and oleic acid (up to 71.3%) [39].
iii) Linseed which is also referred to as Linum usitassimum is another
potential non-edible feedstock for biodiesel production. Linseed can
produce oil up to 47% under optimum conditions [40]. The main fatty
acids which are mainly found in linseed oils are linolenic (up to 51%),
oleic acid (up to 21%), linoleic acid (up to 15%), and small amounts of
Non-Edible Feedstock for Biodiesel Production 297

Table 10.2 Current research on non-edible vegetable oils as low-cost feedstocks


for biodiesel production via transesterification.
Non-edible
vegetable Methanol/ Temperature Time Yield
source Catalyst oil (°C) (h) (%) Ref.

Jatropha KOH, 1 wt% 6:1 60 1 96.1 [42]


curca

Linseed KOH, 1 wt% 9:1 60 2 95.5 [43]

Karanja γ-Alumina 1:9 50 0.83 69.3 [44]

Neem Cu/ZnO 10:1 55 1 74% [45]


nanocatalyst,
10 wt%

Rapeseed 6 wt% Na/FAP 10:1 120 8 98.5 [46]

Cotton seed KOH, 1 wt% 6:1 25 5 min - [39]

Castor KOH, 1.25 wt% 12:1 60 1 95 [47]

Jojoba KOH, 1 wt% 6:1 50 1 78.5 [48]

palmitic acid (6%) and stearic acid (5%) [41]. Other non-edible vegeta-
ble feedstocks include but not limited to neem, jojoba, desert date, sea
mango, rubber, tobacco, and castor, among others. Recent studies on
non-edible vegetable oils as sources for the production of biodiesel are
shown in Table 10.2.

10.7.2 Waste Cooking Oil


WCOs are the left over oils after a deep-frying procedure and the used oil
can be a suitable feedstock for biodiesel production. According to Loizides
et al. (2019), about 16.5 million tons of WCO is produced annually [49].
The lower solubility of these oils makes their disposal very challenging.
Therefore, the conversion of WCO into biodiesel will greatly help to mit-
igate the disposal problem while helping to solve the energy crisis. WCO
can be categorized according to the source such as household, food indus-
try, and non-food industry. Acid, base, and enzymes can be used to pro-
mote the conversion of WCO into biodiesel. Recently, it was reported that
demonstrated that sulfonated catalysts can effectively convert WCO into
biodiesel [50]. The FAME content obtained from pyrolyzed rice straw was
around 97.7%, conversion efficiency was 90.4%, and 91.1% free fatty acid
conversion rate with 10 wt% catalyst, methanol-to-oil ratio of 20:1 at 70°C
298 Biodiesel Technology and Applications

Table 10.3 Recent studies on waste cooking oils as feedstocks for the production
of biodiesel via transesterification reaction.
Loading Temperature
Catalyst (wt%) Alcohol Oil/alcohol (°C) Time Yield (%) Ref.

KOH/Clinoptilolite 9.1 MeOH 2.25:1 61 13.4 min 97.5 [52]

ZnAl2O4 5 MeOH 1:18 100 3h 94.9 [53]

RS-SO3H 10 MeOH 1:20 70 6h 97.7 [50]

CaO 4 MeOH 1:8 65 75 min 98.2 [27]

Kaolinite/K+
15 MeOH 1:14 70 3h 94.8 [54]

for 6 h [50]. Similarly, Maneerung and co-workers produced biodiesel with


90% FAME yield using calcined chicken manure as a precursor of CaO cat-
alyst [51]. The result was obtained under optimum conditions at 7.5 wt%
catalyst, the obtained methanol-to-oil ratio was 1:15 at 65°C [51]. Recent
studies on WCOs as feedstocks for the production of biodiesel via transes-
terification reaction are tabulated below (see Table 10.3).

10.7.3 Algal Oil


Algal species, which are regarded as the third-generation feedstock, are
a promising non-edible source of biodiesel. Algae can grow in both nat-
ural and artificial environment, and they are more economically viable
than edible oils. There are about 44,000 species of microalgae and a careful
selection of an appropriate algal strain is crucial to the overall performance
of biodiesel production [55]. Different microalgae including Chlorella sp.,
Chlorella vulgaris and Dunaliella salina algae, among others, have been
reported to be valuable in biodiesel production (see also Table 10.4). Algal
biodiesel has no sulfur and performs just like petroleum diesel, while min-
imizing the emission of harmful gases like CO, SOx, and NOx. Algal oils
have been one of the best choices for researchers as algae provides more
oil yield per area of arable land. To make microalgae more economically
viable, attention should be shifted to minimizing the cost of feedstocks,
inducing the lipid content, improving extraction efficiency, increasing the
area productivity, and helps to convert of algal lipids to biodiesel. Of all
these, the feedstock alone equals about 80% of the total production cost,
which can be drastically lowered by the use of waste and less costly raw
materials [56].
Transesterification is the most preferred process to produce biodiesel
from algal oil. This process can be economical and time-saving if key factors
Non-Edible Feedstock for Biodiesel Production 299

Table 10.4 Recent studies on algal oils as feedstocks for biodiesel production
via transesterification method.
Loading Temperature
Feedstock Catalyst (wt%) Oil/MeOH (°C) Time (min) Yield (%) Ref.

Chlorella NaOH 38 1:600 60 10 96 [57]


Vulgaris

Algal lipids Ca(OCH3)2 3 1:30 80 150 99 [58]

Algal oil CaO 1.25 1:9 55 - 96.3 [55]

Algal oil Waste clay/ 3.5 1:9 60 240 97.4 [59]


ZnO

Algal lipid Immobilized 1.39 1:19 70 180 92.03 [60]


C. rugiza
lipase

like time of reaction, catalyst amount, and oil/alcohol ratio are optimized.
Narula et al. (2017) produced 88.9% yield of biodiesel from algal oil via
transesterification reaction by using CaO and CaO.Al2O3 as catalysts [55].

10.7.4 Waste Animal Fat/Oil


Animal fat is classified under the third-generation feedstock for biodiesel
production. The co-product of meat and fishery is animal fat. Examples
include beef tallow or mutton and yellow grease. These feedstocks, unlike
edible oils, are important for their food security, economic, and environ-
mental benefits. Presently, the co-products obtained from animal fats have a
very low market price and hence can be instrumental as an important source
of biodiesel production (see Table 10.5). In addition, most animal fats are
no longer allowed to be used as food any longer due to the many infections
that ravaged animals. For instance, tallow from diseased livestock is a key
feedstock for biodiesel production for the same reasons above. However,
inconsistent supply poses a great challenge for all these feedstocks because
animal fat is not only produced for biodiesel production [61].

10.8 Fuel Properties of Biodiesel Obtained


From Non-Edible Feedstock
Different factors affect the fuel properties of biodiesel obtained from
non-edible feedstock. This includes, but not limited to, quality of raw
300 Biodiesel Technology and Applications

Table 10.5 Recent studies on waste animal fat/oil as feedstocks for production of
production via transesterification process.
Waste animal Loading Temperature Time
fats/oils Catalyst (wt%) Oil/MeOH (°C) (min) Yield (%) Ref.

Animal tallow NaOH - 1:6 60 180 - [62]


oil

Mutton fat KOH/MgO 4 1:22 65 20 98 [63]

Fish oil CaO-Ca3Al2O6 10 1:12 54 90 96.4 [64]

Waste shark NaOH 5.9 1:6 65 60 96 [65]


liver

Waste chicken CaO/CuFe2O4 3 1:15 70 240 94.52 [66]


fat

materials and fatty acid compositions, method of production, pro-


cesses involved in refining, and final production conditions [60]. There
are different measures to classify fuel properties of biodiesel, the most
important of which are: influence of activities taking place in the engine
(such as ignition quality, calorific value, combustion of air-fuel mixture,
and formation of exhaust gas among others), low temperature prop-
erties (like cloud point, pour point, etc.), transportation and storage
properties (flash point, oxidation and hydrolytic stability, etc.), wear of
engine parts (viscosity, lubrication, etc.) [50]. In addition, these prop-
erties must meet the internationally established worldwide standards
in order to make the biodiesel useful for commercial purposes, which
are American Standards for Testing Materials (ASTM D6751) and
European (EN 14214) Standards for biodiesel fuel. In order to obtain
high quality biodiesel, as per ASTM and EN standards, optimization
of the pretreatment processes, separation of the product, and the final
purification steps should be done optimally. Additionally, the free fatty
acid contained within the oil blend to a large extent influences quality
and the properties of the fuel.
The literature reports that factors like cetane number, flash point, viscos-
ity, cloud point, and iodine value are the most important physicochemical
qualities of biodiesel and are strongly affected by the composition of FAME
[67]. A summary of the physicochemical properties of biodiesel obtained
from selected non-edible feedstock vis-à-vis the ASTM and EN standards
are given in Table 10.6.
Table 10.6 Summary of the physicochemical features of biodiesel produced from non-edible feedstocks.
Non-edible Cetane Viscosity Flash Density Cloud
feedstock number (mm2/s) point °C Kg/m3 point Standard Ref.
Waste 59 4.63 161 887 - ASTM D [25]
cooking 6751
oil
Cotton seed 54 4.06 200 850 −10 EN 14214, [39]
ASTM
D6751
Castor 80 13.75 149 927 - ASTM [47]
D6751

Jatropha 59.64 5.65 184 862 - ASTM [37]


curcas D6751
Chicken fat - 5.3 171 858 18 ASTM [66]
D6751
Chlorella sp 56.1 4.6 113 886 −2.2 ASTM [24]
D6751,
EN
Non-Edible Feedstock for Biodiesel Production
301
302 Biodiesel Technology and Applications

10.9 Advantages of Non-Edible Feedstocks


The various results obtained from different researchers across the globe
have shown that non-edible feedstock have the following advantages with
respect to biodiesel production [68].

1. Non-edible feedstocks can be cultivated in marginal land


and non-agricultural lands.
2. Non-edible sources do not contend with available agricul-
tural resources.
3. They do not threaten food security because they are inap-
propriate for human consumption due to the non-toxic
components they contain.
4. Most of the non-edible feedstocks are free from pests and
diseases.
5. Oils obtained from non-edible feedstocks are readily avail-
able, renewable, and biodegradable, with low sulfur and aro-
matic contents.
6. Non-edible feedstocks produce useful by products. For
example, the seed cakes obtained after oil expelling can
enhance soil enrichment.
7. Most non-edible feedstocks have a considerable amount
of short-chain fatty acids, which gives special features to
biodiesel.
8. They are more eco-friendly than non-edible feedstocks.
9. Non-edible feedstocks can be cultivated in wastelands that
are not ideal for human food crops.
10. They have the potentials to restore degraded lands and cre-
ate job opportunities.

10.10 Economic Importance of Biodiesel Production


Capital cost, raw materials cost, process technology, plant capacity, and
chemical costs are key economic factors in biodiesel production. Of all,
the cost of feedstocks alone is equivalent to 80% of the total production
cost, while the cost of catalyst, methanol, and utilities are equally import-
ant [69]. For instance, it costs 0.82 USD/litre to make biodiesel from palm
oil including the feedstock price of 0.73, while it takes 0.6 USD/litre to pro-
duce the same biodiesel from tallow fat including the feedstock cost of 0.4
USD [70]. Most importantly, biodiesel production helps to boost economy
Non-Edible Feedstock for Biodiesel Production 303

in both developing and developed countries by creating job opportunities


for the rural community, minimizing greenhouse gas emissions, boosting
income tax revenue, and reducing country’s over dependence on crude
oils. Recently, the use of non-edible feedstocks that do not contend with
the food crops is a reasonable way of improving biofuel production.
In terms of socio-economic effects, biodiesel, a sustainable and renew-
able energy source, is good substitutes for petroleum fuels. Hence, they can
help to reduce greenhouse gases that bring about global warming, promotes
regional development, and boosts food security and supply. Biodiesel also
have health, environmental, safety, and other benefits in addition to eco-
nomic advantages [3].

10.11 Conclusions
In this chapter, we discussed the production pathways of biodiesel from
non-edible feedstocks, with respect to their benefits, up-to-date technol-
ogy, properties of the fuel produced, the economic benefits, and the envi-
ronmental concerns which provides some key conclusions.
Different methods can be used to produce biodiesel from oils. These
include pyrolysis, dilution, micro-emulsion, and transesterification.
Among these, the most economically viable is transesterification and the
produced biodiesel compares favorably with the petroleum-based biodiesel
Choice of feedstocks is a crucial factor in the synthesis of biodiesel.
Based on the various feedstocks, non-edible oils, algal oils, waste animal
fats, and WCOs are promising sources for biodiesel production.
Production of biodiesel from non-edible sources has become attractive
because they are renewable and sustainable source that guarantees food
security, improves local economy, and reduces pollution.
Of the various costs of biodiesel production, the feedstock price is almost
80% of the total production cost. A better way of lowering the cost of bio-
diesel production is to develop technology that will make use of bye products
and makes judicious and favorable choice of the feedstocks. The biodiesel
made from non-edible sources meet internationally recognized standards.

Acknowledgments
The authors are grateful for the financial support provided by the
University of South Africa (UNISA), University of the Witwatersrand,
National Research Foundation (NRF) of South Africa, and the Institute for
304 Biodiesel Technology and Applications

the Development of Energy for African Sustainability (IDEAS) research


unit at UNISA.

References
1. G. L. Alexandrino, J. Malmborg, F. Augusto, and J. H. Christensen,
Investigating weathering in light diesel oils using comprehensive two-
dimensional gas chromatography–High resolution mass spectrometry and
pixel-based analysis: Possibilities and limitations, J. Chromatogr. A, vol. 1591,
pp. 155–161, 2019.
2. F. C. Y. Wang, Comprehensive three-dimensional gas chromatography mass
spectrometry separation of diesel, J. Chromatogr. A, vol. 1489, pp. 126–133,
2017.
3. K. Gillingham and J. H. Stock, The Cost of Reducing Greenhouse Gas
Emissions, J. Econ. Perspect., vol. 32, no. 4, pp. 53–72, 2018.
4. International Monetary Fund. Fiscal Affairs Dept, Fiscal monitor, How to
Mitigate Climate Change. 2019. https://www.imf.org/en/Publications/FM/
Issues/2019/09/12/fiscal-monitor-october-2019, (accessed 15 February
2020).
5. Paris agreement, 2015. https://unfccc.int/process-and-meetings/the-paris-
agreement/the-paris-agreement, (accessed 15 February 2020).
6. United Nations, Copenhagen accord, 2009. tps://unfccc.int/resource/
docs/2009/cop15/eng/l07.pdf, (accessed 20 February 2020).
7. T. S. United Nations, Transforming our world: the 2030 Agenda for
Sustainable Development, 2015. https://sustainabledevelopment.un.org/
post2015/transformingourworld, (accessed 20 February 2020).
8. Africa Renewable Energy Initiative, 2015. Increasing Renewable Energy
Capacity on the African Continent, https://unfccc.int/news/africa-
renewable-energy-initiative-increasing-renewable-energy-capacity-on-the-
african-continent, (accessed 20 February 2020).
9. REEESAP, Renewables Energy and Energy Efficiency Strategy and Action
Plan for the Southern African Decelopment Community, 2017. https://www.
sacreee.org/document/reeesap-southern-africa-renewable-energy-and-
energy-efficiency-strategy-and-action-plan, (accessed 20 February 2020).
10. United Nations, World Economic and Social Survey 2014 / 2015 Learning
from national policies supporting MDG implementation, 2016. https://
www.un.org/development/desa/dpad/publication/world-economic-
and-social-survey-20142015-learning-from-national-policies-supporting-
mdg-implementation/, (accessed 22 February 2020).
11. B. Ki-moon, Kyoto Protocol Reference Manual, 2008. https://unfccc.int/
resource/docs/publications/08_unfccc_kp_ref_manual.pdf, (accessed 22
February 2020).
Non-Edible Feedstock for Biodiesel Production 305

12. J. R. Millar, Z. R. Nicholls, P. Friedlingstein, and M. R. Allen, A modified


impulse-response representation of the global near-surface air tempera-
ture and atmospheric concentration response to carbon dioxide emissions,
Atmos. Chem. Phys., vol. 17, no. 11, pp. 7213–7228, 2017.
13. F. Alam, A. Date, R. Rasjidin, S. Mobin, H. Moria, and A. Baqui, Biofuel from
algae Is it a viable alternative, Procedia Eng., vol. 49, no. 49, pp. 221–227,
2012.
14. M. A. Martin, First generation biofuels compete, N. Biotechnol., vol. 27,
pp. 597–607, 2010.
15. C. R. Carere, R. Sparling, N. Cicek, and D. B. Levin, Third Generation
Biofuels via Direct Cellulose Fermentation, Int. J. Mol. Sci., vol. 9, pp. 1342–
1360, 2008.
16. S. Khan et al., Biodiesel Production From Algae to Overcome the Energy
Crisis _ Elsevier Enhanced Reader, Hayati J. Biosci., vol. 24, no. 24, pp. 163–
167, 2017.
17. G. Dragone, B. Fernandes, A. A. Vicente, and J. A. Teixeira, Third generation
biofuels from microalgae, pp. 1355–1366, 2010.
18. S. M. M. Shanab, Water Hyacinth as Non-edible Source for Biofuel
Production Water Hyacinth as Non-edible Source for Biofuel Production,
Waste and Biomass Valorization, vol. 9, no. 2, pp. 255–264, 2017.
19. N. Matsumoto, D. Sano, and M. Elder, Biofuel initiatives in Japan: Strategies,
policies, and future potential, Appl. Energy, vol. 86, no. SUPPL. 1, pp. S69–
S76, 2009.
20. H. C. Ong, A. S. Silitonga, H. H. Masjuki, T. M. I. Mahlia, W. T. Chong,
and M. H. Boosroh, Production and comparative fuel properties of biodiesel
from non-edible oils: Jatropha curcas, Sterculia foetida and Ceiba pentandra,
Energy Convers. Manag., vol. 73, pp. 245–255, 2013.
21. J. Yan and Y. Yan, Biodiesel Production and Technologies, vol. 3. Elsevier,
2017.
22. S. Rezania et al., Review on transesterification of non-edible sources for
biodiesel production with a focus on economic aspects, fuel properties and
by-product applications, Energy Convers. Manag., vol. 201, no. October,
pp. 112155, 2019.
23. F. Yaşar, Biodiesel production via waste eggshell as a low-cost heterogeneous
catalyst: Its effects on some critical fuel properties and comparison with
CaO, Fuel, vol. 255, no. April, pp. 115828, 2019.
24. S. Ghosh, S. Banerjee, and D. Das, Process intensification of biodiesel pro-
duction from Chlorella sp. MJ 11/11 by single step transesterification, Algal
Res., vol. 27, no. March, pp. 12–20, 2017.
25. M. J. Borah, A. Das, V. Das, N. Bhuyan, and D. Deka, Transesterification
of waste cooking oil for biodiesel production catalyzed by Zn substituted
waste egg shell derived CaO nanocatalyst, Fuel, vol. 242, no. May 2018,
pp. 345–354, 2019.
306 Biodiesel Technology and Applications

26. M. E. Günay, L. Türker, and N. A. Tapan, Significant parameters and tech-


nological advancements in biodiesel production systems, Fuel, vol. 250, no.
September 2018, pp. 27–41, 2019.
27. Z. L. Chung et al., Life cycle assessment of waste cooking oil for biodiesel
production using waste chicken eggshell derived CaO as catalyst via trans-
esterification, Biocatal. Agric. Biotechnol., vol. 21, no. September, p. 101317,
2019.
28. M. E. Borges and L. Díaz, Recent developments on heterogeneous catalysts
for biodiesel production by oil esterification and transesterification reac-
tions: A review, Renew. Sustain. Energy Rev., vol. 16, no. 5, pp. 2839–2849,
2012.
29. S. H. Dhawane, T. Kumar, and G. Halder, Process optimisation and para-
metric effects on synthesis of lipase immobilised carbonaceous catalyst for
conversion of rubber seed oil to biodiesel, Energy Convers. Manag., vol. 176,
no. June, pp. 55–68, 2018.
30. T. A. Andrade, M. Errico, and K. V. Christensen, Influence of the reaction
conditions on the enzyme catalyzed transesterification of castor oil: A pos-
sible step in biodiesel production, Bioresour. Technol., vol. 243, pp. 366–374,
2017.
31. F. Moazeni, Y. C. Chen, and G. Zhang, Enzymatic transesterification for bio-
diesel production from used cooking oil, a review, J. Clean. Prod., vol. 216,
pp. 117–128, 2019.
32. A. E. Atabani et al., Non-edible vegetable oils: A critical evaluation of oil
extraction, fatty acid compositions, biodiesel production, characteristics,
engine performance and emissions production, Renew. Sustain. Energy Rev.,
vol. 18, pp. 211–245, 2013.
33. A. B. Fadhil, E. T. B. Al-Tikrity, and M. A. Albadree, Biodiesel production
from mixed non-edible oils, castor seed oil and waste fish oil, Fuel, vol. 210,
no. September, pp. 721–728, 2017.
34. A. Arumugam and V. Ponnusami, Biodiesel production from Calophyllum
inophyllum oil a potential non-edible feedstock: An overview, Renew. Energy,
vol. 131, pp. 459–471, 2019.
35. B. Sajjadi, A. A. A. Raman, and H. Arandiyan, A comprehensive review
on properties of edible and non-edible vegetable oil-based biodiesel:
Composition, specifications and prediction models, Renew. Sustain. Energy
Rev., vol. 63, pp. 62–92, 2016.
36. D. Kumar, T. Das, B. S. Giri, E. R. Rene, and B. Verma, Biodiesel production
from hybrid non-edible oil using bio-support beads immobilized with lipase
from Pseudomonas cepacia, Fuel, vol. 255, no. February, p. 115801, 2019.
37. M. J. Salar-García, V. M. Ortiz-Martínez, P. Olivares-Carrillo, J. Quesada-
Medina, A. P. De Los Ríos, and F. J. Hernández-Fernández, Analysis of opti-
mal conditions for biodiesel production from Jatropha oil in supercritical
methanol: Quantification of thermal decomposition degree and analysis of
FAMEs, J. Supercrit. Fluids, vol. 112, pp. 1–6, 2016.
Non-Edible Feedstock for Biodiesel Production 307

38. R. L. Patel and C. D. Sankhavara, Biodiesel production from Karanja oil and
its use in diesel engine: A review, Renew. Sustain. Energy Rev., vol. 71, no.
April 2015, pp. 464–474, 2017.
39. X. Fan, X. Wang, and F. Chen, Ultrasonically assisted production of biodiesel
from crude cottonseed oil, Int. J. Green Energy, vol. 7, no. 2, pp. 117–127,
2010.
40. A. Demirbas, Production of biodiesel fuels from linseed oil using methanol
and ethanol in non-catalytic SCF conditions, Biomass and Bioenergy, vol. 33,
no. 1, pp. 113–118, 2009.
41. M. Taherkhani and S. M. Sadrameli, An improvement and optimization
study of biodiesel production from linseed via in-situ transesterification
using a co-solvent, Renew. Energy, vol. 119, pp. 787–794, 2018.
42. D. A. Kamel, H. A. Farag, N. K. Amin, A. A. Zatout, and R. M. Ali, Smart
utilization of jatropha (Jatropha curcas Linnaeus) seeds for biodiesel produc-
tion: Optimization and mechanism, Ind. Crops Prod., vol. 111, no. October
2017, pp. 407–413, 2018.
43. R. Kumar, P. Tiwari, and S. Garg, Alkali transesterification of linseed oil
for biodiesel production, Fuel, vol. 104, pp. 553–560, 2013.
44. S. S. Kashyap, P. R. Gogate, and S. M. Joshi, Ultrasound assisted intensified
production of biodiesel from sustainable source as karanja oil using interes-
terification based on heterogeneous catalyst (Γ-alumina), Chem. Eng. Process.
- Process Intensif., vol. 136, pp. 11–16, 2019.
45. B. Gurunathan and A. Ravi, Process optimization and kinetics of biodiesel
production from neem oil using copper doped zinc oxide heterogeneous
nanocatalyst, Bioresour. Technol., vol. 190, pp. 424–428, 2015.
46. Y. Essamlali, O. Amadine, A. Fihri, and M. Zahouily, Sodium modified fluo-
rapatite as a sustainable solid bi-functional catalyst for biodiesel production
from rapeseed oil, Renew. Energy, vol. 133, pp. 1295–1307, 2019.
47. R. K. Elango, K. Sathiasivan, C. Muthukumaran, V. Thangavelu, M. Rajesh,
and K. Tamilarasan, Transesterification of castor oil for biodiesel production:
Process optimization and characterization, Microchem. J., vol. 145, pp. 1162–
1168, 2019.
48. A. Bouaid, L. Bajo, M. Martinez, and J. Aracil, Optimization of biodiesel
production from jojoba oil, Process Saf. Environ. Prot., vol. 85, no. 5 B,
pp. 378–382, 2007.
49. M. I. Loizides, X. I. Loizidou, D. L. Orthodoxou, and D. Petsa, Circular
bioeconomy in action: Collection and recycling of domestic used cook-
ing oil through a social, reverse logistics system, Recycling, vol. 4, no. 2,
2019.
50. R. M. Mohamed, G. A. Kadry, H. A. Abdel-Samad, and M. E. Awad, High
operative heterogeneous catalyst in biodiesel production from waste
cooking oil, Egypt. J. Pet., no. xxxx, 2019.
51. T. Maneerung, S. Kawi, Y. Dai, and C. H. Wang, Sustainable biodiesel
production via transesterification of waste cooking oil by using CaO
308 Biodiesel Technology and Applications

catalysts prepared from chicken manure, Energy Convers. Manag., vol. 123, pp.
487–497, 2016.
52. M. Mohadesi, B. Aghel, M. Maleki, and A. Ansari, The use of KOH/
Clinoptilolite catalyst in pilot of microreactor for biodiesel production from
waste cooking oil, Fuel, vol. 263, no. September, p. 116659, 2020.
53. C. T. Alves et al., Transesterification of waste frying oils using ZnAl2O 4 as
heterogeneous catalyst, Procedia Eng., vol. 42, pp. 1928–1945, 2012.
54. M. R. Abukhadra and M. A. Sayed, K+ trapped kaolinite (Kaol/K+) as low
cost and eco-friendly basic heterogeneous catalyst in the transesterification
of commercial waste cooking oil into biodiesel, Energy Convers. Manag., vol.
177, pp. 468–476, Dec. 2018.
55. V. Narula, M. F. Khan, A. Negi, S. Kalra, A. Thakur, and S. Jain, Low tem-
perature optimization of biodiesel production from algal oil using CaO and
CaO/Al2O3 as catalyst by the application of response surface methodology,
Energy, vol. 140, pp. 879–884, 2017.
56. A. Xaaldi Kalhor, A. D. Mohammadi Nassab, E. Abedi, A. Bahrami, and A.
Movafeghi, Biodiesel production in crude oil contaminated environment
using Chlorella vulgaris, Bioresour. Technol., vol. 222, pp. 190–194, 2016.
57. K. A. Salam, S. B. Velasquez-Orta, and A. P. Harvey, Kinetics of fast alkali
reactive extraction/in situ transesterification of Chlorella vulgaris that iden-
tifies process conditions for a significant enhanced rate and water tolerance,
Fuel Process. Technol., vol. 144, pp. 212–219, 2016.
58. S. H. Teo, A. Islam, T. Yusaf, and Y. H. Taufiq-Yap, Transesterification of
Nannochloropsis oculata microalga’s oil to biodiesel using calcium methox-
ide catalyst, Energy, vol. 78, pp. 63–71, 2014.
59. G. Kalavathy and G. Baskar, Synergism of clay with zinc oxide as nanocata-
lyst for production of biodiesel from marine Ulva lactuca, Bioresour. Technol.,
vol. 281, no. January, pp. 234–238, 2019.
60. S. Wu, L. Song, M. Sommerfeld, Q. Hu, and W. Chen, Optimization of an
effective method for the conversion of crude algal lipids into biodiesel, Fuel,
vol. 197, pp. 467–473, 2017.
61. U. Rajak and T. N. Verma, Effect of emission from ethylic biodiesel of edible
and non-edible vegetable oil, animal fats, waste oil and alcohol in CI engine,
Energy Convers. Manag., vol. 166, no. X, pp. 704–718, 2018.
62. C. Öner and Ş. Altun, Biodiesel production from inedible animal tallow and
an experimental investigation of its use as alternative fuel in a direct injection
diesel engine, Appl. Energy, vol. 86, no. 10, pp. 2114–2120, 2009.
63. V. Mutreja, S. Singh, and A. Ali, Biodiesel from mutton fat using KOH
impregnated MgO as heterogeneous catalysts, Renew. Energy, vol. 36, no. 8,
pp. 2253–2258, 2011.
64. D. Papargyriou et al., Investigation of solid base catalysts for biodiesel pro-
duction from fish oil, Renew. Energy, vol. 139, pp. 661–669, 2019.
Non-Edible Feedstock for Biodiesel Production 309

65. A. S. Al Hatrooshi, V. C. Eze, and A. P. Harvey, Production of biodiesel


from waste shark liver oil for biofuel applications, Renew. Energy, vol. 145,
pp. 99–105, 2020.
66. K. Seffati, B. Honarvar, H. Esmaeili, and N. Esfandiari, Enhanced biodiesel
production from chicken fat using CaO/CuFe 2 O 4 nanocatalyst and its
combination with diesel to improve fuel properties, Fuel, vol. 235, no. August
2018, pp. 1238–1244, 2019.
67. P. R. Pandit and M. H. Fulekar, Biodiesel production from microalgal bio-
mass using CaO catalyst synthesized from natural waste material, Renew.
Energy, vol. 136, pp. 837–845, 2019.
68. B. Karmakar and G. Halder, Progress and future of biodiesel synthesis:
Advancements in oil extraction and conversion technologies, Energy Convers.
Manag., vol. 182, no. December 2018, pp. 307–339, 2019.
69. V. B. Veljković et al., Biodiesel production from corn oil: A review, Renew.
Sustain. Energy Rev., vol. 91, no. April 2017, pp. 531–548, 2018.
70. S. N. Gebremariam and J. M. Marchetti, Economics of biodiesel production:
Review, Energy Convers. Manag., vol. 168, no. May, pp. 74–84, 2018.
11
Oleochemical Resources for
Biodiesel Production
Gayathri R., Ranjitha J. and Vijayalakshmi Shankar*

CO2 Research and Green Technologies Centre, VIT University,


Vellore, Tamil Nadu, India

Abstract
Oleochemicals are materials composed of high energy molecules made up of tri-
glycerides (lipids) with more than 96% present inform of MG, DG, and FFA. The
energy produced when the oleochemicals heated is nearly 91% close to diesel fuel.
The major difference between the oil of fossil fuel and plant oils is due to variation
in the percentage of oxygen present in it. The direct usage of triglyceride in diesel
engines is possible but it causes lot of problems as result of reduced volatile, thick-
ness, and its properties in cold flow, which will be rectified by improving the prop-
erties of plant oil. Non-edible feedstocks including AFW have recently gained high
interest due to the quality of biodiesel produced from them and they are low cost,
eco-friendly, with reduced emission of NOx and properties such as high cetane
number and oxidative stability. Oleochemicals can be used to produce three major
biofuels which includes biodiesel, bio-oil, and renewable diesel. The advantages
of renewable biofuels are that they are similar to fossil fuels but eco-friendly with
reduced emission of greenhouse gases.

Keywords: Oleochemicals, plant oils, animal oils, biodiesel production,


optimization, purification, biodiesel properties

11.1 Introduction
Biodiesel (FAME/FAEE) is an alternative, inexhaustible, and green energy
resource than fossil fuels in order to meet the energy crisis. Utilization of

*Corresponding author: vijimicro21@gmail.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (311–340) © 2021 Scrivener Publishing LLC

311
312 Biodiesel Technology and Applications

fossil fuel results in the elevated emission of environmental pollutants/


contaminants including nitric oxides, sulfur oxides, and carbon monoxide
and other toxic volatile organic compounds which are the major causes for
deforestation, depletion of ozone, global warming, photochemical smog,
and Eutrophication of aquatic ecosystem. The amount of the emitted envi-
ronmental pollutants/contaminants can be highly reduced by replacing the
fossil fuel with biodiesel [1].
There are variety of raw materials/feedstocks available for the produc-
tion of biodiesel which include first-generation feedstock mainly focusing
on food crops, second-generation feedstocks are materials that are pro-
cessed form of first-generation feedstock such as derived products of plants
and animals in form of vegetable oil and lignocellulosic materials, animal
products including fat, oil, waxes obtained from tallow, and dairy products.
These second-generation residues are categorized into four groups. Primary
residues are materials obtained while cultivating the desired plants and in
the forest. Secondary residue comprises of intermediate and by-products
formed during the processing of food crops into their final products (e.g.,
food processing waste). Tertiary residue includes the final products result
from the derivative products of biomass available only after its consumption
by humans or animals in form municipal solid waste (MSW) that are fur-
ther subjected waste water/sewage treatment for proper disposal. The last
group involves the utilization of algal biomass. Third-generation feedstocks
are materials with high carbon content utilized for biodiesel production and
also include chemical and enzymatic conversion of algae [2].
Utilization of used waste frying oil (WFO) for FAME synthesis through
process of recycling is a potential and eco-friendly technique. This tech-
nique gives promising results in the yield of biodiesel ranging from 60% to
90%. The yield can be increased up to 95% by adjusting some physiochem-
ical and biological parameters during the process of production and sim-
ple pre-treatment before subjecting it for biochemical reactions [3]. This
method is economically feasible by cutting down the production cost up to
60%–70% [1]. This method is an eco-friendly technique which involves the
proper disposal of waste cooking oil (WCO) that is unpalatable because of
elevated level of FFA [4]. Recycling of the WCO is subjected for oleochem-
ical production [2].

11.2 Definition of Oleochemicals


Chemicals produced from natural oils and fats are called oleochemicals.
From the overall production of oils and fats (around 105 million), 80% is
Oleochemical Resources for Biodiesel Production 313

used for human consumption, 5% for animal feed, and 15% for chemical
production. For industrial production of chemicals every year 16 million
tons of fats and oils were utilized. There are two categories of oils-lauric
oils (e.g., coconut and palm kernel) with rich carbon chains involving of 12
and 14 linked carbon atoms, and (tallow and palm) with 16 and18 carbon
atoms. Rapeseed, soybean, and sunflower oil are also used for oleochemi-
cal production [5].
Oleochemicals are typically derived products of fatty acids (FAs) and
glycerol obtained through variety of chemical/biological reactions in
which separation of the triglyceride (TG) structure of oil and fats occurs.
Oleochemicals can also be synthesized from FA through modification of
their carboxylic group. The basic oleochemicals include FA, FAME, fatty
alcohols, fatty amines, glycerol/glycerine, and products obtained by hydro-
genation of FAME and fatty alcohols. Fatty chemical derivative such as
methyl ester (ME) and FAs plays a major role in the field of oleochemistry
and oleochemical-based industries [6].

11.3 Oleochemical Types


There are five different types of oleochemicals which includes ME, FAs,
fatty alcohol, fatty amines, and glycerol. FA is a lipid molecule made up of a
long fatty chain combined with a carboxyl group at the end. FAs are present
in two forms, the saturated (absence of double bond) and unsaturated FAs
(presence of double bond) based on the types of bond present in that FA
chain. Lipids such as fats and oil are made up of TGs (single glycerol mole-
cule bounded with three FA chains) and these TGs can be split into FA and
glycerol by hydrolyzing the oil/fat and then subjecting it for further reac-
tion under high temperature, pressure, or lipase catalyst under 37°C with
controlled pressure and stirring speed. The TGs present in oils and fats
are converted into ME by a process called methanolysis. The process was
illustrated in Figure 11.1. When oil/fat is subjected to react with methanol
and catalyst, the FAs get detached from the glycerol and bond with meth-
anol thus resulting in a compound called MEs and glycerol is recovered as
by-product. MEs are reported to be superior than FAs for production of
various FA derivatives. Fatty alcohols are long-chain high molecular weight
primary alcohols derived from oil and fats (lipids). They are produced by
methanolysis process in which the esters react with alcohol after the for-
mation of ME. Fatty amines are derivative products of fatty alcohol and
FAs. The production of primary amines is gradually declining, but there
is an increased utilization and demand for secondary and tertiary amines.
314 Biodiesel Technology and Applications

Animal
products

Edible
products

Fish industry Poultry


Cattle and sheep products
waste products

Crushing process
Wet
crushing Dry crushing

Pressing

Separation of solid and


liquid materials

Rendering process

Evaporation of
liquid

solid waste of solid waste solid waste of


Chicken fat of chicken cattle/sheep
fish

Fish oil and Protein meal Pet feed and


Protein meal tallow

Fish oil, tallow, chicken


fat usedas feedstock for
biodiesel production

Figure 11.1 Process flow of methyl ester production from animal fat.
Oleochemical Resources for Biodiesel Production 315

They are used as biocides, mineral oils, corrosion inhibitors, in the oil field
as additives (demulsifiers), used in textiles, fibers, industrial dyes and pig-
ments, soft detergents, softening agents, and also used in the following
fields, as anti-cracking agents for road construction, etc. Fatty amines con-
taining ca. 25% are used as anti-cracking agents and ca. 75% derived product
are used as conditioning agents in cosmetics, amine oxides, and amine-
EO-adducts [7]. Glycerine occurs as a by-product when oil/fats are sub-
jected for any reactions in which water (hydrolysis), methanol (meth-
anolysis), and acid/base catalyst (saponification) takes place, this due to
breaking of glycerol from FAs. Glycerine is primarily used for its viscosity
and moisture retaining capacity [1, 4–8].

11.4 Production of Biodiesel


The synthesis and utilization of FAME have created a great interest in the
financial investment and implementation of biodiesel production plants
among various countries. Based on the report of the European Biodiesel
Board, the production of Biodiesel by the European Union in 2011 was
24.7 × 109 L yr−1 and 4.16 × 109 L yr−1 by the USA in 2011 [4]. The amount
of biodiesel produced by Canada has increased twice when comparing its
annual production in 0.210 × 109 L yr−1 (2012) and 0.471 × 109 L yr−1 (2013)
[20]. In USA, Europe, and Canada, the biodiesel was produced from rape-
seed oil, soybean oil, and canola oil with low content of erucic acid [9].
There is an increased demand for production of biodiesel in Malaysia,
Canada, Europe, USA, and China. So, in order to meet the demand, there
is a need for use of rendered oils, palm oil (Elaeis guineensis jacq.) and
rendered animal fats. The animal fat was classified as feeds by the interna-
tional market, and their price become low. Hence, the cost of biodiesel will
be extremely lowered by the AFWs to increase the production of biodiesel
to compete with Petro-diesel.
The various feedstocks to produce ME include animal fats and vegeta-
ble oils both edible and non-edible oils [10–14]. The two major limiting
factors involved in the use of various substrates for the production of bio-
diesel are their obtainability and price. The industries store their feedstock
which serves as a limiting factor for sustainability of biodiesel regardless of
their current growth. There are limitations in the sustainability of biodiesel
production regardless of its current growth is due to inefficiency of the
industries to safely store their cheapest feedstocks. The chief source for the
production of FAME in Canada is Canola oil. The feedstocks used by USA
and Europe is edible vegetable oil but this leads to an increased demand for
316 Biodiesel Technology and Applications

vegetable oil supply that will result in food security, ethical issues, and pro-
voke worldwide arguments regarding the level of effect on the diversity of
food crops at global level. This could be sought by replacing the feedstocks
that produce good quality biodiesel with low capital cost and not a threat
to food safety [15].
The feedstocks that can be used to avoid such issues include chicken fat,
tallow, lard, and solid waste materials produced in leather industry, etc.
In some regions, the animal fat waste (AFWs) are abundant that result in
lowering the cost of feedstocks. The advantage of using waste animal fat for
producing biodiesel is that it emits NOX which will be below or equal to
vegetable oil biodiesel with a reduced amount of environmental pollutant
and thus preventing environmental damage. Quality of the produced diesel
fuel is determined by the cetane number. The ratio of saturated FAs is high
in AFWs which is the major reason for high cetane number (460) when
compared with most of the vegetable oils. The advantage of high cetane
number is that it reduces the temperature during the early combustion
process resulting in low NOx emission. The presence of saturated FA in
the AFWs is responsible for better oxidative stability of biodiesel [16–19].
There is a huge challenge in producing biodiesel by alkali-catalyzed trans-
esterification because of the presence of more free FAs in AFWs. Efficiency
of the biodiesel is based on the techniques involved during the construction/
development of the biodiesel plant. Chemical and biological techniques are
the two important techniques currently used for the synthesis of FAME
from plant oil and animal fats. Synthesis of biodiesel using chemical tech-
niques includes acidic or basic catalytic conversion reactions. Enzymatic
conversion of biodiesel takes place in biological technique. Lipase enzyme
is mainly used in biological techniques. Other than these two techniques,
the use of other techniques such as non-catalytic and superficial tempera-
ture conversion methods is still under the process of investigation. Even
though chemical-based biodiesel production is widely used in industrial
scale, they have certain limiting factors such as high energy, equipment, and
alcohol requirements, and the downstream and purification process was
complex. Through enzyme catalyzed transesterification process, there is an
advantage of overcoming a specific limitation in producing diesel with the
feedstock having low FFA content and moisture that are required for chem-
ical techniques. Development of high molar ratio of animal fat and soap
required for production of biodiesel during which chemical-based catalysis
will be comparatively low in the biological method. Enzyme immobilized
techniques made the lipase enzyme reusable and developed high alcohol
tolerant lipases; thus, this technique become cost efficient by overcoming
the inactivation of enzyme lipase and its higher cost [15, 20].
Oleochemical Resources for Biodiesel Production 317

There are many new techniques emerging for biodiesel production


which involves many treatments for maximum yield with the reduced
reactor operating time. The microwave-assisted (MA) transesterification
and ultrasound-assisted (UA) transesterification are the two major emerg-
ing techniques for biodiesel production. The use of cheap reagents, short
reaction time, and simple process reactors created a great interest in the
UA techniques for biodiesel production. There is also a positive approach
toward MA technique due to its great energy efficiency, improved yield,
short reaction period, and highly purified products [15].

11.5 Types of Feedstocks


11.5.1 Non-Edible Feedstocks
There is a global demand for edible oil and animal fat due to increased pop-
ulation and resulted in elevating their price on the world market that many
developing countries had limited the utilization of plant oil and animal
fat as fuel. Several studies stated that grease and fat obtained from animal
source waste are feasible, good, and eco-friendly feedstocks for FAME pro-
duction. These materials are easily accessible in many developing countries
through placing order in units of teragrams. There is a worldwide availabil-
ity of non-edible oil plants. But there will be a problem leading to strong
debate due to the inaccessibility of land for food crops vs. non-edible oil
crops [15].

11.5.2 Non-Edible Vegetable Oil


Oil from non-edible sources for synthesizing biodiesel from Oleagnious,
Leguminosae, Brassicaceae, and Euphorbiaceae have been under several
investigations in recent years. These plants produce oil-rich seeds but their
high FFA level made them toxic and inedible for both animals and human
consumption. The use of non-edible oil can directly gain opposition with
respect to safe and availability of food. Pongamia, Jatropha, Brassica,
Madhuca, Gossypium, Cerbera, Nicotiana, Thevettia, Ricinus, and Hevea
sps are the common non-edible crops used for the production of biodiesel.
The non-edible plants used for the research of biodiesel production and
their cultivation varies according to its place of growth. For example, sev-
eral types of researches mainly concentrating on oleaginous microorgan-
ism for lipid-based biodiesel feedstocks are currently going in the USA,
Canada, and EU [15].
318 Biodiesel Technology and Applications

11.5.3 Tall Oil


Tall oil is the secondary product produced by wood pulp industry. It con-
tains 42% resins and 45% of FAs. Distillation method is used to separate
the resin and FAs which are present together and only in the form of free
acids [21].

11.5.4 Waste Cooking Oils


There is no direct opposition for the usage of WFO as a source for biodiesel
concerned with land usage and availability of food. Yellow grease and brown
grease are the two groups of waste cooking or frying oil (WFO). WFO is
highly concentrated with grease after cooking bacon, meat appetizers, and
hamburgers, and this is due to the method used for the extraction of cook-
ing oils. Since the cooking oil is extracted by the melting of fat obtained
from AFW and heating of vegetable oil that is used for cooking various
foods including meat, fish, and vegetable oils. This is the major reason for
the formation of grease after cooking. WFO collected from the restaurant or
industrial operations with the presence of less than 15% of FFA in the grease
produced from WFO, fats, and oil act as feasible potent and inexpensive raw
material for biodiesel production. Restaurants and industries involve the use
of grease trap to collect brown grease in order to separate oil and grease from
the wastewater for municipal sewage facilities. Grease traps allow floating
of light grease and oil at the top of the trap when flushed in the drain and
thus facilitating the free flow of wastewater into the water treatment unit.
Difficulties in the potential conversion of biodiesel from brown grease with
greater than 15% FFA are due to the presence of H2O content in it. The tech-
niques used for conversion of biodiesel from WFO include various types
of transesterification process such as acid-catalyzed transesterification, alka-
line-catalyzed transesterification, two-step transesterification, enzymatic
catalysis, and supercritical temperature processing methods. Heating virgin
oils for long duration leads to higher FFA concentration of FFA in WFO.
Biodiesel yield from WCO is lowered and involves a complex downstream
process during the separation of ME, C3H8O3, and washing H2O, and this is
due to the saponification reaction taking place when the alkaline catalyst acts
on FFA during the alkaline transesterification process [15].

11.5.5 Animal Fats


The by-products of processed and rendered animal meats from facilities are
the primary source used for deriving fats from animals. From cattle processing
Oleochemical Resources for Biodiesel Production 319

facilities and rendering process, tallow can be obtained; grease and lard can
be obtained from processing the pork, and poultry fat can be obtained from
chicken and other birds processing; oil fish industry and waste from leather
industry are the major animal fats used for synthesis of FAME. Chicken, tal-
low, and lard fats are some of the animal fat-based feedstocks used for large
industrial scaled biodiesel production. Compared with edible plant oils, AFW
used as feedstock for FAME production have economical environmental and
food security advantages. Production of FAME from AFWs involves complex
techniques when saturated FAs (FFA) is present at a high level, and the result-
ing biodiesel will have low chemical and physical quality. There are certain
advantages in low unsaturated FAs of AFWs such as increased cetane num-
ber, increased oxidation stability, and increased calorific value [15].

11.5.6 Chicken Fat


During the feather meal preparation, the chicken fat can be produced.
Head, intestine, undeveloped eggs, and feet other than feathers are clean
carcass available in poultry by-products found in different portions of wet
rendered or dry grounded forms. Based on the type of feather used, the fat
thus obtained varies from 2% to 12%. Biodiesel has been produced by Guru
et al. (2010) [66], with fat from chicken using catalytic process which con-
tains two steps involving catalyst such as CH3OH, H2SO4, and NaOH. The
effects on performance in engine and exhaust emission when injected in a
diesel engine with produced biodiesel from synthetic Mg additive chicken
fat have been reported by the author. The decrease in the viscosity (5.184
to 4.812) and flash point (129°C to 122°C) due to rise in concentration of
Mg have been reported by the author. There was a decrease of 7°C pour
point in the chicken fat ME with the addition of Mg. Yield of ME was about
87.4% when the conditions for animal fat was optimized by pre-treatment.
When converting chicken fat into biodiesel the point at which best TG
conversion and glycerol decomposition were obtained at 400°C, 300 bar
pressure, 9:1 molar ratio, with 6-min time [15].

11.5.7 Lard
Rendered fat of pig is termed as lard. Dry or wet process is used for render-
ing the pig fat. Wet rendering involves high temperature boiling of pig fat in
water/steam and the insoluble lard is formed as an upper layer in the mix-
ture which is skimmed off or centrifuged. Pig fat is subjected to higher heat
treatment in the absence of water in oven or pan. These two processes result
in the yield of two different products. Lard with a neutral flavor, light color,
320 Biodiesel Technology and Applications

and high smoke point can be obtained from the process of wet rendering
process and dry rendering results in lard production with brown color and
low smoke point. According to Dias et al. [67] on the generation biodiesel
by acid transesterification, the lard was pretreated with KOH to enable it
to be a potential feedstock for FAME production with suitable quality but
the yield is low with only 65% by weight. The yield of biodiesel from waste
lard and WFO ranges from 81.7 to 88.0 (wt.%) has been reported from the
investigation done on the synthesis of FAME from mixtures of WFO; thus,
the yield is low when only WFO and lard have been used as raw materi-
als. Investigations have been done on the conversion of biodiesel from lard
using the Candida sp. (fungus) as a biocatalyst in transesterification process
shows the optimal condition for the reaction in order to process a gram
of lard involves immobilized lipase (0.2 g), 8-ml n-hexane, a ratio of 20%
of water to the weight of fat, 40°C and adding CH3OH produced 87.4% of
FAME. This investigation gives a detailed report about the effect of biocata-
lyst used in transesterification of lard involving the parameters such as tem-
perature, water content, the quantity of enzyme used, solvent, in three-step
methanolysis process for biodiesel production. The ME production was
directly affected by the concentration of catalyst used and agitation speed,
and the optimal agitation speed is found to be 600 rpm found in a regression
model used to the predict concentration of ME with sufficient experimental
parameters. According to the experiment of Shin et al. [68], on production
of biodiesel with lard without pre-treatment and using the technique with
supercritical without adding the catalyst under optimal reaction conditions
produced biodiesel which is similar to biodiesel synthesized from refined
lard, although it has free FA and water in it [15].

11.5.8 Tallow
Animal by-products from slaughter house can be rendered to produce fat and
protein meal. More than 50% of FAs in the tallow are in saturated form. The
high melting point and viscosity is because of the presence of palmitic and
stearic acid in the tallow leads to makes it solid at room temperature. Edible
or non-edible tallow are derivatives of beef or mutton which are mostly used
to produce FAME. For alkaline transesterification reaction process, edible
tallow acts as a potential substrate having low FFA content. Biodiesel pro-
duction from beef tallow that was studied and reported conclusively stated
that tallow from beef obtained from cattle house remarkably acts as an inex-
pensive feedstock for FAME production with an immense energy, along with
economic and environmental advantages and this report also included the
economic feasibility, energy efficiency, and resource availability of beef tallow
Oleochemical Resources for Biodiesel Production 321

and its conversion into biodiesel. The yield of FAME from edible tallow and
the effect of catalyst, FFA, water content, and mixing intensities have been
investigated by Ma and Hanna and reported that reaction takes place only
when NaOH and CH3OH were added and mixed with melted beef tallow
with limited stirring speed and also stated that higher stirring speed for a lon-
ger period is required to mix the two phases when a blend of sodium hydrox-
ide and methanol solution is added before stirring the melted beef tallow.
Biodiesel production from non-edible tallow requires highly expensive pro-
cessing methods, which leads to higher FFA. Viability of FAME conversion
from inedible tallow has also been reported. Based on the report of Bhatti
et al. (2008) [69] and their studies on biodiesel production, its reaction param-
eters for the conversion of 5 grams of tallow waste are 60°C, 1:3 molar ratio,
and addition of sulfuric acid (2.5g) [22]. Oner and Altun (2009) have done an
experiment in which they directly injected the biodiesel produced from non-
edible tallow in a diesel engine and studied its emission and performance.
This experimental report clearly shows that there was a significant decrease
in the emission of carbon monoxide (15%), NOX (38.5%), sulfur dioxide
SO2 (72.7%), and smoke opacity 56.8% for ME (B100) from tallow [23].
Teixeira et al. (2009) did a comparative study on FAME production from
beef tallow by using ultrasonic and conventional method and reported that
only the time for reaction is reduced but the reaction rate and quality of
biodiesel remained the same for both. Kumar et al. (2013) [24] reported the
effect of enzyme ns88001with CH3OH and n-hexane used as solvent. This
report shows that n-hexane reduced the toxicity of alcohol by stabilizing the
enzyme and the NS88001-enzyme is reduced a little bit after 10 cycles [25].
Immobilized PS-30 lipase enzyme used for transesterification of Primary
alcohol (methanol) resulted in the effective conversion of biodiesel ranging
from 82% to 94% when compared with free lipase enzyme with low conver-
sion efficiency ranging from 47% to 89%, and ME showed poorest yield [15].

11.5.9 Leather Industry Solid Waste Fat


Leather industry generates huge amount solid waste fat during various
processes. This fat remains solid at room temperature with high H2O con-
tent in it. The water content is eliminated by treating the fat at 110°C for
1 hour and insoluble materials are removed by filtering it. A study [26]
experimentally proved the utilization of these fuels in engine (without any
modifications) and synthesized from the fat produced in tannery flesh-
ing waste. The reaction parameters were studied by Isler et al. (2010) [70]
which includes molar ratio (1:6), weight percentage (0.75%), and cata-
lyst and temperature for production of ME from raw fleshing oil through
322 Biodiesel Technology and Applications

base-catalyzed transesterification reaction and found that the efficient bio-


diesel yield can be achieved 1:6 molar ratio, temperature (50°C), 0.75 per-
centage weight of catalyst, and time of reaction is about 15 min. Evaluation
of FAME production from leather industry waste as feedstock in alka-
line catalyzed transesterification reaction was reported by Alptekin et al.
(2012) [27]. Property of produced ME met the biodiesel standard, and to
enhance the cold flow property or cold flow, enhancers can be added to
it. Supercritical CH3OH approach in transesterification of leather tanning
waste for production of FAME was first applied by Ong et al. (2013) [65].
A model was developed by the author for the kinetic reactions of FAME
production from leather industry waste [15].

11.5.10 Fish Oil


FAME can be produced from oil-rich waste obtained in a significant
amount during the processing of fish from the fish processing industry.
Large fraction of the effluent contains oil (6%–11%) will be utilized as a
feasible feedstock for FAME production with good yield can be obtained
with proper separation techniques. Recovery of fish oil from the effluent
involves the following methods such as grinding/­homogenizing of solid
fish waste followed by heat treatment for 15–20 min at 95°C–100°C, screw
pressing of the solid waste to remove liquids, centrifugation of the liquid
to separate oil and remove wastewater, and final polishing of oil by water
washing. Colak et al. (2005) [26] investigated production of FAME with
fish oil and reported that catalyst concentration gives greater yield and
quality of biodiesel. This report shows that optimum condition for bio-
diesel production is 1 wt.% catalyst and 60 vol% of CH3OH solution. Waste
characterization and quality assessment of fish processing plant effluent for
FAME production have been investigated by Tanwar et al. (2013) [28] and
Jayasinghe and Hawboldt (2013) [29]. The major factors which determine
the suitability of the transesterification technique for FAME production
are the type and quality of the feedstock used.

11.6 Uses of Oleochemicals


11.6.1 Polymer Applications
Linoleum is produced from the linseed oil. The demand for linoleum has
been increased from 10,000 tons to 50,000 tons from the year 1975–1998.
Epoxidized soybean oil is used for plastic and coating additives, and the
Oleochemical Resources for Biodiesel Production 323

production is around 100,000 tons/year. Industrial production of dicar-


boxylic acid involves ozonolysis of oleic acid to produce diacids [4].
Hydrocarbons and chlorinated compounds are used in large volume as
solvents for coating and other various polymer applications. Ester solvents
are the largest groups representing the Green solvents naturally obtained
from the FAs and alcohols [5].

11.6.2 Application of Plant Oil as a Substitute for Petro-Diesel


Ground nut was used as a fuel for demonstration in 1900 [30]. In 1930s and
1940s, experimental work has been done with plant oil in diesel engine.
Due to depleting non-renewable energy and fuel energy crisis during the
year 1970–1980 lead to the development of alternatives to petroleum-based
fuels. This gives rise to many investigations in finding various alternatives
for Petro-diesel and resulted in the production of biodiesel. Currently, bio-
diesel have become commercial and have a global level demand. Several
plant oils were investigated for the production of FAME. FAME produc-
tion will be determined upon availability of substrates such as palm oil,
coconut oil, soybean oil, rapeseed oil, sunflower oil, safflower oil, and some
vegetable oil used for cooking. Plant oil directly in diesel engine is still
problematic because of its high viscosity; further development is needed to
overcome this problem. Following techniques can be applied to maximize
the quality of the fuel from plant oil which includes dilution of oil from
plant (25 parts) with Petro-diesel (75 parts). The physiochemical prop-
erties of biodiesel (FA esters) are similar to Petro-diesel. The production
process is simple and the burning efficiency of methyl or ethyl ester of FA
in diesel engine does not require any modifications. When compared to
microemulsion and dilution techniques, transesterification is a best choice
for biodiesel production [5].

11.6.3 Used as Surfactants


Oleochemicals are used as surfactants used to bring two different types of
compounds together. Oleochemicals with FA chains or long-chain alcohol
are used for this purpose. Oleochemical used as surfactant involve com-
pounds such as bio-based hydrophiles including amino acid [31], poly-
saccharide, and FA chains with carboxylate groups. Bio-based surfactant
should have its hydrophobic or hydrophilic group to be derived. Group
of bio-based surfactants includes oleochemical-based fatty esters that are
used as hydrophobe group to combined with ethylene oxide as hydrophile
group. Carbohydrates were used as hydrophilic group of surfactants for
324 Biodiesel Technology and Applications

more than half of a century. Carbohydrates such as sorbitol (lignin, lecithin


to sugars) are used surfactants. Some surfactant has both the hydrophilic
and hydrophobic groups to be bio-based such as sugar and FA, which
is beneficial in case of SpanTM 60. Glycerol produced in excess during
biodiesel production can be used as a cheap material for the bio-based
surfactant production as a precursor. Polymers of glycerol such as such
as polyglycerides act as good hydrophobes. Surfactant with polyglycerol
properties will be produced either from compound contains epoxide such
as glycidol or by polymerization of glycerol. The utilization of surfactants
involves various fields. The most important field where more than 50% of
surfactants used includes washing and cleansing sector, textile treatment,
and cosmetics. Other fields such as mining, protection of crop, paints,
coatings, inks, and adhesives also involve the application of surfactants [5].

11.6.4 Oleochemicals Used in Pesticide


Plant-based oils and fats act as an important raw material for chemicals
industry [32]. Plant-based oleochemicals are more advantageous than min-
eral oil that they are renewable, biodegradable, and widely available (by-
products from industries). Pesticide formulation is mainly composed of
two major components such as inert and active ingredients [33].

11.6.5 Oleochemicals Used in Spray Adjuvants and Solvents


Oleochemicals have been used as spray adjuvants due to their high adher-
ence capacity to the leaves of plants or insects. The polyunsaturated nature
of this plant oil made them persistent on leaves and insects even after rain-
fall. Palm olein in liquid fraction from palm oil also used as spray adjuvants.
MEs and hydrocarbons act as potential vehicles for transporting pesticides
on the surface of plant and insects. This is due to high viscosity and surface
tension of methyl eater and hydrocarbons. They are biodegradable with
less toxicity and less viscous with better solvency. The solvency property to
pesticide is better in ME produced from plant oils [5].

11.7 Methyl Ester or Biodiesel Production


Biodiesel (ME) from plant oil and fat resulting with the glycerol release
[15].
Conversion of plant oil into biodiesel includes TGs’ reaction such
as methanol with catalyst resulting in FAME and C3H8O3. Initially, in
Oleochemical Resources for Biodiesel Production 325

transesterification process, the TGs are converted into DG followed by


a series of reactions involving larger glycerides converted into smaller
glycerides and finally into glycerol (Figure 11.2), with one ME molecule
at the end of each steps of the subsequent reactions [10]. In this revers-
ible reaction, the equilibrium is shifted toward the product side by using
excess alcohol [15]. Solubility of alcohol will be improved by the catalyst,
thus result in increasing the reaction rate and yield. Utilization of catalyst
depends upon the substrate and content of free FA. The FFA and H2O con-
tent should be below 0.5 and 0.05 wt.%, to prevent saponification during
the reaction [34]. FFA can be converted into FAEE through esterification
reaction as a pre-treatment process for effective conversion of oil into
biodiesel [35]. The TG and DG can be hydrolyzed into FFA using water.
Temperature is an important parameter influencing the rate of reaction,
which is fast when the temperature is high [36]. Many techniques were
used in the transesterification reaction which includes the usage of cata-
lysts of heterogeneous type (e.g., magnesium oxide, calcium oxide, and sul-
fated zirconia) [35]. BIOX process is also a noncatalytic transesterification
method used for ME production from WCO, animal fat, and vegetable oil
[7, 37–39]. For transesterification process, the WCO has to be pre-treated

Step 1

Esterification
feed stock oil extraction
(pre-treatment)

Step 2

Transesterif ication • addition of methanol/ethanol

Ester recovery • release of by-product (glycerol)

Biodiesel
FAME/FAEE

Figure 11.2 Production of biodiesel by transesterification reaction. Step1. Pre-treatment


of oil to treat FFA and water through esterification reaction. Step 2. Transesterification of
pre-treated oil to produce fatty acid methyl ester/ fatty acid ethyl ester (biodiesel).
326 Biodiesel Technology and Applications

due to presence of high FFA (0.5%–15%) content for proper conversion


and yield [36].

11.7.1 Palm Oil


Acidic crude palm oil (ACPO) can be utilized as potential substrate for
the production of ME production. Cost for production can be lowered
by using the by-products of APCO [40]. The catalyst commonly used for
the pre-treatment of acidic oils in esterification reaction includes sulfu-
ric acid—conventional catalyst [41] (lipase-biocatalyst, Ferric sulfate-
heterogeneous catalyst (widely used). Ethane sulfonic acid is a catalyst used
for treating high FFA content in industrial ACPO by [41] for esterification
process. Methane sulfonic acid (MSA) is a catalyst used for pre-treatment
of FFA in APCO by (Adeeb Hayyan et al., 2012) [40].
Materials used in this reaction are APCO, CH3OH 99.8 %, potassium
hydroxide pellets 85%, and MSA. Before starting the esterification reac-
tion, the ACPO has to be melted at a 65°C. Then, the FFA content was
measured and reported. For FAME production, 200g of ACPO was taken
for each experiment in which both transesterification and esterification
reactions were involved. Conversion of FFA into FAME through ester-
ification reaction is a preliminary stage. There is a high yield of FAME,
and enhanced esterification reaction due to addition of MSA catalyst was
observed by Adeeb Hayyan et al. (2012) [40]. The liquid phase of APCO
and MSA were found to be responsible for enhanced esterification reac-
tion. It also converted triacylglycerols (TAGs) into FAMEs. In the second
reaction, the TAG is converted into FAME by using KOH as a catalyst [40].

11.7.2 Sunflower Oil


Production of ME from sunflower oil by alkali-catalyzed transesterifica-
tion reaction was done by Umer Rashida et al. (2008) [42]. About 500 g
of sunflower oil was preheated for setting the temperature that ranges from
(30°C, 45°C, or 60°C) [43] with known catalysts and mixed well. Each
experiment was subjected for 120 min at 600 rpm. Different molar ratio
was used including 3:1, 6:1, 9:1, 12:1, 15:1, and 18:1 [44]. The sunflower
oil methyl oil (SOME) was purified by distillation around 80°C in which
the left out methanol is removed by subsequent washing with distilled
water and Na2SO4 along with filtration. The unreacted catalyst is treated by
neutralization with concentrated sulfuric acid and decomposition of soap
formed during the reaction. Methanol is recovered by vacuum treatment
under 80°C to purify the C3H8O3.
Oleochemical Resources for Biodiesel Production 327

Results reported from this experiment were mentioned in the following:


In this experiment, temperature variation has no significant effect on bio-
diesel yield. After the reaction was completed, it is more than 80% without
the influence of temperature change. The yield was reported to be 97.1% at
60°C, 92.8% at 45°C, and 90.9% at 30°C, and this indicates the efficiency of
base catalyst to be high nearly to the boiling point of alcohol. Different con-
centrations of NaOH (0%, 0.25%, 0.50%, 0.75%, 1.00%, 1.25%, and 1.50%
w/w) [45] were used among four different catalysts in same experimental
conditions mentioned above, and the yield was found to be 97.1% at a con-
centration of 1.00% NaOH. Six different molar ratios were used in each of
the experiments ranging from 3:1, 6:1, 9:1, 12:1, 15:1, and 18:1, and 97.1%
was found at 6:1. This experiment shows that the optimal reaction condition
for the production of ME from sunflower oil through transesterification was
found to be at 60°C,6:1, concentration of catalyst 1.00% with 97.1% yield.

11.7.3 ME From AFW


Production of ME from various animal fat sources was explained in Figure
11.3. Guru et al. (2010) [66] reported the production of FAME from
chicken fat. They reported that yield of biodiesel to 87.4% with an opti-
mal reaction condition of 20% H2SO4 at 40:1 methanol molar ratio with
13.45% FFA in chicken fat was subjected to reaction for 8 min at 60°C. The
yield of FAME was 88% when the lard was subjected under the following
reaction conditions: 1-g lard, with 20% H2SO4 based on weight of the fat,
0.2g of immobilized lipase catalyst, and 8 ml of n-hexane. The ME yield
was reported to be 93.21% when 5 g of mutton tallow was treated with 2.5g
of H2SO4 at CH3OH molar ratio of 1:3 for 24 h under 60°C [15].

11.8 Parameters Affecting the Yield of Biodiesel


11.8.1 Reaction Conditions
The production of FAME can be affected by the reaction conditions such as oil
to alcohol ratio, temperature, catalyst concentration, and type of reactor used.

11.8.2 Catalyst
11.8.2.1 Alkali Catalyst
Saponification reaction reduces the FA yield and forms emulsions which
cause difficulties in glycerol and biodiesel separation. Singh et al. reported
328 Biodiesel Technology and Applications

Seed storage

Removal of cortex
(Decortication)

Size reduction by flaking

Breaking of seeds

Pre-expelling of
crude oil

Solvent extraction

Oil cake or oil meal Crude oil

Oil storage

Refining of oil

Figure 11.3 Process of extraction of vegetable oil (e.g., sunflower seed oil extraction
process).
Oleochemical Resources for Biodiesel Production 329

that better yield was obtained when hydroxide catalysts is replaced with
methoxide catalysts and sodium-based catalysts replaced by potassium-­
based catalysts [46].

11.8.2.2 Acid Catalyst


There is a slow reaction rate when acid catalyst is used and it requires more
TGs and high ratio of alcohol; hence, they are usually a best choice for
feedstock pretreatment with maximum FFA where conversion of soap into
ester is required [15].

11.8.2.3 Biocatalyst
Lipase immobilized transesterification is a very slow process and requires
more reaction time and temperature ranging from 4 to 40 h and 35°C
to 45°C, respectively. But, this process gives promising results only for
feedstocks containing high FFA contents. The FA ME produced by using
various strains of fungus Aspergillus sp. has been reported to have high
biodiesel properties (lubrication and fuel properties) [47].

11.8.2.4 Heterogeneous Catalyst


CaO and Ca(OCH3)2 are usually used heterogeneous catalyst for biodiesel
production. Slow reaction is seen due to low mass transfer ability and
requires a co-solvent with a boiling point similar to methanol for enhanced
mass transfer among the reactants and thus resulting in the fast reaction
without leaving the catalyst as residue with a reaction period of 5 to 10 min
at 30°C. Alkali doped metal oxides were used as an alternative for homog-
enous catalyst [48].

11.8.2.5 ME Conversion by Supercritical Method


Supercritical methanol transesterification is a modified method used to
avoid initiation time lag during biodiesel production. The solvents are
kept in a temperature and pressure beyond its critical point there will
be only liquid phase which facilitates simple, fast, and eco-friendly pro-
duction of biodiesel. The recovery of glycerol and separation of biodiesel
becomes much easier due to the absence of catalyst. This method involves
alcohol to oil ratio of 42:1 under 80 atm pressure at 350°C to 500°C in 2
to 4 min [15].
330 Biodiesel Technology and Applications

11.8.3 Properties of Feedstock


Quality of produced biodiesel is based on feedstocks used for its produc-
tion. The factors that affect the biodiesel production are FA contents and its
composition present in the selected feedstocks. Other factors such as elim-
ination of unsaponifiables and water and titer property oil also influence
the properties of biodiesel produced [15].

11.8.3.1 Composition of FA
The FAs present in the feedstock in different compositions highly influence
the physical and chemical properties. FAs and its composition of various
feedstocks were presented in Table 11.1 [15].

11.8.3.2 FFA
These are FAs produced as a consequence of breakage in the long car-
bon chain in a FA that result in the detachment of TG molecule due to
heating/overheating of the oil/fat. Over-heating of oils and fats results
in the production of high FFA contents and they present in the ratio
of FA wt.% in oil. This FFA leads to saponification reaction during the
transesterification process and slow down the biodiesel production
[15, 49].

11.8.3.3 Heat
The calorific value of feedstock greatly influences the energy content of the
biodiesel. Fuels with high saturation give high energy (on weight basis),
and high unsaturation provides lower energy content. The fuel has high
energy when it is denser. Heat is an important factor that affects biodiesel
quality [50].

11.8.3.4 Presence of Unwanted Materials


The feedstock contains undesirable materials that hinder the biodiesel pro-
duction, which has to be eliminated before or in the purification process.
Water content and impurities are filterable solids (including solid wastes,
bone fragments, and food particles) and unsaponifiable are non-TGs
which are incapable of producing mono alkyl fatty esters when subjected
for esterification or transesterification process [15].
Table 11.1 Fatty acid contents of vegetable oils and animal fats.
Soybean oil Palm oil Sunflower oil WCO Tallow Lard Poultry oil Fish oil
Fatty acids (wt.%) (wt.%) (wt.%) (wt.%) (wt.%) (wt.%) (wt.%) (wt.%)
Palmitic acid 10.2 42.8 5.8 8.5 23.3 23.6 23.9 14.3
Stearic acid 3.7 4.5 2.9 3.1 19.4 14.2 6.0 3.0
Oleic acid 22.8 40.5 35.0 21.2 42.4 44.2 44.0 15.0
Linoleic acid 53.72 10.1 74.5 55.2 2.9 10.7 19.9 1.4
Lauric acid 0.1 0.1 – – 0.1 0.1 – 0.2
Myristic acid 0.1 1.0 0.9 – 2.8 1.4 – 6.1
Linolenic acid 8.6 0.2 1.5 5.9 0.9 0.4 1.9 0.7
Oleochemical Resources for Biodiesel Production
331
332 Biodiesel Technology and Applications

11.8.3.5 Titer
Titer is the temperature required for melting of oil from solid into liquid
phase. Titer is an important factor which influences the biodiesel produc-
tion. Oil with high titer leads to high production cost [15].

11.8.4 Characteristic of Feedstock


The characteristics of feedstock are directly reflected in the biodiesel
obtained from it. Since the transesterification process does not have any
effect on the composition of FAs that already exist in the feedstock, this
greatly influences the quality and physiochemical properties of the bio-
diesel. Factors which affect the quality of biodiesel are oil or fat with high
FFA content which leads to high production cost, presence of MIU involv-
ing additional processing methods including filtration, heat treatment and
centrifuge, and some critical parameters like cetane number, cold flow
properties, saturation level of FAs, cloud point, flash point, and oxidative
stability [15].

11.9 Optimization of Reactions Conditions for High


Yield and Quality of Biodiesel
11.9.1 Pre-Treatment of Feedstock
11.9.1.1 Elimination of Water
During the transesterification reaction, the water present in the feedstock
at high temperature forms FFA by hydrolyzing the TG into DG, and as
a result, saponification reaction takes place even a very low amount FFA
(less than 1%) is present. To avoid the saponification reaction, the water
has to be removed from the feedstock in order to obtain good biodiesel
yield from transesterification reaction. Heat treatment and centrifuge are
some of the methods used to remove water content from the feedstock
[51].

11.9.1.2 Elimination of Insoluble Impurities


The feedstock must be filtered before subjecting it to the conversion reac-
tion in order to remove the presence of solid impurities in the form of
sand, dirt, fragments, etc., in vegetable oil and small fragments of gums,
bones, etc., and in animal fats and WCO [51].
Oleochemical Resources for Biodiesel Production 333

11.9.1.3 Elimination of Unsaponifiables


Unsaponifiables are organic compounds such as hydrocarbons, alcohol
with high molecular weight, sterols, waxes, and pigments from which soaps
cannot be formed during transesterification process. These compounds are
removed during refining process of crude oil or with water washing tech-
nique during glycerol phase if the crude oil is directly used for biodiesel
production [15, 51].

11.9.2 Characterization and Selection of Feedstocks


The followings conditions should be taken into account for selection of the
feedstock to obtain better yield and quality in biodiesel.
FFA content of the feedstock should be low, the FA composition should
be considered since it directly influences the properties of the biodiesel,
either the feedstock should be free from MIU or pre-treated before subject-
ing it transesterification reaction, and the saturation level of FAs should be
considered for better quality of biodiesel [15, 51].

11.9.3 Selection of Reaction Conditions


Biocatalyst is better, and as a pre-treatment step for producing FAAE from
FFA, acid catalyst transesterification method is suitable. Heterogeneous cata-
lyst such as calcium methoxide can be used when solvent with boiling point
equal to methanol is used. To avoid initiation time lag, supercritical methanol
transesterification can be adopted. In order to get a better yield and higher
quality biodiesel, reaction parameters have to be optimized by proper selec-
tion of feedstock, and reaction conditions play an important role [15, 51].

11.10 Oil Recovery


11.10.1 Alkaline Flooding Method
Alkaline flooding is a method used to improve oil recovery from crude
oil. NaOH, Na2CO3, and sodium orthosilicate are some of the alkali com-
pounds used for oil recovery [52]. Based on the results reported from the
experiment done by Abhijit Samanta et al. (2011) [53], there is conclusive
proof that alkali flooding enhances the oil recovery by formation of in situ
surfactant. When the alkali slug concentration was increased, the recovery
of oil was found to be increased about 15% more than the conventional
water flooding in which the recovery of oil is ~50%.
334 Biodiesel Technology and Applications

11.10.2 Additives
Hydrocarbon solvents are commonly used for extraction of TG oil
directly derived from seeds of feedstocks, e.g., sunflower seeds. Hexane
is widely as solvent in plant oil industry [54]. Fishwick and Wright
(1997) [55] have experimentally proven that chloroform-methyl alco-
hol can be used as potential solvent for extraction of plant lipids such as
sterol lipids, phospholipids (both bounded and free), and glycolipids.
Supercritical carbon dioxide is also used as a solvent for oil recovery
[56].

11.11 Quality Improvement of Biodiesel


The properties of FAME/FAEE can be enhanced by addition of different
types of additives based on the requirements.

11.11.1 Additives for Improving Combustion Ability


Barium, platinum, iron, cerium, manganese, and calcium can be used as
additives with metal based to enhance the ability to combust and less emis-
sion [57, 58].

11.11.2 Additives for Enhancing the Octane Number


Oxygenated additives are fuel containing oxygen. It is used for enhance-
ment of octane rating and improving the combustion quality. Commonly
used oxygenate additives include alcohols, ether, and ester. Alcohol
includes C2H5OH, CH3OH, C4H10O, and C₃H₈O. Ethers used are C5H12O,
C6H14O, (C2H5)2O, C2H6O, and C6H14O, etc. [59].

11.11.3 Additives for Improving the Stability


Antioxidant additives are used improve the stability and deterioration of
biodiesel during storage. Alkyl phenols, hindered phenols, and aromatic
diamines are commonly used as antioxidant additives [60–62].

11.11.4 Additives to Enhance Cold Flow Property


Ethylene vinyl acetates are used as cold flow improver additives [63].
Oleochemical Resources for Biodiesel Production 335

11.11.5 Additives to Enhance Lubricity


These surface-active compounds allow the protective film formation for
assisting the solubility of fuel and to improve the lubrication property of
biodiesel [64]. Ester, FAs, and amides are used as lubricity improver addi-
tives to reverse the lubricity lost during refining of oil [65].

11.11.6 Additives to Enhance Cetane Number


Alkyl nitrates especially 2-ethyl hexyl nitrate (2-EHN) is an important cetane
number improver additive used to enhance the cetane number. Tertiary
butyl peroxide can be used for improving the quality of biodiesel [64].

11.12 Conclusion
Oleochemicals are a capable substrate for renewable and production of
biofuel and it has advantages over fossil fuels due to its carbon renewabil-
ity, but there is an issue due to the threating of food safety and security
and loss of agrobiodiversity when edible feedstocks are used. Techniques
such as transesterification, pyrolysis, and hydrotreatment are the success-
ful process adopted for large-scale biofuel production including biodiesel,
bio-oil, and renewable diesel. This would act as an alternative method for
glycerol production that, in turn, produces a good impact on biodiesel pro-
duction. Biodiesel blend that is available commercially has the potential-
ity to perform as same as Petro-diesel in diesel engines in an eco-friendly
manner with reduced emission of greenhouse gases. This production may
be increased when there is an increased demand for renewable fuel occurs.

Abbreviations
AFW Animal Fat waste
ACPO Acid crude palm oil
DG Diglyceride
FA Fatty acid
FAEE Fatty Acid Ethyl Ester
FAME Fatty Acid Methyl Ester
FFA Free Fatty Acid
ME Methyl ester
MG Monoglyceride
336 Biodiesel Technology and Applications

MSA Methane Sulfonic Acid


MSW Municipal Solid Waste
PTSA p-toluene sulfhonic monohydrate acid
TAG Triacylglycerol
TG Triglyceride
WCO Waste Cooking Oil
WFO Waste Frying Oil

References
1. Sumitkumar Joshi, Pradipkumar Hadiya, Manan Shah, Anirbid Sircar,
Techno-economical and Experimental Analysis of Biodiesel Production
from Used Cooking Oil. Biophy. Econo. Resour. Quality, 4, 2, 2019.
2. Sze Ying Lee, Revathy Sankaran, Kit Wayne Chew, Chung Hong Tan,
Rambabu Krishnamoorthy, Dinh-Toi Chu, Pau-Loke Show., Waste to bio-
energy: a review on the recent conversion technologies. BMC Energy, 1, 4,
2019.
3. Gemma Vicente, Mercedes Martínez, José Aracil., Comparative Study of
Vegetable Oils for Biodiesel Production in Spain. Ene. Fuels, 20, 394–398,
2006.
4. Karlheinz Hill, Fats and oils as oleochemical raw materials. Pure Appl. Chem.,
72, 7, 1255–1264, 2000.
5. Jumat Salimon, Nadia Salih, Emad Yousif, Industrial development and appli-
cations of plant oils and their biobased oleochemicals. Arab. J. Chem., 5,
135–145, 2012.
6. Richtler, H.J., Knaut, J., Henkel KGaA., Challenges to a Mature Industry:
Marketing and Economics of Oleochemicals in Western Europe. JAOCS., 61,
2, 1984.
7. Meira, M., Quintella, C.M., Ribeiro, E.M.O., Silva, H.R.G., Guimarães, A.K.,
Overview of the challenges in the production of biodiesel. Biomass. Conv.
Bioref., 5, 321–329, 2015.
8. Mehdizadeh, A., Handy, L.L., Further investigation of high temperature
alkaline floods. SPE. Reservoir. Eng., 171–177, 1989.
9. Gander, K.F., Fats and Oils as Feedstocks for Oleochemicals. JAOCS., 61, 2
February 1984.
10. Balat, M., Potential alternatives to edible oils for biodiesel production–A
review of current work. Energy. Convers. Manage., 52, 1479–92, 2011.
11. Sanli, H., Canakci, M., Alptekin, E., Predicting the higher heating values of
waste frying oils as potential biodiesel feedstock. Fuel, 2013.
12. Badday, A.S., Abdullah, A.Z., Lee, K.T., Optimization of biodiesel produc-
tion process from Jatropha oil using supported heteropolyacid catalyst and
assisted by ultrasonic energy. Ren. Energy, 50, 427–32, 2013.
Oleochemical Resources for Biodiesel Production 337

13. Kumar, G., Kumar, D., Poonam Johari, R., Singh, C.P., Enzymatic trans-
esterification of Jatropha curcas oil assisted by ultrasonication. Ultrason.
Sonochem., 18, 923–7, 2011.
14. Heidrich, J.F., Oleochemicals: Feedstock or Auxiliary, JAOCS., 61, 2 Feb 1984
15. Peter Adewale, Marie-Josée Dumontn, Michael Ngadinn., Recent trends of
biodiesel production from animal fat wastes and associated production tech-
niques. Ren. Sustain. Ene. Rev., 45, 574–588, 2015.
16. Tong, D., Hu, C., Jiang, K., Li, Y., Cetane number prediction of biodiesel from
the composition of the fatty acid methyl esters. J. Am. Oil. Chem. Soc., 88,
415–23, 2011.
17. Ramírez-Verduzco, L.F., Rodríguez-Rodríguez, J.E., Jaramillo-Jacob, A.d.R.,
Predicting Cetane number, kinematic viscosity, density and higher heat-
ing value of biodiesel from its fatty acid methyl ester composition. Fuel, 91,
102–11, 2012.
18. Adewale, P., Dumont, M.J., Ngadi, M., Rheological, thermal and physico-
chemical characterization of animal fat wastes destined for biodiesel produc-
tion. Energy. Technol., 2, 634–42, 2014.
19. Knothe, G., Some aspects of biodiesel oxidative stability. Fuel. Process.
Technol., 88, 669–77, 2007.
20. Luković, N., Knežević-Jugović, Z., Bezbradica, D., Biodiesel fuel production
by enzymatic transesterification of oils: recent trends, challenges and future
perspectives. Alter. Fuel. Tech. Europe., 47–72, 2011.
21. Hanna Brännström, Hemanathan Kumar, Raimo Alén, Current and Potential
Biofuel Production from Plant Oils. BioEne. Research., 11, 592–613, 2018.
22. Öner, C., Altun, Ş., Biodiesel production from inedible animal tallow and an
experimental investigation of its use as alternative fuel in a direct injection
diesel engine., Appl. Energy, 86, 2114–20, 2009.
23. Teixeira, L.S.G., Assis, J.C.R., Mendonça, D.R., Santos, I.T.V., Guimarães,
P.R.B., Pontes L.A.M., Comparison between conventional and ultrasonic
preparation of beef tallow biodiesel. Fuel. Process. Technol., 90,1164–6, 2009.
24. Kumar, S., Ghaly, A., Brooks, M., Budge, S., Dave, D., Effectiveness of enzy-
matic transesterification of beef tallow using experimental enzyme Ns88001
with methanol and hexane. Enzyme. Eng., 2, 2, 2013.
25. Hsu, A.F., Jones, K., Marmer, W., Foglia, T., Production of alkyl esters
from tallow and grease using lipase immobilized in a phyllosilicate sol–gel.
JAOCS., 78, 585–8, 2001.
26. Colak, S., Zengin, G., Ozgunay, H., Sarikahya, H., Sari, O., Yuceer, L.,
Utilization of leather industry pre-fleshings in biodiesel production. J. Am.
Leather. Chem. Assoc., 100, 137–41, 2005.
27. Alptekin, E., Canakci, M., Sanli, H., Evaluation of leather industry wastes as
a feedstock for biodiesel production. Fuel, 95, 214–20, 2012.
28. Tanwar, D., Tanwar, A., Sharma, D., Mathur, Y.P., Khatri, K.K., Soni, S.L.,
Production and characterization of fish oil methyl ester. IJITR., 1, 209–17,
2013.
338 Biodiesel Technology and Applications

29. Jayasinghe, P., Hawboldt, K., Biofuels from fish processing plant effluents-
waste characterization and oil extraction and quality. Sustain. Energy.
Technol. Assess., 4, 36–44, 2013.
30. Kirakosyan, A., Kaufman, P.B., Plants as Sources of Energy. Recent. Adv. in.
Plant. Biotech., 163–210, 2010.
31. Reznik, G.O., Vishwanath, P., Pynn, M.A., Sitnik, J.M., Todd, J.J., Wu, J.,
Jiang, Y., Keenan, B.G., Castle, A.B., Haskell, R.F., Smith, T.F., Smith, T.F.,
Somasundaran, P., Jarrell, K.A., Use of sustainable chemistry to produce
an acyl amino acid surfactant. Appl. Microbiol. Biotechnol., 86, 1387–1397,
2010.
32. Keng, P.S., Basria, M., Zakaria, M.R., Abdul Rahman, M.B., Ariff, A.B.,
Abdul Rahman, R.N., Salleh, A.B., Newly synthesized palm esters for cos-
metics industry. Ind. Crops. Prod., 29, 37–44, 2009.
33. Abhilash, P.C., Singh, N., Pesticide use and application: an Indian scenario.
J. Hazard. Mater., 165, 1–12, 2009.
34. Christopher, L.P., Kumar, H., Zambare, V.P., Enzymatic biodiesel: challenges
and opportunities. Appl. Energy., 119, 497–520, 2014.
35. Larissa, P., Lima, Francisco, F.P., Santos, Enio Costa, Fabiano, A.N.,
Fernandes., Production of free fatty acids from waste oil by application of
ultrasound. Biomass. Conver. Bioref., 2, 309–315, 2012.
36. Karmakar, A., Karmakar, S., Mukherjee, S., Properties of various plants and
animal feedstocks for biodiesel production. Bioresour. Technol., 101,19, 7201-
10, 2010.
37. Demirbas, A., Biodiesel fuels from vegetable oils via catalytic and non-­
catalytic supercritical alcohol transesterifications and other methods: a sur-
vey. Energy. Convers. Manag., 44, 2093–2109, 2003.
38. Abbaszaadeh, A., Ghobadian, B., Omidkhah, M.R., Najafi, G., Current
biodiesel production technologies: a comparative review. Energy. Convers.
Manag., 63, 138–148, 2012.
39. http://www.bioxcorp.com. BIOX Process., Dec 2008.
40. Adeeb Hayyan, Farouq, S., Mjalli, Mohamed, E.S., Mirghani, Mohd Ali
Hashim, Maan Hayyan, Inas, M., AlNashef, Saeed M., Al-Zahrani., Treatment
of acidic palm oil for fatty acid methyl esters production. Chem. Papers, 66, 1,
39–46, 2012.
41. Hayyan, A., Mjalli, F. S., Hashim, M. A., Hayyan, M., AlNashef, I. M.,
Al-Zahrani, S. M., & Al-Saadi, M. A. Ethane sulfonic acid-based esterifica-
tion of industrial acidic crude palm oil for biodiesel production. Bioresour.
Technol., 102, 9564–9570, 2011.
42. Umer Rashida, Farooq Anwara, Bryan R. Moserb, Samia Ashrafa., Production
of sunflower oil methyl esters by optimized alkali-catalyzed methanolysis.
Biomass. Ener., 32, 1202–1205, 2008.
43. Jeong, G.T., Park, D.H., Kang, C.H., Lee, W.T., Sunwoo, C.S., Yoon, C.H.,
Production of biodiesel fuel by transesterification of rapeseed oil. Appl.
Biochem. Biotechno., 113–116, 747–58, 2004.
Oleochemical Resources for Biodiesel Production 339

44. Antolin, G., Tinaut, F.V., Briceno, Y., Castano, V., Perez, C., Ramirez, A.I.,
Optimization of biodiesel production by sunflower oil transesterification.
Bioresour. Technol., 83, 111–4, 2002.
45. Encinar, J.M., Gonzalez, J.F., Sabio, E., Ramiro, M.J., Preparation and proper-
ties of biodiesel from Cynara cardunculus L. oil. Indus. Eng. Chem. Research.,
38, 2927–31, 1999.
46. Singh, A., He, B., Thompson, J., Process optimization of biodiesel production
using alkaline catalysts. Appl. Eng. Agri., 22, 597–600, 2006.
47. Venkata Subhash, G., Venkata Mohan, S., Sustainable biodiesel production
through bioconversion of lignocellulosic wastewater by oleaginous fungi.
Biomass. Conv. Bioref., 5, 215–226, 2015.
48. Georgios Karavalakis, Georgios Anastopoulos, and Stamos Stournas., Methyl
Ester Production from Sunflower and Waste Cooking Oils Using Alkali-
Doped Metal Oxide Catalysts. Ind. Eng. Chem. Res., 49, 12168–12172, 2010.
49. Hayyan, A., Alam, Md. Z., Mirghani, M. E. S., Kabbashi, N. A., Hakimi, N.
I. N. M., Siran, Y. M., & Tahiruddin, S., Reduction of high content of free
fatty acid in sludge palm oil via acid catalyst for biodiesel production. Fuel.
Processing. Technol., 92, 920–924. 2011.
50. Sevil Özgül-Yücel, Selma Türkay, Variables Affecting the Yields of Methyl
Esters Derived from in situ Esterification of Rice Bran Oil- parameters affect-
ing ME. JAOCS., 79, 6, 2002.
51. Van Gerpen, J., Shanks, B., Pruszko, R., Clements, D., Knothe, G., Biodiesel
Production Technology., Subcontractor Report. Nat. Ren. Ene. Lab./SR.,
510–36244, 2004.
52. Mehdizadeh, A., Handy, L.L., Further investigation of high temperature alka-
line floods. SPE. Reservoir. Eng., 171–177, 1989.
53. Abhijit Samanta, Keka Ojha, and Ajay Mandal., Interactions between Acidic
Crude Oil and Alkali and Their Effects on Enhanced Oil Recovery. Ene.
Fuels, 25, 1642–1649, 2011.
54. Kevin, J., Harrington, Catherine, D’Arcy-Evans., Transesterification in Situ of
Sunflower Seed Oil. Ind. Eng. Chem. Prod. Res. Dev., 24, 314–310, 1985.
55. Fishwick, M. J and Wright, A. J., phytochemistry, 16, 1507, 1977.
56. Rashedul, H.K., Masjuki, H.H., Kalam, M.A., Ashraful, A.M., Ashrafur
Rahman, S.M., Shahir, S.A., The effect of additives on properties, perfor-
mance and emission of biodiesel fuelled compression ignition engine. Ener.
Conver. Manage., 88 348–364, 2014.
57. Kannan, G., Karvembu, R., Anand, R., Effect of metal-based additive on per-
formance emission and combustion characteristics of diesel engine fuelled
with biodiesel. Appl. Energy, 88, 3694–703, 2011.
58. Campanella, A., Rustoy, E., Baldessari, A., Baltaná s, M.A., Lubricants from
chemically modified plant oils. Bioresour. Technol., 101, 245–254, 2010.
59. Rahmat, N., Abdullah, A.Z., Mohamed, A.R., Recent progress on innovative
and potential technologies for glycerol transformation into fuel additives: a
critical review. Renew. Sustain. Energy. Rev., 14, 987–1000, 2010.
340 Biodiesel Technology and Applications

60. Joshi, G., Lamba, B.Y., Rawat, D.S., Mallick, S, Murthy, K., Evaluation of
additive effects on oxidation stability of Jatropha curcas biodiesel blends with
conventional diesel sold at retail outlets. Ind. Eng. Chem. Res., 52, 7586–92,
2013.
61. Karavalakis, G., Hilari, D., Givalou, L., Karonis, D., Stournas, S., Storage sta-
bility and ageing effect of biodiesel blends treated with different antioxidants.
Energy, 36, 369–74, 2011.
62. Schober, S., Mittelbach, M., The impact of antioxidants on biodiesel oxida-
tion stability. Eur. J. Lipid. Sci. Technol.,106, 382–9, 2004.
63. Hamada, H., Kato, H., Ito, N., Takase, Y., Nanbu, H., Mishima, S., Effects of
polyglycerol esters of fatty acids and ethylene-vinyl acetate co-polymer on
crystallization behaviour of biodiesel. Eur. J. Lipid. Sci. Technol., 12, 1323–30,
2010.
64. Kumar, H., Enhancing biodiesel stability by using fuel additives–a review.
Int. J. Res. Biochem. Process. Eng., 1, 2012.
65. Ong, L.K, Kurniawan, A., Suwandi, A.C., Lin, C.X., Zhao, X.S., Ismadji, S.,
Transesterification of leather tanning waste to biodiesel at supercritical con-
dition: kinetics and thermodynamics studies. J. Supercrit. Fluid., 75, 11–20,
2013.
66. Gürü, M, Koca, A., Can, Ö., Çınar, C., Şahin, F., Biodiesel production from
waste chicken fat-based sources and evaluation with Mg based additive in a
diesel engine. Ren.@Energy., 35, 637–43, 2010.
67. Dias, J.M., Alvim-Ferraz, M., Almeida, M.F., Production of biodiesel from
acid waste lard. Bioresour. Technol., 100, 6355–61, 2009.
68. Shin, H.Y., An, S.H., Sheikh, R., Park, Y.H., Bae, S.Y., Transesterification of
used vegetable oils with a Cs-doped heteropoly acid catalyst in supercritical
methanol. Fuel, 96, 572–8, 2012.
69. Bhatti, H.N., Hanif, M.A., Qasim, M., Ataur, R., Biodiesel production from
waste tallow. Fuel, 87, 2961–6, 2008.
70. İşler, A., Sundu, S., Tüter, M., Karaosmanoğlu, F., Transesterification reaction
of the fat originated from solid waste of the leather industry. Waste. Manage.,
30, 2631–5, 2010.
12
Overview on Different Reactors
for Biodiesel Production
V. C. Akubude1*, K.F. Jaiyeoba2, T.F Oyewusi2, E.C. Abbah1,
J.A. Oyedokun3 and V.C. Okafor1
1
Department of Agricultural and Bioresource Engineering, Federal University of
Technology, Owerri, Nigeria
2
Department of Agricultural Engineering, Adeleke University, Ede, Nigeria
3
National Centre for Agricultural Mechanization, Ilori, Nigeria

Abstract
Biodiesel is alternative source of energy which has gained wide research input.
Notwithstanding, its wide utilization in energy systems like fossil fuel is still facing
challenges due to high production cost. This bottleneck can only be eradicated by
developing a better cost-effective and productive approach that can lead to afford-
able end product at the long run. Biodiesel reactor plays a key function in the over-
all production cost of biodiesel; therefore, a careful selection of a reactor that can
help eliminate the challenges faced by convectional reactors is paramount. Batch
reactors are widely used for biodiesel production but its disadvantages has led to
development of continuous flow reactors and even novel reactors like microreac-
tors. This chapter features the general concept of various configuration of reactors
currently been utilized in biodiesel production. Their merit and limitations as well
as their properties are captured.

Keywords: Biodiesel, reactor, batch reactors, continuous flow reactors,


microreactors

12.1 Introduction
Alternative source of energy has been the current center of research in
developing countries to provide the global energy need with its rising pop-
ulation and limited source of petroleum. Also, a cleaner, sustainable, and

*Corresponding author: akubudevivianc@gmail.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (341–360) © 2021 Scrivener Publishing LLC

341
342 Biodiesel Technology and Applications

eco-friendly source of energy [1, 2] is another aim of the research commu-


nity to reduce the pollution caused by the fossil fuel. Biodiesel is energy
product that is gotten from biological sources like animal fat, vegetable, or
waste oil. It is derived from chemical reaction between bio-oil and alco-
hol in the presence of a catalyst [2]. It is a type of biofuel comprising of
mono-alkyl esters of long-chain fatty acids produced by transforming veg-
etable oil via transesterification reaction [3]. It offers lots of advantage to
the environment, society, and economy including lower sulfur and aro-
matic content, non-toxic, renewable, biodegradable, clean burning, reduce
greenhouse gas emission, reduce the carbon monoxide, and good lubricity
[4–8]. Despite the numerous benefits of biodiesel, it has not gain public
acceptance because of the price when compared with fossil fuel. Hence,
to make biodiesel more appealing to the society, well-developed strategy
that is void of the various problems relating to the production processis
essential [9].
Reactor is known as a compartment where chemical conversion takes
place during any processing activity. It is a chamber used in transform-
ing raw materials to desired products via chemical reaction. A reactor
must provide enough time for chemical reaction to occur [10]. Reactor
design should consider the following key factors: construction mate-
rial, stirring system, size of reactor, hydrodynamics, physical properties
of the reactants, and methods for in-take or discharge of heat as this
will ensure high effectiveness and productivity [11]. Several configura-
tions of continuous reactors have been developed. This work reviews
many technological developments within literature with respect to these
reactors.

12.2 Biodiesel Production Reactors


Generally, reactors are classified as batch or continuous [12, 13]. The design
of the reactor depends on how the biodiesel is produced, either in batches
or as a continuous process. Many investigations have been carried out in
order to realize the right type of reactor to surmount some technical chal-
lenges during production of biodiesel by transesterification. In the recent
time, the technical methods or processes in production of biodiesel have
improved tremendously with the advent of various types of reactor in dif-
ferent sizes. This study therefore examines several types of reactor such as
batch reactors (BRs), continuous stirred tank reactors (CSTRs), fixed bed
reactors (FBRs), bubble column reactors (BCRs), microchannel reactors
Different Reactors for Biodiesel Production 343

(MCRs), membrane reactors (MRs), reactive distillation, and hybrid cata-


lytic plasma reactors.

12.2.1 Batch Reactor


Batch production of biodiesel via the use of BR is the basic method that has
been used for biodiesel synthesis [14] before the emerging of continuous
processes [15]. BRs are as simple as tanks with different forms of agitat-
ing equipment as shown in Figure 12.1 [16]. The chamber/tank remains
occupied with the processing reactants (catalyst, oil, and alcohol) with
operation of the agitators for a certain period. In the reactor, the reactants
are charged at a predetermined period of time. The chemical structure in
the reactor transforms with time along with the product concentration
by closing the reactor and then taking it to the required reaction states
(pressures, temperatures, and agitation rates). Subsequently, the result of
reaction inhibition by reaction products can be reduced. After completion
of the reaction time, the chemical substances from the reactor are isolated
and sent for consequent processes. One of the major features of BRs is that
it begins with unreacted substance, cause it to react, and then finish with
reacted substance after a while. Other helpful characteristics of BRs include
suitable mixing features and comparative ease of controlling homogeneous
catalysts as used in the biodiesel transesterification reaction.

Cooling water Condenser


Speed controller

Thermometer and
sampling outlet Cooling water

Agitator/stirrer

Three-neck flask

Hot plate

Figure 12.1 Schematic diagram of batch reactor apparatus [16].


344 Biodiesel Technology and Applications

In some literature, reports have mostly focused on maximizing the


experimental factors of BRs to obtain high biodiesel yield. Price et al. [17]
performed a modeling study on optimal yield of biodiesel in a BR by means
of a liquid lipase formulation. A simulated biodiesel yield of 90.8% (wt.)
was reported at optimum parameters for the model. Leadbeater et al. [18]
reported a typical reaction procedure of a BR which involved two-neck
round bottomed flask of 31 containing a stirrer with soybean oil (1 kg),
sulfuric acid solution (5% by weight, 27 ml) with 630 ml of butanol. This
blend was laid inside hollow of a microwave through an immersion in one
side of the flask necks with a glass connector into one another. In addi-
tion, an optic’s fiber probe was dipped inside the immersion. This blend
was allowed to heat from room temperature to 120°C by means of a pre-
liminary microwave powered at 1,600 W and allowed to stay for 2 min
at 120°C. Then, reaction compartment was detached from the hollow of
the microwave, and upon cooling to normal temperature, the substance
was transferred into another chamber to settle. Thereafter, the substance
was rinsed with water followed by brine. By solution of nuclear magnetic
resonance spectroscopy with respect to hydrogen (HNMR), the degree of
transesterification of oil was measured. The groups of −OCH2in the butyl-
ester-chain with α-methylene proton presents inside the triglyceride (TG)
by-products of the oil were related signs selected for incorporation. After
estimating the conversion completely from the incorporated area of the
signal, 93% conversion of biodiesel was the highest reported.
However, BRs have drawbacks, for instance, bulky reactor size, wide sep-
aration processes, and high cost of labor [19]. More so, the main restriction
of BRs is commonly connected near mass transfer control relating the oil
and solvent [20]. To rise above these challenges, several forms of reactors
were developed with improving usage of solvent, catalyst, and energy; fur-
thermore, to streamline improvement phases by incorporated separation
techniques. The extent of mixing in BRs increases as the totality of energy
presented from the stirrer increases and vice-versa [11, 12].

12.2.2 Continuous Stirred Tank Reactor


Most CSTRs are well configured in order to give room to continuous
pumping of the reactants [21] for efficient working of the reactors as shown
in Figure 12.2 as given by [22]. At first glance, a CSTR seems to be similar
to a BR. Actually, the reactors can be so alike except that CSTR requires
some controls to be set up inside the reactor. At the beginning, after load-
ing the continuous reactor with reaction substances, the process is agitated
adequately until the reaction is accomplished. The major components in
Different Reactors for Biodiesel Production 345

electronic
reactant 1 valve
control level gauge
reactant 2

to vent
system

floor level

products
overflow
from
coolant reactor

stirrer

Figure 12.2 Schematic diagram of continuous stiired tank reactor.

CSTRs include the agitation speed and in-out flow rate which influences
the residence time, rate of mass transfer, and efficiency of mixing [23].
Thorough mixing is required for assurance of adequate homogeneity of
the products; this always caused more energy to be consumed. Thus, many
efforts have been made with the use of different kinds of CSTR to exchange
the mechanical agitators in order to minimize the energy consumed. So
far, CSTR has been extensively employed for production of biodiesel in
large scale due to its adequate technologies and wide understanding of
operations.
While the continuous steady tank reactor worked continuously at a
steady condition, the ideal thing is that the concentrations of some chem-
icals used should be nearly constant all time at everywhere in the reactor.
However, this ideal condition is rare to achieve; hence, adjustment needs
to be done to working parameters to make sure reaction is completed. It
must be noted that, two reactors may be used sometimes. For the two-stage
process, almost 80% of catalyst with alcohol can be combined with oil in
the first stage of CSTR. After which, the reacted flow passes within glycerol
removal phase prior to its entrance to second stage of CSTR. The other 20%
of catalyst with alcohol can be placed inside the reactor at the second stage.
This process offers an extremely completed reaction with the possibility of
346 Biodiesel Technology and Applications

utilizing a smaller amount of alcohol than single phase process. So, due to
consistency of quality products required and inadequate capital with oper-
ational cost for each product, transesterification by CSTR process is prefer-
able to BR process [24]. When comparing CSTR with BR, CSTR provides
quality operation to improve thermal and mass transfer, less production
cost, provides a uniformity of product, and supports scale up [25, 26].

12.2.3 Fixed Bed Reactor


FBR as shown in Figure 12.3 [16], sometimes, called packed bed reactor,
is a cylindrical-tube full of catalytic agents (pellets) packed in a fixed bed
with the flowing of the reactants within the bed which are transformed
to fatty acid methyl esters (FAMEs). It uses heterogeneous catalysts to
simplify actions in improving process since no separation technique
common to the catalyst and product is needed. Also, FBR improves the
heterogeneous reaction of the catalyst because of the delay in deactivation

outlet
stainless steel
column
(Φ25х450mm)

water jacket

pump
inlet
pump

accumulation
tank

feedstock thermostat
tank water bath

Heater

Figure 12.3 Schematic diagram of fixed bed reactor [16].


Different Reactors for Biodiesel Production 347

and longevity in duration of the catalytic agent; hence, there is reduction


in cost of production. However, the main parameters in this reactor are
amount and size of the catalytic agent, feed flowrate, bed height, molar
ratio, and residence time. When comparing with another heterogeneous
catalytic reactor, FBR lowers the reaction time and raises the link combin-
ing bio-oil and alcohol with catalyst; yet, a greater molar ratio is required.
Likewise, the by-product (resultant glycerol) dwells in the bottom most
part of the reactor and accumulates on the surface of the catalytic agent.
Hence, it lowers the effectiveness of the catalyst which causes a further
process removal [27].
Previous works on FBR have been reported by some researchers [28]
who extensively examined the production of FAME from esterification of
acidified oil alongside methanol in FBR with cation exchange resin, and a
better result was attained. Also, esterification of crude triolein with ethanol
in FBR was examined at 60-min residence time, and thereafter, about 80%
of FAME was achieved [29]. In a recent study, esterification of processed
coconut oil using methanol in FBR was carried out with mass transfer as
recovery step and 78% conversion rate was obtained [30]. An optimum
FBR for biodiesel production using soybean oil in a solvent process of
tert-butanol achieved a molar conversion of 83.31%. The FBR can work
over 30 days with no meaningful loss noticed in the substrate conversion.
Also, the FBR shows the ability for industrial-scale production of FAME
utilizing a tert-butanol solvent system. Some common advantages of using
FBR are constant discharge of glycerol with excessive alcohol, effectual
reuse of enzymes, and preservation of particulate enzymes from mechan-
ical shear stress [31–33]. Conversely, short-term operation of FBR is a big
problem; however, using a solvent-free system can operate for 3 to 7 days
without reduction in ester yield [34, 35].

12.2.4 Bubble Column Reactor


This is broadly used in industrial production of biodiesel. It has been gen-
erally utilized for performing gas-to-liquid reaction for many industrial
applications. It is a better reactor for production of biodiesel that requires
a substantial interaction area for gas-to-liquid mass transfer and effective
blending for reacting species. Figure 12.4 [36] gives a schematic diagram of
BCR. The heterogeneous somatic scheme in the BCR is classified into two
fragments: vapor and liquid phase. Vapor phase contains water vapor and
methanol, whereas the liquid phase contains oil, biodiesel, methanol, water
and catalyst (H2SO4). For simplicity, both phases are assumed to be well
348 Biodiesel Technology and Applications

Gaseous Phase
(Unreacted Methanol)
Condenser

Reactor Vessel
(520-620 K, 0.1 MPa)

Liquid Phase
Vegetable Oil

FAME
(Biodiesel Fuel)
Dehydration
Column
Glycerol
Methanol (By-Product)
Heater

Figure 12.4 Schematic view of bubble column reactor for biodiesel production [36].

blended without spatial concentration gradient inside the reactor. Often,


BCR makes use of acid catalysts without soap but produces high tempera-
ture that boils the methanol. The bubbling methanol produces agitation
and removes the by-product water. Production of biodiesel with super-
heated methanol-vapor occurs under atmospheric pressure and higher
temperature between 250°C to 300°C, this necessitates for stainless-steel
reactor with insulation [37].
Kocsisova et al. [38] reported that supplying methanol into BCR with
operation at temperature greater than 100oC under atmospheric pressure
supports water ejection from oil during transesterification that enhances
conversion of the final biodiesel. Suwannakaran et al. [39] discovered the
viability of utilizing blended feedstocks [TGs and free fatty acid (FFA)] for
production of biodiesel using solid catalyst (Calcined Tungstated Zirconia)
through BCR at high temperature of 130°C. Conversion of 85% (FFA)
and 22% (TG) was achieved under 2 h of reaction. The fairly low conver-
sion rate achieved is on account of the competition between FFA and TG
during early protonation of their carboxylic functionality following nucle-
ophilic attack by methanol producing biodiesel. Also, BCR has been oper-
ated previously at upper temperatures (250°C to 290°C) by means of palm
oil in deficit of catalytic agent to envisage its functionality with increased
methanol flow rate but low conversion of 60% was recorded [40]. In a sub-
sequent report, an increased conversion rate of 95% was observed at 290°C
but esterification rate increases as temperature increased [41].
Different Reactors for Biodiesel Production 349

In another related development, utilization of BCR in preparing bio-


diesel was reported by Joelianingsih et al. [40, 41]. The studies established
that catalyst free biodiesel can be successfully produced using BCR at
ambient pressure. BCR provides the opportunity of achieving vigorous
agitation as gas is circulated in a sparger that brings out bubbles through a
vertical-ylindrical column [42, 43]. Reduction in enzymes’ wears and tears,
and provision of longer duration using a highly viscous and s­ olvent-free
system agitated by gas bubbling has been reported by Manman et al. [43].
Benefit of this process is that catalytic agent is not needed; thus, esterifica-
tion of oil with high FFA substance can be achieved with no initial treat-
ment. Also, methanol bubbles coming up in the reactor provides adequate
agitation to permit dispersal of methanol steam to enter the oil phase for
reaction; hence, there is reduction in the cost of mechanical agitation [44].
However, the main drawbacks of using BCR in production of biodiesel
are the condition for even bubble size dispersal and gas/vapor delay pro-
file. These drawbacks are as a result of somekey factors including column
height and diameter, geometry of sparger, liquid phase viscosity, and gas
velocity [42]. The problem of improving these factors is the major con-
straint to the use of BCR for industrial-scale production. Also, it exhibits
ill-effect when biocatalyst (enzyme) is used since the enzyme carrier can
be broken-down through mechanical agitation thereby deactivate the cat-
alytic activity.

12.2.5 Reactive Distillation Column


This is a reactor type in which chemical reactions and product separations
occur concurrently in a single unit. Reactive distillation column (RDC)
consists of a reactive section, non-reactive rectifying and stripping sections
as shown in Figure 12.5 [45]. The product removal and formation hap-
pens simultaneously giving rise to its capacity to overbear the equilibrium
thermodynamics of the reaction, reaching high conversion and selectivity.
Therefore, it can serve as a good option to the conventional combination
of reactor and separation units [46–49]. Usually, vegetable oil is fed to the
top of the column and alcohol is fed to the bottom as vapor. The product
formed (biodiesel) is pumped from the bottom while the water by-­product
is distilled from the top [21]. The merits of using this process include
maintaining a high alcohol-to-oil molar ratio, reduction in excess alcohol
usage, and combination of reactor and separation chamber in a single unit
thereby saving cost [50–52]. This technique has been utilized in biodiesel
production by researchers [53].
350 Biodiesel Technology and Applications

Rectifying section
Feed

Reactive section

Stripping section

Figure 12.5 The general configuration of reactive distillation [45].

12.2.6 Hybrid Catalytic Plasma Reactor


The potential of the plasma technology is the utilization of high energetic
electrons because of a collision between the high energetic electrons and
the reaction mixtures. The high energetic electron is generated from a
high-voltage power supply through high-voltage electrode. The advantages
of this plasma electro-catalysis system only require very short time reac-
tion, do not need a catalyst, do not form soap, and do not produce glycerol
as a by-product. However, there was difficulty in regulating the main reac-
tion process during the plasma process due to the action of high energetic
electrons [54]. Figure 12.6 shows a schematic diagram of hybrid catalytic
plasma reactor [16].

12.2.7 Microreactors Technology


They are of different configuration, namely, MCR, MR, microtube micro-
reactor (MM), and oscillatory flow reactors (OFRs). The sizes of these reac-
tors are quite small compared to the conventional types. This will result to
reduction in capital cost.
MCR as shown in Figure 12.7 [55] is made up of channels whose diam-
eters are very small like less than few millimeters [55]. The merits of
MCR are reduced reaction time, fast phase separation, excellent blending,
Different Reactors for Biodiesel Production 351

Oil Tank Methanol Tank


flowmeter

AC/DC High Voltage


Voltage Generator Settler
Regulator (AC/DC)
BIODIESEL
PRODUCT
High Mixing Tank
Voltage
Electrode
Ground
Electrode
Hybrid Catalytic Plasma Reactor

Figure 12.6 A schematic diagram of hybrid catalytic plasma reactor system in the future
works [16].

Outlet

Middle Sheet

Inlet

Zig-zag Mirochannel
(with 90º turn)

Figure 12.7 Schematic diagram of a microchannel microreactor [55].

less cost of scale-up, large surface area–to-volume ratio, effective mass and
thermal, and safe and stable operations. Its demerit is limited capacity [21].
MR as shown in Figure 12.8 [9] is a reactor type was reaction, and mem-
brane-based separation occurs concurrently in the same chamber [21, 56].
It has high selectivity and ability to regulate the blending of components
between two phases and provide high surface areas per unit volume. The
membrane acts as a selective barrier thereby controlling the movement of
352 Biodiesel Technology and Applications

Pressure
guage
Membrane reactor
Heat exchanger
Oil tank Back
pressure
Circulating pump valve

chiller

Methanol tank Mixing tank


Methanol recovery
Biodiesel tank unit

Figure 12.8 Schematic diagram of a membrane reactor for biodiesel production [9].

substances using different mass transfer rates. During production process,


the membrane isolates the glycerol from the product stream [57, 58] or
keeps the unreacted TGs within the membrane [59–61]. Various forms
of membrane that have been utilized in biodiesel production include as
ceramic carbon membranes, polymeric membranes, and polyethersulfone
[59–63]. Study by [56] shows that using ceramic membrane gave a yield of
96.42% under optimal conditions.
MM as shown in Figure 12.9 [55] is a type of reactor where heat and
mass transfer are expected to remain notably reinforced due to its small

High Pressure Pump


(HPLC)

Hot Water Tube Containing


Microreactor Cooling Tube

Mixing
Substrate Unit
Product
(I-Mixer)

Hot Water Hot Water Cold Water Cold Water


Inlet Outlet Inlet Inlet

Ice-Water Bath
Methanol+Catalyst

Figure 12.9 Schematic diagram of microtubemicroreactor [55].


Different Reactors for Biodiesel Production 353

space with a large surface area–to-volume ratio. Resulting in molecular


diffusions via the edge and in the two-phase reaction, it should come to be
much less considerable [64].

12.2.8 Oscillatory Flow Reactors


This is a continuous process reactor type that is made up of tubes enclosing
equally spaced orifice plate baffles on a piston that create an oscillatory
motion generating recirculation flow pattern that ensures effective blend-
ing of fluid inside it, resulting in increased heat and mass transfer. This
piston is electrically or pneumatically operated. It helps to enhance the
residence time for the reaction [65, 66]. Production of biodiesel using an
oscillatory flow biodiesel reactor (OFBR) has been reported to be another
effective continuous process [14].

12.2.9 Other Novel Reactors


Novel reactors such as sonochemical reactor, cavitation reactor, and micro-
wave reactor have been used as intensification technology for biodiesel
production.
Cavitation is the formation of small bubbles that grows and collapses
resulting in discharge of high level of energy [2, 67, 68], generating very
high pressure and temperature gradient which improves the mass transfer
rate of the reaction by creating a state of intense violent movement and
liquid microcirculation currents in the reactor [67–69].
Sonochemical reactor is a novel reactor that works based on the genera-
tion of cavitation using energy from ultrasonic irradiation [68]. The merits
include savings on energy and cost, high yield and selectivity, and safer
operating conditions [70]. Also, studies report that in ultrasound-assisted
process, the operating parameters are meaningfully decreased [71, 72].
However, its demerit is that of erosion and particle shedding at the delivery
tip surface because of the high surface energy intensity [2, 68].
Microwave reactor utilizes microwave radiation in conveying energy
directly into the reactants and thus speeds up the transesterification pro-
cess. It plays a vital function in heating of reactants to the desired tem-
perature [73]. It is more advantageous in terms of energy and time saving,
steady internal heating, and safe operating conditions. It is easy to regulate
and gives higher yield under mild reaction conditions and less environ-
mental pollution [67].
354 Biodiesel Technology and Applications

12.3 Future Prospects


BRs are basic mode of biodiesel production. However, several continuous
process reactors are emerging with specific benefits to tackle the draw-
backs of BRs. Most of the reactors have been utilized at laboratory level
and needs a scale up for large-scale production. Therefore, there is need for
more research on how to improve on the existing technologies for large-
scale production. Also, the cost that will be incurred to build the reactors
for scale up should be considered carefully as this will have effect on the
cost of production of biodiesel.

12.4 Conclusion
Biodiesel are produced using reactor which vary in type, configuration,
and design. Reactor has a direct effect on the biodiesel produced. The var-
ious types of bioreactors include batch, CSTR, FBR, BCR, RDC, hybrid
catalytic plasma reactor, microreactors technology, and OFRs. Each con-
figuration of reactor has varying degrees of merits and demerits; most of
the reactors have been utilized at laboratory level and needs a scale up for
large-scale production thus calling for more research on improvement
of existing technologies for large-scale production considering cost of
production.

References
1. Rahman, M.M., Hazrat, M.A., Rasul, M.G., Mahmudul, H.M. Comparative
evaluation of edible and non-edible oil methyl ester performance in a vehic-
ular engine. Energy Procedia 75, 37–43, 2015.
2. Bhuiya, K.M.M., Rasul, M.G., Khan, M.M.K., Ashwath, N., Azad, A.K.
Prospects of generation biodiesel as a sustainable fuel—Part: 1 selection of
feedstocks, oil extraction techniques and conversion technologies. Renew.
Sustain. Energy Rev. 55, 1109–1128, 2016.
3. Akubude, V. C., Nwaigwe, K. N. Dintwa, E. Production of biodiesel from
microalgae via nanocatalyzedtransesterification process: A review, Materials
Science for Energy Technologies 2, 216–225, 2019.
4. Atabani A.E., Silitonga A.S., IrfanAnjumBadruddin, Mahlia T.M.I., Masjuki
H.H. and Mekhilef S. A comprehensive review on biodiesel as an alternative
energy resource and its characteristics. Renewable and Sustainable Energy
Reviews, 16, 2070–2093, 2012.
Different Reactors for Biodiesel Production 355

5. Daming Huang, Haining Zhou and Lin Lin. Biodiesel: an alternative to con-
ventional fuel. Energy Procedia, 16, 1874–1885, 2012.
6. Rhoda H. G., Iwekumo W. and OkekogeneEfeonah I E. Simulation model for
biodiesel production using nonisothermal (CSTR) mode: Membrane reac-
tor. Chemical and Process Engineering Research, 21–34, 2013.
7. Saxena P., Jawale S. and Joshipura M.H. A review on prediction of properties
of biodiesel and blends of biodiesel. Procedia Engineering 51, 395–402, 2013.
8. Coniglio L., Bennadji H., Glaude P.A., Herbinet O. and Billaud F. Combustion
chemical kinetics of biodiesel and related compounds (methyl and ethyl
esters): experiments and modelling - advances and future refinements.
Progress in Energy and Combustion Science 39, 340–382, 2013.
9. Olagunju, O. A. and Musonge. P. Production of Biodiesel Using a Membrane
Reactor to Minimize Separation Cost. IOP Conf. Series: Earth and
Environmental Science 78, 2017.
10. Gary, F.L. Reactors in Process Engineering, https://www.researchgate.net/
publication/241765470, 2017.
11. Ajala, E., O., Aberuagba, F., Olaniyan, A. M., Ajala, M. A., Sunmonu, M. O.
and Odewole. M. M. Design, construction and performance evaluation of a
mini-scale batch reactor for biodiesel production: A case study of shea butter,
Songklanakarin J. Sci. Technol. 40, 1066–1075, 2018.
12. Gerpen, J. V. Biodiesel processing and production, Fuel Processing Technology,
86,1097–1107, 2005.
13. Knothe, G., Van Gerpen, J. H. and Krahl, J. The biodiesel handbook, AOCS
Press, Champaign, Ill, 2005.
14. Azhari T. I. Mohd. Ghazi, M. F. M. GunamResul, R. Yunus, T. C. Shean Yaw.
Preliminary design of oscillatory flow biodiesel reactor for continuous bio-
diesel production from jatropha triglycerides. Journal of Engineering Science
and Technology 3, 138–145, 2008.
15. Kraai, G. N., Schuur, B., Van Zwol, F., Van de Bovenkamp, H. H., Heeres, H.
J. Novel highly integrated biodiesel production technology in a centrifugal
contactor separator device. Chemical Engineering Journal 154, 384–389, 2009.
16. Buchori, L., Istadi, I., Purwanto, P. Advanced Chemical Reactor Technologies
for Biodiesel Production from Vegetable Oils - A Review. Bulletin of
Chemical Reaction Engineering & Catalysis, 11 (3), 406–430, 2016.
17. Price, J., Hofmann, B., Silva, V.T.L., Nordblad, M. Woodley, J.M. and Huusom,
J.K. Mechanistic modeling of biodiesel production using a liquid lipase for-
mulation. Biotechnol. Prog. 30, 1277–90, 2014.
18. Leadbeater, N.E., Barnard, T.M. and Stencel. L.M. Batch and continuous-flow
preparationof biodiesel derived from butanol and facilitated by microwave
heating. Energy & Fuels. 22(3), 2008.
19. Anton, A. K. and Costin, S. B. (2012). A review of biodiesel production by
integrated reactive separation technologies. Journal of Chemical Technology
and Biotechnology 87, 861–879, 2012.
356 Biodiesel Technology and Applications

20. Abdurakhman Y.B., Putra Z.A. and Bilad M.R. Aspen HYSYS simulation
for biodiesel production from waste cooking oil using membrane reactor.
Proceedings IOP Conference Series: Materials Science and Engineering, 180,
2017.
21. KhairulAzlyZahana, Manabu Kano Technological Progress in Biodiesel
Production: An Overview on Different Types of Reactors Energy Procedia,
156, 452–457, 2019.
22. Department Of Chemistry of York, UK. The essential chemical industry:
chemical reactors http://www.essentialchemicalindustry.org/processes/
chemical-reactors.htm, 2018.
23. Janajreh I. and Al-Shrah M. Numerical simulation of multiple step transes-
terification of waste oilin tubular reactor. Journal of Infrastructure Systems,
22(4), 2016.
24. Madhu, A., Sunny, S., Kailash, S., Chaurasia, S. P. and Dohare, R. K. Biodiesel
YieldAssessment in Continuous-Flow Reactors Using Batch Reactor
Conditions, International Journalof Green Energy, 10:1, 28–40, 2013.
25. Eflita, Y., Moh, E. Y., Diyono, I., Aditya M. N., and Ristiyanti P. The devel-
opment ofthe super-biodiesel production continuously from Sunan pecan
oil through the process of reactive distillation. AIP Conference Proceedings,
1737, 2016.
26. Bao-Quan, Q., Dan, Z., Gen, L., Jian-Zhong, Y., Song, X. and Jiao L.
Processenhancement of supercritical methanol biodiesel production by
packing beds. Bioresource Technology 228, 298–304, 2017.
27. Shinji, H., Ayumi, Y., Naoki, T., Hideo, N. and Akihiko, K. Enzymatic pro-
duction of biodiesel from waste cooking oil in a packed-bed reactor: An
engineering approach to separation of hydrophilic impurities. Bioresource
Technology, 135, 417–421, 2013.
28. Feng, Y., Zhang, A., Li, J., and He, B. A continuous process for biodiesel pro-
duction in a fixed bed reactor packed with cation-exchange resin as hetero-
geneous catalyst. Bioresource Technology, 102, 3607–3609, 2011.
29. Shibasaki-Kitakawa, N., Honda, H., Kuribayashi, H., Toda, T., Fukumura, T.,
and Yonemoto, T. Biodiesel production using anionic ion-exchange resin as
heterogeneous catalyst. Bioresource Technology, 98(2), 416–421, 2007.
30. Edric, C., Millicent, C., Angelo, J., Diamante, Rainier, L.C. and Raymond,
Y.R. Internalmass-transfer limitations on the transesterification of coconut
oil using an anionic ion exchangeresin in a packed bed reactor. Catal. Today
174, 54–58, 2011.
31. Royon, D., Daz, M., Ellenrieder, G. and Locatelli, S. (2007). Enzymatic
production of biodiesel from cotton seed oil using t-butanol as a sol-
vent, Bioresource Technology, 98(3), 648–653, 2007.
32. Nielsen, P. M., Brask, J. and Fjerbaek, L. Enzymatic biodiesel production:
technical and economic considerations, European Journal of Lipid Science
and Technology, 110 (8), 692–700, 2008.
Different Reactors for Biodiesel Production 357

33. Robles-Medina, A., González-Moreno, P.A., Esteban-Cerdán, L. and


Molina-Grima, E.Biocatalysis: towards ever greener biodiesel produc-
tion, Biotechnology Advances, 27(4), 398–408, 2009.
34. Chang, C., Chen, J.H., Chang, C.M. J., Wu, T.T. and Shieh, C.J. Optimization
of lipase-catalyzed biodiesel by isopropanolysis in a continuous packed-bed
reactor using response surface methodology, New Biotechnology, 26 (3–4),
187–192, 2009.
35. Ognjanovic, N., Bezbradica, D. and Knezevic-Jugovic, Z. Enzymatic conver-
sion of sunflower oil to biodiesel in a solvent-free system: process optimiza-
tion and the immobilized system stability. Bioresource Technology, 100(21),
5146–5154, 2009.
36. Hagiwara, S., Nabetani, H and M Nakajima. Non-catalytic alcoholysis pro-
cess for production of biodiesel fuel by using bubble column reactor. Journal
of Physics: Conference Series, 596, 2015.
37. Wulandani, D. Determination on CFD Modeling for Bubble Column Reactor
to ImproveBiodiesel Fuel Production. SustaiN’2010, 2010: 41–44, 2010
38. Kocsisová, T., Cvengros, J., Lutisan, J. High-temperature esterification of fat-
tyacids with methanol at ambient pressure. European Journal of Lipid Science
and Technology, 107(2), 87–92, 2005.
39. Suwannakarn, K., Lotero, E., Ngaosuwan, K. and Goodwin, J. G.
SimultaneousFree Fatty Acid Esterification and Triglyceride Tran­
sesterification Using a Solid Acid Catalyst with in Situ Removal of Water and
Unreacted Methanol. Industrial & Engineering Chemistry Research, 48 (6),
2810-2818, 2009.
40. Joelianingsih, H.S., Nabetani, H., Sagara, Y., Soerawidjaya, T.H., Tambunan,
A.H. and Abdullah, K. Performance of a Bubble Column Reactor for the-
Non-Catalytic Methyl Esterification of Free Fatty Acids at Atmospheric
Pressure. Journal of Chemical Engineering of Japan, 40 (9), 780–785, 2007
41. Joelianingsih, M.H., Hagiwara, S., Nabetani, H., Sagara, Y., Soerawidjaya,
T.H., Tambunan, A.H. and Abdullah, K. Biodiesel fuels from palm oil via the
non-catalytic transesterificationina bubble column reactor at atmospheric
pressure: Akinetic study. Renewable Energy, 33 (7), 1629–1636, 2008.
42. Dyah, W., Fajri, I., Yayan, F., Ahmad, I.S., Hiroshi, N. and Shoji, H.
Modification of biodiesel reactor by using of triple obstacle within the bub-
blecolumn reactor. Energy Procedia, 65: 83–89, 2015.
43. Manman, L., Junning, F., Yinglai, T., Zhen, Z., Ning, Z. and Yong, W. Fast
production of diacylglycerol in a solvent free system via lipase catalyzed
esterification using a bubble columnreactor. Journal of the American Oil
Chemists’ Society (93): 637–648, 2016.
44. Mollenhauer, T., Klemm, W., Lauterbach, M., Ondruschka, B. and Haupt,
J. ProcessEngineering Study of the Homogenously Catalyzed Biodiesel
Synthesis in a Bubble Column Reactor. Industrial and Engineering Chemistry
Research, 49(24), 12390–12398, 2010.
358 Biodiesel Technology and Applications

45. Shinde, G. B., Sapkal, V. S., Sapkal. R.S. and Raut, N. B. Transesterification by
Reactive Distillation for Synthesis and Characterization of Biodiesel., https://
www.researchgate.net/publication/221919441, 2011.
46. Giessler, S., Danilov, R.Y., Pisarenko, R.Y., Serafimov, L.A., Hasebe, S. &
Hashimoto, I. Systematic structure generation for reactive distillation pro-
cesses. Computers & Chemical Engineering, 25 (1), 49–60, 2001
47. Pai, R.A., Doherty, M.F. & Malone, M.F. Design of reactive extraction sys-
tems for bioproduct recovery. AIChE Journal, 48 (3), 514–526, 2002.
48. Chin, S.Y., Mohamed, A.R., Ahmad, A.L. & Bhatia, S. Esterification of pal-
mitic acid with iso-propanol in a catalytic distillation column: Modeling and
simulation studies. International Journal of Chemical Reactor Engineering, 4
(4) 32, 2006.
49. He, B.B., Singh, A.P. & Thompson, J.C. A novel continuous-flow reactor using
reactive distillation for biodiesel production. Transactions of the ASABE, 49
(1), 107–112, 2006.
50. Omota, F., Dimian, A.C. & Bliek, A. Fatty acid esterification by reactive dis-
tillation. Chemical Engineering Science. 58, 3159–3174, 2003
51. Singh, A.P., Thompson, J.C. & He, B.B. In A continuous-flow reactive distil-
lation reactor for biodiesel preparation from seed oils, ASAE/CSAE Annual
International Meeting, Ottawa, Ontario, Canada, 1–4 August, 2004; Ottawa,
Ontario, Canada, 2004.
52. Kiss, A.A., Omota, F., Dimian, A.C. & Rothenberg, G. The heterogeneous
advantage: biodiesel by catalytic reactive distillation. Topics in Catalysis, 40
(1), 141–150, 2006.
53. MadhuAgarwal, Kailash Singh, S.P. Chaurasia. Simulation and sensitivity
analysis for biodiesel production in a reactive distillation column. Polish
Journal of Chemical Technology, 14(3), 59, 2012.
54. Istadi, I., Yudhistira, A.D., Anggoro, D.D., Buchori, L. Electro-catalysis sys-
tem for biodiesel synthesis from palm oil over dielectric-barrier discharge
plasma reactor. Bull. Chem. React. Eng. Catal., 9(2): 111–120, 2014.
55. Akansha Madhawan & Arzoo Arora & Jyoti Das & ArindamKuila & Vinay
SharmaMicroreactor technology for biodiesel production: a review. Biomass
Conversion and Biorefinery, https://doi.org/10.1007/s13399-017-0296-0,
2017.
56. Motahareh V., Mohammad R. S.E. and Kambiz T. A Novel Membrane
Reactor for Production of High-Purity Biodiesel European Online Journal of
Natural and Social Sciences, 3(3), 2014.
57. Guerreiro L., Castanheiro J.E., Fonseca I.M., Martin-Aranda R.M., Ramos
A.M., Vital J. Transesterification of soybean oil oversulfonic acid function-
alised polymeric membranes. Catal Today, 118(1–2):166–171, 2006
58. Saleh J., Tremblay A.Y., Dubé M.A. Glycerol removal from biodiesel using
membrane separation technology. Fuel, 89(9):2260–2266, 2010.
59. Dubé, M. A., Tremblay A.Y., Liu J. Biodiesel production using a membrane
reactor. BioresourTechnol, 98(3):639–647, 2007.
Different Reactors for Biodiesel Production 359

60. Cao P., Dubé M.A., Tremblay A.Y. High-purity fatty acid methyl ester pro-
duction from canola, soybean, palm, and yellow greaselipids by means of a
membrane reactor. Biomass Bioenergy, 32(11): 1028–1036, 2008.
61. Baroutian S., Aroua M.K., Raman A.A.A., SulaimanN.M..N. A packed bed
membrane reactor for production of biodiesel usingactivated carbon sup-
ported catalyst. BioresourTechnol, 102(2): 1095–1102, 2010.
62. Barredo-Damas S, Alcaina-Miranda M.I., Bes-Piá A., Iborra-Clar M.I.,
Iborra- Clar A., Mendoza-Roca J.A. Ceramic membrane behavior in textile
wastewater ultrafiltration. Desalination, 250(2): 623–628, 2010.
63. Achmadin LM, Misri G, Mohammad N, Mohammad N, Siswa S, Young JY
Membrane microreactor in biocatalytictransesterification of triolein for bio-
diesel production. Biotechnol Bioprocess Eng, 15:911–916, 2010.
64. Afia Mehboob, Shafaq Nisar, Umer Rashid, Thomas Shean Yaw Choong,
Talha Khalid and Hafiz Abdul Qadeer. Reactor designs for the production
of biodiesel. International Journal of Chemical and Biochemical Sciences,
10(2016):87–94.
65. Mackley, M. R. Process innovation using oscillatory flow within baffled
tubes. Trans IchemE, 69(A), 197–199, 1991.
66. Stonestreet, P., and Harvey, A. P. A mixing-based design methodology for
continuous oscillatory flow reactors. Institution of Chemical Engineers, 80(A),
31–44, 2002.
67. Qiu, Z., Zhao, L., Weatherley, L. Process intensification technologies in con-
tinuousbiodiesel production. Chem. Eng. Process, 49, 323–330, 2010.
68. Gogate, P. R., Kabadi, A. M. A review of applications of cavitation in bio-
chemicalengineering/biotechnology. Biochem. Eng. J. 44, 60–72, 2009.
69. Kelkar, M. A., Gogate, P. R., Pandit, A. B. Intensification of esterification
ofacids for synthesis of biodiesel using acoustic and hydrodynamic cavita-
tion. Ultrason. Sonochem. 15, 188–194, 2008.
70. Colucci, J., Borrero, E., Alape, F. Biodiesel from an alkaline transesterifica-
tionreaction of soybean oil using ultrasonic mixing. J. Am. Oil. Chem. Soc.
82,525–530, 2005.
71. Deshmane, V. G., Gogate, P. R., Pandit, A. B. Ultrasound-assisted synthesisof
biodiesel from palm fatty acid distilate. Ind. Eng. Chem. Res., 48, 7923–7927,
2008.
72. Kalva, A., Sivasankar, T., Moholkar, V. S. Physical mechanism of ultrasoun-
dassistedsynthesis of biodiesel. Ind. Eng. Chem. Res., 48, 534–544, 2008.
73. Barnard, T. M., Leadbeater, N. E., Boucher, M. B., Stencel, L. M., Wilhite,
B. A., Continuous-flow preparation of biodiesel using microwave heating.
Energy.Fuel. 211777–1781.
13
Patents on Biodiesel
Azira Abdul Razak, Mohamad Azuwa Mohamed* and Darfizzi Derawi

Department of Chemical Sciences, Faculty of Science and Technology, Universiti


Kebangsaan Malaysia, UKM Bangi, Malaysia

Abstract
Biodiesel is a renewable fuel that is made from edible or non-edible feedstock
and it has high potential to compete with the diesel fuel. In this paper, the pub-
lished patents from 2003 to 2018 related to biodiesel were reviewed, and they were
retrieved from Justia Patent Database. This paper is divided into five sections:
started with the introduction to biodiesel, the generation of biodiesel that depends
on the feedstock used, the development of catalysts, the latest method for biodiesel
production, and the technology of reactor used to produce biodiesel. The second
generation of biodiesel has gained interest as it can be produced from waste which
lowering the production cost. The use of a catalyst can facilitate the reaction, but
it is crucial to choose the best type of catalyst. There are various methods used to
produce biodiesel, and the most common method used is transesterification due
to its simplest and efficient method for biodiesel production. To produce biodiesel
in large amounts, the reactor’s technology is used and it was found that continuous
stirred tank reactor is a more practical and simpler reactor that can be used.
Keywords: Biodiesel, feedstock, catalyst, production method, reactor technology

13.1 Introduction
The change in the global climate is one of the most serious problems faced
by the world nowadays. The major contribution to this problem is the high
utilization of fossil fuel, where the burning of them leads to the release of
carbon dioxide (CO2) to the atmosphere and cause global warming [5].
Based on the data obtained by International Energy Agency, the utilization

*Corresponding author: mazuwa@ukm.edu.my

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (361–376) © 2021 Scrivener Publishing LLC

361
362 Biodiesel Technology and Applications

of energy continues to increase within the years, and it is reported that the
energy demand for oil showed the highest amount, where it has increased
from 3,232,737 kiloton (1990) to 4,449,499 kiloton (2017). High demand
for fossil fuel not only becomes the environmental issue, but it is also a
non-renewable source of energy. If the consumption continues to increase,
then fossil fuel is expected to be completely used up by 2050 [1–3].
To overcome the problems discussed earlier, biofuels show a high poten-
tial to be an alternative solution to replace the usage of fossil fuels as energy
and expected to satisfy the global energy demand. Biofuels are known as
liquid or gaseous fuels produced through a biological and chemical process
or derived from biomass of living organisms such as microalgae, plants,
and bacteria and used for transport sectors [2]. Biodiesel is one of the types
of biofuel that most commonly used in European Countries. Biodiesel can
be produced by converting fats or oils from animal fat and vegetable oil
chemically to esters because they have similar properties to mineral diesel.
U.S. Energy Information Administration indicated that the United States
(U.S.) showed the highest amount of biodiesel production, which is 6.9
billion liters [4]. Indonesia has become the leading for Asian Countries,
which produced 4.0 billion liters of biodiesel and fall as the third-highest
production of biodiesel in 2018.

13.2 Generation of Biodiesel


Biodiesel can be classified into different generations depending on the
feedstock used either edible, non-edible, or waste oil which determines the
qualities of biodiesel produced. It also depends on the major sources of
feedstock from the countries to produce biodiesel. For example, the com-
mon feedstock used in the U.S. is refined vegetable oil such as soybean and
canola oil. However, they also produced biodiesel from used cooking oil
and animal fats to cut the production cost [4]. The patent for biodiesel pro-
duction from first to the fourth generation, which is still in an early stage.
The first generation of biodiesel is quite popular at the beginning of
the biodiesel era, where edible oil used as feedstock to produce biodiesel.
For E.U. and U.S., rapeseed oil and soybean oil are widely used, while for
Asian countries such as Malaysia and Indonesia, palm and coconut oil is
the most common feedstock used. The difference in feedstock used will
affect the production process, quality, and cost because the produced bio-
diesel depends on the types of fatty acid that attached to the triglycerides
molecules [6]. Saka and Kusdiana (2001) have proposed the transesterifi-
cation of rapeseed oil in supercritical methanol without catalyst consumed.
Patents on Biodiesel 363

They found that by using the new supercritical methanol, the process
became shorter in time and has a simpler purification procedure [7].
However, there is a limitation for this generation because of the competi-
tion between food supplies that increase the production cost is the main
drawback for edible oil-based biodiesel. The disadvantages showed by the
first generation of biodiesel caused the development of the second gener-
ation of biodiesel. The feedstock for second-generation biodiesel is non-­
edible oil such as Neem oil, Jatropha oil, Karanja oil, Rubber seed oil, and
waste oil. This type of biodiesel has advantages compared to the first gen-
eration because there is no competition between the food supply and also
less production cost [8].
However, there is still a need for land for growth which causes competi-
tion between food and fiber production. The high competition will lead to
limited nutrient supply for the land and it will be hard to maintain the qual-
ity of oil produced [9]. Algae-based biodiesel, known as third-­generation
biodiesel, has gained much attention to overcome the disadvantages of the
first and second generation of biodiesel. They could produce less green-
house effect, elevated growth rate and productivity, lesser struggle toward
farming land, and a higher amount of oil. However, the limitations for
third-generation biodiesel are higher production cost, high need for sun-
light, need to produce on large scale, and difficult for oil extraction.

13.3 Development of Catalyst


Catalyst is a substance that facilitates a chemical reaction, which makes
the reaction go faster without itself being consumed in the reaction.
Catalysis can be divided into three categories, which are homogeneous,
heterogeneous, and bio-catalysis. A homogeneous catalyst involves the
reaction between a catalyst and the substrate in different phases. Recently,
the heterogeneous catalyst is preferable over the homogeneous catalyst in
the production of biofuels. This is because it is easier to separate from the
mixture after the reaction process by filtering [10], their reusability, and
low production cost due to a single-step reaction [11]. The catalyst used
for industrial production of biodiesel is the alkali-based catalyst method,
where Sodium Hydroxide and Potassium Hydroxide are commonly used
as a homogeneous catalyst. By using alkali catalyst, the reaction can be car-
ried out in mild temperature and pressure condition. However, this process
requires additional steps of separating and removing the catalyst used. The
free fatty acid (FFA) produced also will react with the alkali catalyst to
form soap which will be lowering the function of the catalyst [12].
364 Biodiesel Technology and Applications

13.3.1 Homogeneous Catalyst


Homogeneous catalyst refers to a catalytic reaction involving catalyst and
reactant within the same phase that can be divided into two types which
are alkali and acid catalyst. Alkali catalyst (NaOH, KOH) for biodiesel
production is widely used because they are cheap and readily available.
Transesterification of low-grade safflower oil with methanol in the pres-
ence of alkaline-catalyst (NaOH) has produced a high yield of methyl
ester [13]. However, the utilization of sodium or potassium hydroxide as
an alkaline catalyst caused the decrease of biodiesel yield. This is due to
the presence of the hydroxide group originally attached from triglyceride,
which will dissolve with glycerol [14].
An acid catalyst such as sulfuric, sulfonic, and hydrochloric acid is
the most common acid catalyst used in transesterification and preferable
because it can produce a high yield of the product. However, the reaction
rate when using an acid catalyst is too slow and the heat is required to
accelerate the reaction which makes it unsuitable for larger-scale produc-
tion [15]. Other than that, the homogeneous catalyst also difficult to be
separated from the reactant which makes it unsuitable to reuse and pres-
ence acid could lead to corrosion.
The present invention for the homogeneous catalyst is to provide a
neutral, non-corrosive homogeneous catalyst for transesterification of
triglycerides. The design for the catalyst is based on organometallic com-
pounds that contain at least one −OCH3 functional group that will catalyze
the transesterification and esterification efficiently to produce biodiesel.
This invention made to solve the disadvantages of using a homogeneous
catalyst by offering a non-basic and non-acidic homogeneous catalyst [16].

13.3.2 Heterogeneous Catalyst


A heterogeneous catalyst is the type of catalysis that involves different
phases of catalyst and reactant which contrast with homogeneous cata-
lyst. A heterogeneous catalyst is widely used in many industries such as
chemical, petrochemical, agrochemical, and pharmaceutical industries.
The most important factors in choosing good catalysts are the choice of
catalyst support, type of supported catalyst, and interaction between
catalyst and support [17]. In the previous study, several types of hetero-
geneous catalysts have been developed for biodiesel production such as
Amberlyst 36, Purolite CT482, Purolite CT275DR, Amberlyst-15, and
Sulfuric Silica-Acid. In a study by Kuzminnska et al. (2014), they found
that Amberlyst 36, Purolite CT482, and Purolite CT275DR were able to
Patents on Biodiesel 365

produce biodiesel from Trimethylolpropane (TMP) and Oleic Acid [18].


In another study, Akerman et al. (2011) produced heterogeneous catalysts
such as Amberlyst-15 and Silica-Sulfuric Acid for biodiesel production
from TMP and Oleic Acid. They found that the Silica-Acid catalyst shows
better properties and efficient catalysts compared to Amberlist-15 [19].
The present invention provides a heterogeneous catalyst that has high
tolerance toward the presence of water and FFAs in the production of bio-
diesel. The catalyst is designed by a class of zinc and lanthanum oxide het-
erogeneous catalysts, with different ratios of the metal. The metal catalyst
will give effect to both esterifications of fatty acid and transesterification of
oil simultaneously [20].

13.4 Method Producing Biodiesel


The present invention to produce biodiesel and/or glycerine from feed-
stock includes pre-treatment process, esterification of FFA, and transester-
ification of triglycerides. The pre-treatment process involves the separation
of biodiesel feedstock from impurities such as sulfur, phosphorous, phos-
phatides, gums, sterols, metals, and other color bodies. As for esterifica-
tion, the FFA content in biodiesel feedstock will react with methanol in the
presence of a catalyst to form biodiesel and water. Meanwhile, transester-
ification involves the reaction between triglycerides of biodiesel feedstock
with methanol in the presence of a catalyst to produce biodiesel and glyc-
erine [21]. Other methods used for biodiesel production, such as direct use
and blending of oils and pyrolysis methods, also will be discussed further
under this topic.

13.4.1 Pre-Treatment Process


According to U.S. Patent No. 20090071063 (2009), by applying the
pre-treatment process, low quality of feedstock could be used as a source
of FFA to produce good quality of biodiesel. The pre-treatment process
involves the filtration and distillation process, which first started by heat-
ing to 43°C, followed by filtering the feedstock from solid particles to pro-
vide filtrate, which then further distilling repeatedly until the final distillate
is obtained [22]. This process is important to ensure all the impurities are
removed completely from the feedstock before they are used to produce
biodiesel. The production of low-cost biodiesel is a necessity to compete
with petroleum-based diesel. Utilizing waste such as waste cooking oil,
brown greases, and crude corn oil as feedstock could be a better alternative
366 Biodiesel Technology and Applications

for more practical and economical biodiesel. However, high FFA con-
tent in these wastes will become the obstacle which is the reason why the
pre-treatment process becomes necessary to produce biodiesel [23].

13.4.2 Direct Use and Blending of Oils


Direct use of vegetable oils as biodiesel is not preferable because of their
properties due to the effect of cold temperature, which not suitable to be
used as fuel. Therefore, vegetable oil usually will be blended with biodiesel
with a certain concentration. The amount of biodiesel blending with diesel
varies depending on the country which used biodiesel. The most common
are B5 (5% of biodiesel) and B20 (6%–20% of biodiesel). An invention
by U.S. Patent No. 20180223202 proposed four stages of principle which
apply for biodiesel blending. The first principle is an automated method
of blending that provides a distillation stream, separation of distillate, pro-
vides target pf biodiesel content, and measure of actual biodiesel content.
The second principle is an automated system for blending, which applies
the same principle with the addition of an analyzer for measuring actual
biodiesel content. The third and fourth principles are automated methods
and systems for blending, which provide a blending of biodiesel not to
exceed the maximum biodiesel content [24].
Another invention by U.S. Patent No. 7458998 proposed a method of
blended biodiesel by heating the biodiesel and diesel at a temperature
above the cloud point of the first blended. Then, the heated biodiesel will
be added to the unheated diesel to provide the second biodiesel blend.
The blended biodiesel produced by this method is maintained in a heated
environment which compatible with fuel purposes [25]. It is important to
ensure that the blended fuel meets the standard requirement of diesel in
terms of its kinematic viscosity, density, and the flashpoint of the blend.
Arabi et al. (2017) studied the characteristics of palm oil-biodiesel-diesel
fuel blend, where they found that the fuel properties for the blended oil
showed no significant difference compared to diesel fuel up to 30% of the
volume of biodiesel of palm oil [26].

13.4.3 Esterification of FFA


Biodiesel can be produced via an esterification reaction between fatty acid
and methanol with the presence of a catalyst. The most common catalyst used
in esterification is liquid acid such as sulfuric acid, hydrochloric acid, and
p-toluenesulfonic acid (p-TsOH). Esterification is considered as an attractive
method for producing biodiesel because it allows the use of FFA which can
Patents on Biodiesel 367

H2C O R Hydrolysis
O O Esterification O
3 + 3 R OH 3 R + H2O
HC O R R OH R O
O
H2C O R
Triglycerides Fatty acid Alkyl Ester
of Oil (Biodiesel)

Figure 13.1 General reaction for esterification.

be obtained from refinement of vegetable oils and waste oil such as animal
fats and waste frying oil. The use of low-value fatty acid gives the advantage
for biodiesel produced, where it could compete with diesel in terms of their
low cost. Figure 13.1 illustrates the general reaction for esterification.
The invention for the production of biodiesel by esterification of FFA
using niobic acid as a heterogeneous catalyst has been proposed. Long
chain of carboxylic acid-containing 6–24 carbon atoms is used to react with
short-chain alcohol (methanol and ethanol), which act as esterifying agents.
Before the niobic acid catalyst is used in the reaction, it must be calcined to
maximize the acid strength of the catalyst [27]. An invention by U.S. Patent
No. 20080289248 proposed a process of producing biodiesel by involving
the contact of lipid material with alcohol in the presence of Zirconium (IV)
metal salt conjugated to the solid support. The temperature for the process
is between 25°C and 75°C, in which the esterification process runs within
the lipid material. The purpose of this invention is to enhance the conver-
sion of fatty acid and efficiently producing biodiesel [28].

13.4.4 Transesterification of TAG


Transesterification of TAG in biodiesel feedstock is the most common
and widely used method to produce biodiesel, as shown in Figure 13.2.
Transesterification is a reversible reaction, where triglyceride of oil reacted
with short-chain alcohol with the presence of catalyst either a strong base
or strong acid to produce Fatty Acid Alkyl Ester, which is the biodiesel.
In transesterification, the most preferable alcohol used is methanol and
ethanol, especially methanol because of its low cost, polar solvent, and the
shortest chain of alcohol [29]. If methanol is chosen, then the product gen-
erated is Fatty Acid Methyl Ester (FAME), which is the biodiesel that has
been well established in the oleo-chemical industry.
The present invention for the transesterification of biodiesel is by con-
trolling the removal of glycerin, which is the major by-product in biodiesel
368 Biodiesel Technology and Applications

H2C O R H2C OH
O
O
+ 3 R OH 3 R + HC OH
HC O R R O
O

H2C O R H 2C OH

Triglycerides Alkyl Ester Glycerol


of Oil (Biodiesel)

Figure 13.2 General reaction for transesterification.

production. The removal of glycerin can be done in a reaction vessel, either


continuous, semi-continuous, or batch reaction. By removing glycerin from
the biodiesel product, the reaction can be carried out with lower ratios of
alcohol to oil and will increase the rate of production for biodiesel [30].

13.4.5 Pyrolysis
Pyrolysis is defined as a thermochemical reaction which involved reaction
at high temperature (280°C–850°C), with the absence of an oxidizing agent
to produce different products such as solid (biochar), gaseous (biogas), and
liquid (bio-oil) [31]. High oxygen content in biodiesel from the transester-
ification reaction has become a disadvantage as it could lead to corrosion.
In contrast, biodiesel from pyrolysis appears as the solution as no oxygen
involved in the reaction and also it produced hydrocarbon diesel [32]. An
invention by U.S. Patent Application 20070144060 proposed a method
of producing biodiesel from triglycerides feedstock by pre-treating with
thermal cracking or rapid pyrolysis. By applying this method, the con-
taminants from the biodiesel can be removed and will produce distillate
fraction that rich in FFA content [33]. The general reaction of pyrolysis is
shown in Figure 13.3.

H2C O R R + CO2
O

HC O R
O
R + CO + H2O
H2C O R
Triglycerides
of Oil Hydrocarbon

Figure 13.3 General reaction for pyrolysis.


Patents on Biodiesel 369

Table 13.1 Comparison between the pyrolysis method by operating parameters


and product yield [34].
Product yield (%)
Pyrolysis Temperature Residence Heating rate
method (°C) time (s) (°C/s) Oil Char Gas
Slow Medium-High Long Low (10) 30 35 35
(400–500) (450–550)

Fast Medium-High Short (0.5–10) High (100) 50 20 30


(400–650)

Flash High (700–1,000) Very short Very high 75 12 13


(<0.5 s) (>500)

The pyrolysis method is considered a complex process because of


two reactions condition, which is temperature and non-reactive atmo-
sphere that must be kept simultaneously within the reaction. The reaction
involved the breakdown of long chains of carbon, hydrogen, and oxygen
compounds in biomass into smaller molecules. Pyrolysis can be classified
into three main categories: slow, fast, and flash pyrolysis.
Among the methods, fast and flash pyrolysis is considered a better
method for producing biodiesel because these methods are able to produce
high yield of biodiesel with 50 and 75% respectively, in a high reaction
temperature but within a very short residence time as stated in the Table
13.1 [34]. By referring to Takuya (2012), triglycerides of oils decomposed
to fatty acid chains and further decomposed to the hydrocarbon chain at
390°C. By increasing the reaction temperature and heating rate, the prod-
uct yield of oil can be increased [35].

13.5 Reactor’s Technology for Biodiesel Production


General reactors used for biodiesel production are batch reactors, semi-
continuous-flow reactor, and continuous-flow reactor. Reactor becomes an
important technology for producing biodiesel on a large scale with more
economic value. By referring to U.S. Patent No. 20080282606, a biodiesel
reactor is a material that consists of a housing enclosing chamber that con-
tains four parts. The part in a biodiesel reactor is an inlet (for the inflow of
raw material), stir bar (inner part for stirring), baffle (partially segmenting
chamber for mixing), and outflow (for the outflow of the reaction mixture).
It is also stated that the process for producing biodiesel started by selecting
the feedstock oil and followed by measuring alcohol and catalyst need to
370 Biodiesel Technology and Applications

be used. Then, the feedstock, alcohol, and catalyst are mixing in the reactor
to form a mixture. The product and by-product will be separated from the
mixture and further distillate to purify the biodiesel product [36]. For cur-
rent status, there are various types of reactors used nowadays to increase
biodiesel production, which will be discussed further within this topic.

13.5.1 Continuous Stirred Tank Reactor


A continuous stirred tank reactor (CTSR) is an advanced technology for
a reactor that applies the same fundamental processing mechanism as a
batch reactor. CTSR allowed the continuous production of biodiesel in a
simple step, which makes the production cost become lowered. There is
two important part in CTSR: the reactor and phase separator. The first past
of the reactor involved the conversion of alcohol and triglycerides into bio-
diesel. The process followed by the removal of glycerol from the reaction
mixture in the second stage of CTSR. The second stage: phase separator
used to promote the reaction of transesterification through chemical equi-
librium shifting. CTSR has proved to produce a high yield of biodiesel,
which is about 97.3% yield [37]. CTSR has been used widely in industrial
scale due to the simpler and deep understanding operation of CTSR.
An invention from U.S. Patent No. 2005052103 provides an improved
process for the preparation of biodiesel using modified CTSR. The feed-
stock oil is taken into the modified reactor equipped with the alcohol
recycle/recovery system, condenser, thermometer, and feeding funnel. By
using the CTSR, two layers are formed, and the mixture is allowed to settle
for four hours and the top layer is taken for further processing [38].

13.5.2 Fixed Bed Reactor


Fixed bed reactor (FBR) involved the flowing of oil and solvent through a
cylindrical tube filled with catalyst pellets, which packed in static bed for
the conversion of biodiesel shown in Figure 13.4. A heterogeneous cata-
lyst is used in FBR because of their simple recovery steps which do not
involve any separation process between product and catalyst. Other than
that, the heterogeneous catalyst used in FBR would enhance the reaction
due to slow deactivation and longer durability of the catalyst. By offering
these advantages, the production cost can be reduced. However, the higher
molar ratio is needed during the process due to the remaining glycerol (by-­
product), which will remain at the bottom of the reactor and will adsorb on
the surface of the catalyst. This condition will affect the catalyst efficiency
and an additional removal process will be needed [39].
Patents on Biodiesel 371

Condenser

Thermometer
Sampling
Biodiesel

Separator

Feedstock Alcohol and Reactor


oil catalyst
Glycerol
Heating and Stirring

Figure 13.4 Schematic diagram of a single reactor CTRS. Reproduced with permission
from [37]. Copyright 2019 Elsevier.

The invention proposed by U.S. Patent No. 2011/0245551A1 introduced


the process for hydroprocessing an acidic biomass feedstock in a guard
bed, which aims to prevent undesired polymerization from occurring. The
guard bed reactor is used to saturate the olefins with hydrogen and prevent
the olefins and other compounds from polymerizing. The reactor, which
can be either a noble metal or non-metal catalyst, is run at low tempera-
tures and able to recycle back the product into the reactor [40].

13.5.3 Micro-Mixer Reactor


Micro-mixer reactor also known as micro-reactors and micro-
structured reactors has been used widely in biodiesel production. The
micro-mixer reactor uses numerous channels with millimeter range
in diameter, which allowed better separation of oil. This is because of
high surface-to-volume ratios and low distance between heat and mass
transfer for diffusion. An invention from U.S. Patent No 8.404,005 stated
the method and system to produce biodiesel using an improved cat-
alytic transesterification process. In this process, the first and second
reactant is dispersed to form a laminar slug flow pattern within the
micro-channel of microreactor and both of them become immiscible.
The mixing between the reactant will trigger the reaction between them
to produce biodiesel and glycerol [41].
372 Biodiesel Technology and Applications

13.6 Conclusion
This paper has discussed and analyzed patents related to biodiesel, and the
following conclusion can be made. The demand for fuel energy continues
to increase each year and biodiesel has become a solution to replace diesel
fuel for a more sustainable and greener environment. The use of feedstock
for biodiesel production must come from non-edible feedstock to avoid
the competition between food and fuel. For the current status of biodiesel
production, the second generation of biodiesel, which is from waste, is
more preferable because it does not only produce low-cost biodiesel but
also helps in properly managing waste. In biodiesel production, the use of
catalyst is crucial to fasten the reaction and to produce a better quality of
biodiesel. A heterogeneous catalyst is more suitable to use as it is easier to
separate after the reaction and it can be reused for five to six cycles of reac-
tion. There are few methods for biodiesel production that have been dis-
cussed, and most of the patents have been made for the transesterification
process. This shows that the trend nowadays is more to using transesteri-
fication as a method to produce biodiesel due to the easier reaction which
will lower the production cost. Lastly, the utilization of the reactor in pro-
ducing biodiesel is very helpful for the industrial scale. Among the reactor
used, the CTSR has been used by many biodiesel industries because of the
operation method is more practical.

References
1. Abideen. A., Sustainable biofuel production from non-food sources-An
overview. Journal of the Science of Food and Agriculture., 26 (12), 1057–1066,
2014.
2. Rodionova. M.V et al., Biofuel production: Challenges and opportunities.
International Journal of Hydrogen Energy., 42, 8450–8461, 2017.
3. International Energy Agency, Data and Statistic, https://www.iea.org/
data-andstatistics?country=WORLD&fuel=Energy%20supply&indicator=-
Total%20primary%20energy%20supply%20(TPES)%20by%20source, 2017.
4. U.S. Energy Information Administration, Biodiesels produced from certain
feedstocks have distinct properties from petroleum diesel, https://www.eia.
gov/todayinenergy/detail.php?id=36052, 2018.
5. Alalwan, H. A., Alminshid, A.H & Aljaafari, H. A. S, Promising Evolution of
Biofuel Generations. Subject Review. Renewable Energy Focus., 28, 127–139,
2019.
6. Ho, D.P., Ngo, H.H., Guo, W., A mini review on renewable sources for bio-
fuel. Bioresource Technology., 169, 742–749, 2014.
Patents on Biodiesel 373

7. Saka, S., Kusdiana, D., Biodiesel fuel from Rapeseed oil as prepared in super-
critical methanol. Fuel., 80(2), 225–231, 2001.
8. Singh, D., Sharma, D., Soni, S.L., Sharma, S., Sharma, P.K., Jhalani, A., A
review on feedstocks, production processes, and yield for different genera-
tions of biodiesel. Fuel., 262, 2020.
9. Sims, R.E.H., Mabee, W., Saddler, J.N., Taylor, M., An overview of second
generation biofuel technologies. Bioresource Technology., 101, 1570–1580,
2010.
10. Martinez, A., A novel green one-pot synthesis of biodiesel from Ricinus com-
munis seeds by basic heterogeneous catalysis. Journal of Cleaner Production.,
196, 340–349, 2018.
11. Gopinath, S., Efficient mesoporous SO4 2/Zr-KIT-6 solid acid catalyst for
green diesel production from esterification of oleic acid, Fuel., 203, 488–500,
2017.
12. Y. Ueki and M. Tamada, Catalyst for Production of Biodiesel and Its
Production Method, and Method for Producing Biodiesel, US Patent No
20100170145, 2011.
13. Math, M.C., Chandrashekhara, K.N., Optimization of Alkali Catalyzed
Transesterification of Safflower Oil for Production of Biodiesel. Journal of
Engineering., 8, 86–92, 2016.
14. Vyas, A.P., Shukla, P.H., Subrahmanyam, N., Production of Biodiesel
using Homogeneous Alkali Catalyst and its Effect on Vehicular Emission.
International Conference on Current Trends in Technology., 382–481, 2011.
15. Zheng, S., Kates, M., Dube, M.A., Acid-catalyzed production of biodiesel
from waste frying oil. Biomass Bioenergy., 30(2), 67–72, 2006.
16. S. C. Yang, J.R. Chang, M.T. Lee, T. B. Lin, F.M. Lee, C.T. Hong and J.C. Lee,
Homogeneous Catalysts for Biodiesel Production, US Patent No 8624073,
assigned to CPC Corporation, Taiwan, 2014.
17. Bailie, J. E., Hutchings, G. J., O’Leary, S., Supported Catalysts. Encyclopedia of
Materials: Science and Technology, pp. 8986–8990, 2001.
18. Kuzminska, M., Backov, R., Gaigneaux, E.M., Complementarity of
Heterogeneous and Homogeneous Catalysis for Oleic Acid Esterification with
Trimethylolpropane Over Ion-Exchange Resins. Catalysis Communication.,
14, 566–736, 2014.
19. Akerman, C.O., Gaber, Y., Ghani, N.A., Lamsa, M., Kaul, R.H., Clean syn-
thesis of biolubricants for low temperature applications using heterogeneous
catalyst. Journal of Molecular Catalysis B: Enzymatic., 72, 263–269, 2011.
20. S. Yan, S. O. Salley and K. Y. S. Ng, Methods and Catalysts for making
Biodiesel from the Transesterification and Esterification of Unrefined Oils,
US Patent No 8163946, assigned to Wayne State University, 2012.
21. G. Shah and S. Francisco, Transesterification of Biodiesel Feedstock with
Solid Heterogeneous Catalyst, US Patent 8580119, assigned to Menlo Energy
Management, 2013.
374 Biodiesel Technology and Applications

22. M. J. Perrier, Process and System for producing biodiesel fuel, US Patent
20090071063, assigned to Next Energy System, 2009.
23. Tafesh, A., Basheer, S., Pre-treatment Methods in Biodiesel Production
Processes. Green Energy and Technology., pp. 417–434, 2013.
24. R. Fransham and C. Robbins, Controlled Blending of Biodiesel into Distillate
Streams, US Patent 20180223202, 2018.
25. K. Copeland, R. Hardy, J. Johnson, C. Selvidge and K. Walztoni, Blending
Biodiesel with Diesel Fuel in Cold Locations, US Patent 7458998, assigned to
Flint Hills Resources, 2008.
26. Arabi, R., Amin, A., Morsi, A.K., Ibiari, N.N., Diwani, G.I., Study on the
characteristics of palm oil–biodiesel–diesel fuel blend. Egyptian Journal of
Petroleum., 27(2), 187–194, 2017.
27. A. T. Pereira, K. A. Oliveira, R. S. Monteiro, D. A. G. Aranda, R. T. P. Santos
and R. R. Joao, Production Process of Biodiesel from the Esterification of
Free Fatty Acids, US Patent 20070232817, applied to Companhia Brasileira
De metalurgia E Mineracao, 2007.
28. Y. Gao, Immobilized Esterification Catalyst for Producing Fatty Acid Alkyl
Esters, US Patent 20080289248, applied to Southern Illinois University
Carbondale, 2008.
29. Keera, S.T., Sabagh, S.M.E., Taman, A.R., Transesterification of vegetable oil
to biodiesel fuel using alkaline catalyst. Fuel., 90, 42–47, 2011.
30. J. Crawford, J. Crawford and R. Crafts, Transesterification of Oil to form
Biodiesel, US Patent 20070232818, applied to Domestic Energy Leasing,
2007.
31. Gopal, P. M., Sivaram, N. M., Barik, D., Paper Industry Wastes and Energy
Generation from Wastes Energy from Toxic Organic Waste for Heat and
Power Generation, 7, 83–97, 2019.
32. Abdelfattah, M.S.H., Osayed S.M.AA., AbdElmawla, E., Marwa, A., On
Biodiesels from Castor Raw Oil using Catalytic Pyrolysis. Energy., 143, 950–
960, 2017.
33. M. Ikura, Production of Biodiesel from Triglycerides via a Thermal Route,
US Patent Application 20070144060, 2007.
34. Jahirul, M.I., Rasul, M.G., Chowdhury, A.A., Ashwath, N., Biofuels
Production through Biomass Pyrolysis-A Technology Review. Energies., 5,
4952–5001, 2012.
35. Takuya, I., Yusuke, S., Yusuke, K., Motoyuki, S., Katsumi, H., Biodiesel
production from waste animal fats using pyrolysis method. Fuel Processing
Technology., 94:47–52, 2012.
36. J. P. Plaza and B. L. Goodall, System and Process for Producing Biodiesel, US
Patent Application 20080282606, 2008.
37. Wong, K.Y., Han, N.J., Chong, C.T., Lam, S.S & Chong, W.T., Biodiesel pro-
cess intensification through catalytic enhancement and emerging reactor
designs: A critical review. Renewable and Sustainable Energy Reviews., 116,
109399, 2019.
Patents on Biodiesel 375

38. Velappan, Kandukalpatti, Chinnaraj, Saravanan, Subramani, Vedaraman,


Nagarajan, Paruchuri and Gangadhar, an Improved Process for the
Preparation of Bio-Diesel, US Patent 2005052103, 2003.
39. Zahan, K.A., Kano, M., Technological Progress in Biodiesel Production: An
Overview of different types of Reactor. Energy Procedia., 156, 452–457, 2019.
40. T. L. Marker, P. Height, T.A. Brandyold, A. Height and C.P. Leubke, Use of
a Guard Bed Reactor to Improve Conversion of Biofeedstocks to Fuel, US
Patent 2011/0245551A1, assigned to UOP LLC, 2011.
41. B. H. Dennis, R. E. Billo, C. R. Oliver, J. W. Priest, E. S. Kolesar and E. Kolesar,
Methods and Systems for Improved Other Publications Biodiesel Production,
US Patent 8.404,005, assigned to Board of Regents, The University of Texas
System, 2009.
14
Reactions of Carboxylic Acids With
an Alcohol Over Acid Materials
J.E. Castanheiro *

MED, DQ, ECT, Universidade de Évora, Évora, Portugal

Abstract
Biodiesel is an alternative to fossil diesel. It is biodegradable and non-toxic. The
production of biodiesel is carried out by transesterification and/or esterification
reactions.
Usually, reactions between a carboxylic acid with an alcohol are carried out
using sulfuric and phosphoric acids. Mineral acids (e.g., sulfuric acid) have some
problems, such as the separation of reaction mixture. Heterogeneous catalysts can
used to replace the homogenous ones.
These catalysts have been used for biodiesel production. Recently, acid solids,
such as heteropolyacids, activated carbons with sulfonic acid groups, and cation-­
exchange resins, have been used as catalysts on esterification of free fatty acids (FFAs).
This present work is a review about the esterification reactions of carboxylic
acid (fatty acid) with an alcohol over solid materials.

Keywords: Fatty acids, esterification, heterogenous catalysts, biodiesel

14.1 Introduction
Traditionally, the biodisel is obtained by reaction between one molecule
of triglyceride in conjunction with three molecules of an alcohol, using
NaOH as catalyst. However, if the quantity of carboxylic acid presents in
fat is high, processes will become inefficient, due to the soap formation. It
is necessery an esterification reaction step before the transesterification.

Email: jefc@uevora.pt
*

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (377–388) © 2021 Scrivener Publishing LLC

377
378 Biodiesel Technology and Applications

Esterification of a molecule of free fatty acid (FFA) with a molecule of alco-


hol (methanol or ethanol) yields an ester molecule and water molecule.
Figure 14.1 shows a scheme of a fatty acid esterification with an alcohol
(methanol, ethanol, 1-propanol, and 1-butanol) [1, 2].

R1– COOH + R2 – OH ↔ R1 – COOR2 + H2O

Figure 14.1 Scheme of reaction between a carboxylic acid and an alcohol.

where R1 represents a carbon atoms and R2 represents an alkyl group of


the alcohol.
The reactions of carboxylic acids with a alcohol are performed using
homogeneous substances (one phase formed with reactants and catalyst)
and heterogeneous catalysts (two phases: the catalysts are solid and the
reactants are liquid) [1, 2].
To become the biodiesel production as a “eco-friendly processes”, the tradi-
tional sulfuric acid has been substituted for solid catalysts. The homogeneous
catalysts are corrosive, and it is neutralized after reaction. The heterogeneous
catalysts are more environmentally friend and can be reused [3–5].
This chapter focuses on solid acid materials on esterification of FFA.

14.2 Zeolites
Zeolites are aluminosilicates crystalline with micropore system. These
catalysts have an important propriety: shape selective. Due to the surface
acidity, the zeolites can be used as heterogenous catalyst in the synthesis
of different areas such, fine chemistry, and petrochemical industry [6–8].
However, only zeolites with large porous are applied on reactions between
carboxylic acids with alcohols. The zeolite’s activity improves with the
increases of the ratio silicon to aluminum, which can be an indication that
the zeolite reactivity is influenced by the hydrophobic/hydrophilic balance
of its surface The activity of zeolites for esterification reactions is affected
by the aluminum amount present in its framework [6–9].
The reaction between hexadecanoic acid and methanol was done over
four zeolites (H-Y, ZSM-5) as catalysts [10]. When the reaction was carried
out without zeolite (Blank), the conversion was lower than with zeolite.
With H-Y-60 zeolite, a conversion of 100% was obtained. This behavior
was explained due to difference on zeolite’s porous system. H-ZSM-5
contains medium-sized micropores, with avarege pore size (5.6 Å × 5.3
Å channels) lower than H-Y zeolite (7.4 Å). The reaction over H-ZSM-5
Reactions of Carboxylic Acids 379

zeolites, probably, occurs at the external surface area, due to the molec-
ular size of ester. In contrast, H-Y zeolites offer high surface and porous
volume. The H-Y-60 catalyst showed the highest catalytic activity for this
reaction. Under optimazed reaction conditions, a 100% conversion was
obtained. The H-Y-60 zeolite can be recycled [10].

14.3 SO3H as Catalyst


The reaction between oleic acid and ethanol was performed using sulfon-
ated polystyrene [11]. This reaction without polystyrene yielded only about
10% of the ethyl oleate. When it was used sulfonated polystyrene, the yield
increased from 30% to 90%. The optime reaction conditions were found to
be amount of ethanol (219.0 mmol), amount of oleic acid (7.45 mmol), and
amount catalyst (25 mg). Under this reaction conditions, the ethyl oleate
yield was about 90%. These catalysts can be reused several times [11].
When octadec-9-enoic acid reacted together with methanol over
Amberlyst-15, using a continuous reactor, over of Amberlyst-15, the con-
version increased with reaction temperature [12]. The reaction conditions
were as follows: methanol flow 6.4 ml/h, oleic acid flow 9.0 ml/h, and bed
height of 10 cm, and the temprature was 80°C, 100°C, and 120°C. Under
this conditions, the yields were 69.9%, 72.7%, and 84.0%. It was observed
that the catalyst does not loss activity during the reaction [12].
The esterification of palmitic acid with methylic alcohol was done over
PVA cross-linked with sulphosuccinic acid (SSA) and PS cross-linked
with divinylbenzene containing −SO3H [13]. It was observed that the PVA
catalysts showed high activity than the PS. PVA_SSA40 material showed
a good stability [13]. This reaction was also studied over chitosan with
sulfonic acid [14]. Catalyic actvity of these materials increased when the
amount of sulfonic groups increases, until a maximum (sample CT4). The
catalyst sample CT4 showed a good catalytic stability [14].
The reaction of octadec-9-enoic acid and n-butanol was done over poly-
styrene-modified organosilicas with sulfonic acid groups [15]. These mate-
rials demonstrated high activity. The optimal conditions were T = 130°C,
catalyst amount = 2 wt%, and molar ratio oleic acid to n-butanol 1:1.4,
which led a oleic acid conversion about 92% [15].
Activated carbons with −SO3H have been utilized as material for reac-
tions between carboxylic acids with alcohols. These materials can be
obtained from sugar, starch, or cellulose [16, 17].
The reaction between octadec-9-enoic acid and octadecanoic acid and
ethanol were performed using activated carbons prepared from d-glucose
380 Biodiesel Technology and Applications

with −SO3H. The materials exhibited high activity, probably by the high
quantity of −SO3H groups [17].
The reaction between octadec-9-enoic acid and ethyl alcohol was done
over carbons (CMK) with −SO3H [18]. A yield of ethyl oleate about 60%
was obtained, after 6 h of reaction, at 353 K, when it used 0.005 mol of
­octadec-9-enoic acid and 0.05 mol of ethyl alcohol. After four uses, high
catalytic activity was observed [18]. These reaction was studied over acti-
vated carbon prepared from bamboo with sulfanilic acid [19]. The opti-
mized conditions were as follows: m activated carbon = 12%; octadec-9-enoic
acid: ethyl alcohol = 7, T = 85°C; t = 3 h. In these conditions, a catalytic
efficiency of 91% was obtained [19].
The reaction between octadec-9-enoic acid and methyl alcohol was done
on activated carbons [20]. These catalysts are mesoporous materials, and
sulfonic acid groups were introduced in its surface. The performance of
these materials was greater than Amberlyst-15 [20]. The same reaction was
studied using ordered mesoporous carbons (OMCs). The catalysts showed
a methyl oleate yield of 96.25%. After nine reuses, the catalyst is active [21].
The esterification of fatty acid was performed using sulfonated meso-
porous carbon catalysts [22]. After eight cycles, the FFA conversion was
>91% [22].

14.4 Metal Oxides


The reaction between n-octanoic acid and methyl alcohol was carried over
metal oxides (tin and zirconia) sulfated. The sulfated tin material revealed
high performance for the esterification, which can be explained due to its
strong acidity [23].
The reaction between of octadecanoic acid and methyl alcohol was per-
formed over aerogel sulfated zirconia (A-SZr) and xerogel sulfated zirco-
nia (X-SZr) [24]. A-SZr catalyst has greater textural properties, like total
porous volume than X-SZr material. After 7 h of reaction, the A-SZr cat-
alyst showed 88% yield to product. The activity of aerogel material was
better than xerogel. A-SZr catalyst was reused [24].
The reaction of octanoic acid and ethyl alcohol was performed using
zirconia catalysts. The SZ500 material exhibited high performace for this
reaction [25].
The reaction between octadec-9-enoic acid and methy alcohol was done
using titanium oxide [26]. It was observed that the catalyst [TiO0 /SO42− ]
showed the greatest performed of all material used in the esterification
Reactions of Carboxylic Acids 381

(after 3 h, about 82.2% oleic acid conversion) [26]. This reaction was also
studied over WOx immobilized on MCM-41 [27]. It was prepared materi-
als with several WO3 amount. It was observed a conversion near 100% with
catalysts loading of WO3 higher than 10 wt%. The catalyst with 15 wt%
WO3 can be reused [27]. This reaction was studied over sulphated Zr-KIT-6
(x) material [28]. Under optimized conditions (T = 393 K, methy alcohol:
octadec-9-enoic acid = 20:1, mcatalyst = 4 wt%, t = 6 h), it was obtained
an ester yields of 95%. The catalyst was reused. After three cycles, it was
observed that the material has a good activity and stability [28].
The reaction between octadec-9-enoic acid and n-butyl alcohol has
been studied using zirconium sulphate immoblized on silica (ZS) [29]. It
was observed that the performance of ZS was greater than ZS not immo-
bilized. The catalytic activity increased with amount of ZS. However,
when the ZS amount is high, the catalytic activity decreased. The octadec-­
9-enoic acid conversion increased with the amount of solid. Molar ratio of
­octadec-9-enoic acid:n-butyl alcohol was also optimized. The conversion
of octadec-9-enoic acid enhanced with the molar ratio. Silica-supported
ZS was reused [29].
The reaction between octadec-9-enoic acid and ethyl alcohol was done
over oxides of tin sulfated [30]. At T = 353 K with molar ratio 1:10, the
octadec-9-enoic acid conversion improved with the acidity. The mate-
rial can be recycled [30]. This reaction was also studied over mesoporous
SnO2/WO3 [31]. The reaction was permormed using three temperatures:
313, 333, and 353 K. The octadec-9-enoic acid conversion increases with
the reaction temperature. Also, the catalyst loading was studied. It was
observed that increases with the amount of catalyst. The catalyst can be
reused, after regeneration [31]. The reaction between octadec-9-enoic acid
and ethyl alcohol was done over WO3 supported on a USY zeolite [32].
Catalysts with different amounts of WO3 (from 2.5% to 20.5% WO3) were
prepared. It was observed that the WO3/USY catalysts were more active
than USY. The material perormance enhanced with quantity of WO3 sup-
ported on USY, until the top (11.4% of WO3). After this activity maximum,
when the amount of WO3 increases the catalytic activity decreases. The
material could be recycled [32].
The reactions between octadec-9-enoic acid, octadecanoic acid and
dodecanoic acid and methyl alcohol, and ethyl alcohol and butyl alcohol
was carried out over niobium materials [33]. The reaction between dodeca-
noic acid and butyl alcohol was studied over niobic acid and niobium
phosphate, being phosphate material that showed the best performance.
The dodecanoic acid conversion improved with the temperature. The
382 Biodiesel Technology and Applications

effect of carbon chain legth was also studied. The conversion reduce from
dodecanoic acid (97%) to octadecanoic acid (94%). The octadec-9-enoic
acid conversion was smallest of all carboxylic acids. The niobium phos-
phate catalyst was reused without loss its activity [33].
The reaction between octanoic acid and ethyl alcohol was carried out
over SO24− /Fe x Al1− x PO4 acid catalysts [34]. The presence of Fe on this mate-
rials improved the conversion of octanoic acid. This materials can be recy-
cled [34].

14.5 Heteropolyacids
These materials (HPAs) are polyoxometalates, which are catalysts of differ-
ent reactions. The HPAs have got a low surface area. The heteropolyacids
have been supported on different materials like silica, zeolites, zirconia,
and polymers [35, 36].
The reaction between octadec-9-enoic acid and ethyl alcohol was carried
out in the presence of H3PW12O40 immobilized in ZrO2 [37]. Catalysts with
different amount of heteropolyacids were prepared. The octadec-9-enoic
acid conversion enhanced with the amount of H3PW12O40 immobilized on
zirconia. Optimum reaction conditions were obtained with mcatalyst = 20
wt%, T = 100°C, t = 240 min, and an octadec-9-enoic acid:ethyl alcohol=
1:6. In these conditions, it is obtained a 88% of octadec-9-enoic acid con-
version. This material can be recycled [37].
The reaction between acid hexadecanoic (0.01 mol) and methanol (2.5
mmol) was performed using insoluble HPA salts with Cs [38]. The opti-
mum activity was obtained over a material with Cs equal to 2.3. Material
Cs2.3H0.7PW12O40 was reused [38]. This reaction was studied using het-
eropolyacids in SiO2 [39]. The materials activity reduced in the following
order: HPMo-SiO2 < HSiW-SiO2 < HPW-SiO2. Different catalysts with
HPW immobilized in SiO2 were made. The catalyst PW-silica2 reveals
the greatest reation rate. The material was reused [39]. This reaction was
studied over differents amount of HPW (5% to 30% wt) immobilized on
niobia [40]. The conversion of acid hexadecanoic increases with HPW
quantity supported on Nb2O5. High conversion was obtained with mate-
rial (HPW)25%/Nb2O5. When amount of HPW increases, the conversion
do not change to much. The conversion of acid hexadecanoic increased
when the temperature increased as well. Material (HPW) 25%/Nb2O5 was
recycled [40, 44]. This reaction was studied over different heteropolyacids
(HPW, HPMo, and HSiW) immobilized on SBA-15 [41]. Material HPW1
Reactions of Carboxylic Acids 383

showed highest conversion of all prepared materials. It was also made


materials with various HPW quantity. The acid hexadecanoic conversion
improved when the quantity of heteropolyacid supported on silica until
a maximum, which was obtaind using HPW3-SBA-15 catalyst (6.7% wt).
The effect of catalyst loading (HPW3-SBA-15) on esterification was also
studied. The acid hexadecanoic conversion improved when the catalyst
quantity improved. The HPW3-SBA-15 catalyst could be reused [41]. The
reaction between hexadecanoic acid, octadecanoic acid, octadec-9-enoic
acid, and ethyl acohol was done in the presence of HPW immobilized in
SBA-15 (12.5% wt) [42]. The highest conversion was obtained with hexa-
decanoic acid. This material could be recycled [42].
The reaction between hexadecanoic acid and methyl alcohol was done
over HPW on MCM-41 [43]. Materials with various quantities of HPW
were made. The hexadecanoic acid conversion enhanced with quantity of
HPW anchored to MCM-41. However, for catalysts with high amount of
HPW (for 30% and 40% loading), the palmitic acid conversion did not
increase significatively. The material with 30% loading of HPW (HPW3/
MCM-41) was used for optimize the reaction conditions. Under opti-
mized conditions (T = 333 K; mcatalyst = 0.1 g; t = 360 min; hexadecanoic
acid:methyl alcohol = 1:40), a 90% of hexadecanoic acid conversion
was achieved. Catalyst can be reused [43]. This reaction was also done
over 12-tungstophosphoric acid on SnO2 [44]. Material with 15 wt% of
12-tungstophosphoric acid supported on SnO2 revealed the greatest con-
version. Material was reused. After five reaction cycles, the hexadecanoic
acid conversion stayed high level [44].
The reaction between octadec-9-enoic acid and methyl alcohol was
done over HPW anchored to SBA-15, as catalyst [45]. Under optimized
condition (octadec-9-enoic acid: methyl alcohol = 1:40; mcatalyst = 0.1 g;
T = 313 K; t = 240 min), the conversion of octadec-9-enoic acid was 90%.
Materials could be recycled [45].
The reaction between hexanoic acid and methyl alcohol was done on
HPA supported in Nb2O5, ZrO2, and TiO2 [46]. These materials demon-
strated higher performance than Amberlyst-15, H-Beta, and HY zeolite.
Under optimized condition (hexanoic acid:methyl alcohol = 1:20, mcatalyst =
0.1 g, T = 333 K), a 53.2% conversion was obtained [46].
The reaction between dodecanoic acid and ethyl alcohol was done over
H3PW12O40 (HPW) immobilized on Ta2O5 support [47]. The TOF of HPW
(1.62 min−1) is lower than the TOF of HPW/Ta2O5 (5.58 min−1), which is an
indication that the HPW immobilized on Ta2O5 shows more activity than
the HPW. The catalysts showed high stability [47].
384 Biodiesel Technology and Applications

14.6 Other Materials


The reaction between dodecanoic acid and ethyl alcohol was done over clay
mineral (montmorillonite STx1-b) as catalyst [48]. Under optimized con-
dition (dodecanoic acid: ethyl alcohol = 1:6; mcatalyst = 10 wt%; T = 453 K;
t = 1.5 h), the conversion of octadec-9-enoic acid was 90%. Catalyst STx1-b
can reused. After four cycles, a good catalytic activity was observed [48].
This reaction was also done over Zr-MOF [49]. These materials showed
performance similar to other solids [49].
The reaction between dodecanoic acid, octadecanoic acid, octadec-
9-enoic acid, and ethyl alcohol was preformed using montmorillonite K10
as catalyst [50]. Under optimized conditions (carboxylic acid:ethyl alcool =
1:12, mcatalyst = 12%, T = 453 K, t = 240 min), high conversions (about 90%)
were achieved [50].
The reaction between octanoic acid, dodecanoic acid, octadecanoic
acid, octadec-9-enoic acid, and methyl alcohol was carried out over
Brazilian smectite natural clay [51]. The acid treatment of material gave a
catalyst with high catalytic activity. Under optimized condition (carbox-
ylic acid: methyl alcohol = 1:3; mcatalyst = 10 wt%; T = 373 K; t = 240 min),
the conversion of octanoic acid, dodecanoic acid, octadecanoic acid, and
octadec-9-enoic acid was 99%, 98%, 93%, and 80%, respectively. The mate-
rial could be recycled [51].

14.7 Conclusions
The reaction between a carboxylic acid (like octanoic acid, dodecanoic
acid, octadecanoic acid, and octadec-9-enoic acid) and an alcohol is an
important step in biodiesel production. The existence of high amount of
free carboxylic acids in raw material (like waste cooking oil and animal fat)
can complicate the process of transesterification of triglycerides, which is
done with KOH or NaOH. An alternative to produce biodiesel from oils
or fats with high level of carboxylic acids, it is a esterification of carboxylic
acids before transesterification reaction.
Reactions between carboxylic acids and alcohols are carried out over
homogenous catalysts, which have some environment problems. In order
to become the process as a “environmentally friendly process”, sulfuric acid
can be substituted by solid materials. Some materials like zeolites, hetero-
polyacids, materials with sulfonic groups (MCM-41, SBA-15, activated
carbons, organic polymers), and modified inorganic mixed oxides have
been applied on this type of reaction.
Reactions of Carboxylic Acids 385

References
1. Gerpen; J.V., Biodiesel processing and production. Fuel Process Technol., 86,
1097, 2005.
2. Avhad, M.R., Marchettin, J.M., A review on recent advancement in catalytic
materials for biodiesel production. Renew. Sust. Energ. Rev., 50, 696, 2015.
3. Kulkarni, M.G., Dalai, A.K., Waste Cooking Oils-An Economical Source for
Biodiesel: A Review, Ind. Eng. Chem. Res., 45, 2901, 2006.
4. Lam, M.K., Lee, K.T., Mohamed, A.R., Homogeneous, heterogeneous and
enzymatic catalysis for transesterification of high free fatty acid oil (waste
cooking oil) to biodiesel: A review. Biotechnol. Adv., 28, 500, 2010.
5. Karmakar, B., Halder, G., Progress and future of biodiesel synthesis:
Advancements in oil extraction and conversion technologies, Energy
Convers. Manag., 182, 307, 2019.
6. Serrano, D.P., Melero, J.A., Morales, G., Iglesias, J., Pizarro, P., Progress in the
design of zeolite catalysts for biomass conversion into biofuels and bio-based
chemicals, Catal. Rev., 60, 1, 2018.
7. Venuto, P.B., Organic catalysis over zeolites: A perspective on reaction paths
within micropores, Microporous Mater., 2, 297, 1994.
8. Burton, A., Recent trends in the synthesis of high-silica zeolites, Catal. Rev.,
60, 132, 2018.
9. Kiss, A.A., Dimian, A.C., Rothenberg, G., Solid Acid Catalysts for Biodiesel
Production-Towards Sustainable. Energy. Adv. Synth. Catal., 348, 75, 2006.
10. Prinsen, P., Luque, R., González-Arellano, C., Zeolite catalyzed palmitic acid
esterification. Microporous Mesoporous Mat., 262, 133, 2018.
11. Grossi, C. V., Jardim, E.O., de Araujo, M.H., Lago, R.M., Silva, M.J., Sulfonated
polystyrene: A catalyst with acid and superabsorbent properties for the ester-
ification of fatty acids. Fuel, 89, 257, 2010.
12. Son, S., Kimura, M. H., Kusakabe, K., Esterification of oleic acid in a three-
phase, fixed-bed reactor packed with a cation exchange resin catalyst.
Bioresour. Technol., 102, 2130, 2011.
13. Caetano, C.S., Guerreiro, L., Fonseca, I.M., Ramos, A.M., Vital, J.,
Castanheiro, J.E. Esterification of fatty acids to biodiesel over polymers with
sulfonic acid groups. Appl. Catal. A: Gen., 359, 41, 2009.
14. Caetano, C.S., Caiado, M., Farinha, J., Fonseca, I.M., Ramos, A.M., Vital, J,
Castanheiro, J.E., Esterification of free fatty acids over chitosan with sulfonic
acid groups, Chem. Eng. J., 230, 567, 2013.
15. Morales, G., van Grieken, R., Martín, A., Martínez, F., Sulfonated poly-
styrene-modified mesoporous organosilicas for acid-catalyzed processes.
Chem. Eng. J., 161, 388, 2010.
16. Zong, Z., Duan, W., Lou, T.J., Smith, H. Wu, Preparation of a sugar catalyst
and its use for highly efficient production of biodiesel, Green Chem., 9, 434,
2007.
386 Biodiesel Technology and Applications

17. Takagaki, A., Toda, M., Okamura, M., Kondo, J.N., Hayashi, S., Domen, K.,
Hara, M., Esterification of higher fatty acids by a novel strong solid acid.
Catal. Today, 116, 157, 2006.
18. Peng, L., Philippaerts, A., Ke, X., van Noyen, J., de Clippel, F., van Tendeloo,
G., Jacobs, P.A., Sels, B.F., Preparation of sulfonated ordered mesoporous
carbon and its use for the esterification of fatty acids. Catal. Today, 150, 140,
2010.
19. Geng, L., Wang, Y., Yu, G., Zhu, Y., Efficient carbon-based solid acid catalysts
for the esterification of oleic acid. Catal. Commun., 13, 26, 2011.
20. Niu, S., Ning, Y., Lu, C., Han, K., Yu H., Zhou, Y., Esterification of oleic acid
to produce biodiesel catalyzed by sulfonated activated carbon from bamboo.
Energy Convers. Manag., 163, 59, 2018.
21. Gao, Z., Tanga, S., Cuia, X., Tian, S., Zhang, M., Efficient mesoporous car-
bon-based solid catalyst for the esterification of oleic acid. Fuel, 140, 669,
2015.
22. Zhang, M., Sun, A., Meng, Y., Wang, L., Jiang, H., Li, G. High activity ordered
mesoporous carbon-based solid acid catalyst for the esterification of free
fatty acids. Microporous Mesoporous Mat., 204, 210, 2015.
23. Furuta, S., Matsuhashi, H., Arata, K., Biodiesel fuel production with solid
superacid catalysis in fixed bed reactor under atmospheric pressure. Catal.
Commun., 5, 721, 2004.
24. López, D.E., Goodwin Jr., J.G., Bruce, D.A., Furuta, S. Esterification and
transesterification using modified-zirconia catalysts. Appl. Catal. A: Gen.,
339, 76, 2008.
25. Saravanan, K., Tyagi, B., Bajaj, H.C., Nano-crystalline, mesoporous aerogel
sulfated zirconia as an efficient catalyst for esterification of stearic acid with
methanol. Appl Catal B: Env., 192, 161, 2016.
26. Jimenez-Morales, I., Santamaria-Gonzalez, J., Maireles-Torres, P., Jimenez-
Lopez, A., Zirconium doped MCM-41 supported WO3 solid acid catalysts
for the esterification of oleic acid with methanol. Appl. Catal. A: Gen., 379,
61, 2010.
27. Juan, J.C., Zhang, J., Yarmo, M.A., Structure and reactivity of silica-supported
zirconium sulfate for esterification of fatty acid under solvent-free condition.
Appl. Catal. A: Gen., 332, 209, 2007.
28. Gopinath, S., Kumar, P.S.M., Arafath, K.A.Y., Thiruvengadaravi, K.V.,
Sivanesan S., Baskaralingam, P., Efficient mesoporous SO24− /Zr-KIT-6 solid
acid catalyst for green diesel production from esterification of oleic acid.
Fuel, 203, 488, 2017.
29. Ropero-Vega, J. L., Aldana-Pérez, A., Gómez, R., Niño-Gómez, M.E., Sulfated
2−
titania [TiO2 /SO4 ]: A very active solid acid catalyst for the esterification of
free fatty acids with ethanol. Appl. Catal. A: Gen., 379, 24, 2010.
30. Moreno, J.I., Jaimes, R., Gómez, R., Niño-Gómez, M.E., Evaluation of sul-
fated tin oxides in the esterification reaction of free fatty acids. Catal. Today,
172, 34, 2011.
Reactions of Carboxylic Acids 387

31. Sarkara, A., Ghosh, S.K., Pramanik, P., Investigation of the catalytic effi-
ciency of a new mesoporous catalyst SnO2/WO3 towards oleic acid esterifica-
tion. J. Mol. Catal. A: Chem., 327, 73, 2010.
32. Costa, A.A., Braga, P.R.S., Macedo, J.L., Dias, J.A., Dias, S.C.L., Structural
effects of WO3 incorporation on USY zeolite and application to free fatty
acids esterification. Microporous Mesoporous Mat., 147, 142, 2012.
33. Bassan, I.A.L., Nascimento, D.R., Gil, R.A.S.S., Silva, M.I.P., Moreira, C.R.,
Gonzalez, W.A., Faro Jr., A.C., Onfroye, T., Lachter, E.R., Esterification of
fatty acids with alcohols over niobium phosphate. Fuel Process. Technol., 106,
619, 2013.
34. Liu, B., Jiang, P., Zhang, P., Bian, G., Li, M., Preparation and characteriza-
tion of SO24− /FexAl1− x PO4 solid acid catalysts for caprylic acid esterification.
Catal. Commun., 99, 49, 2017.
35. Sanchez, L.M., Thomas, H.J., Climent, M.J., Romanelli, G.P., Iborra, S.,
Heteropolycompounds as catalysts for biomass product transformations.
Catal. Rev. 58, 497, 2016.
36. Patel, A., Narkhede, N., Singh, S., Pathan, S. Keggin-type lacunary and tran-
sition metal substituted polyoxometalates as heterogeneous catalysts: A
recent progress. Catal. Rev. 58, 337, 2016.
37. Oliveira, C.F., Dezaneti, L.M., Garcia, F.A.C., de Macedo, J. L., Dias, J. A.,
Dias, S.C.L., Alvim, K.S.P., Esterification of oleic acid with ethanol by
12-tungstophosphoric acid supported on zirconia. Appl Catal. A:Gen., 372,
153, 2010.
38. Narasimharao, K., Brown, D. R., Lee, A.F., Newman, A.D., Siril, P.F., Tavener,
S.J., Wilson, K., Structure–activity relations in Cs-doped heteropolyacid cat-
alysts for biodiesel production. J. Catal., 248, 226–234, 2007.
39. Caetano, C.S., Fonseca, I.M., Ramos, A.M., Vital, J., Castanheiro, J.E.
Esterification of free fatty acids with methanol using heteropolyacids immo-
bilized on silica. Catal. Commun. 9, 1996, 2008.
40. Srilatha, K., Lingaiah, N., Devi, B.L.A.P., Prasad, R. B. N., Venkateswar, S.,
Prasad, P.S.S., Esterification of free fatty acids for biodiesel production over
heteropoly tungstate supported on niobia catalysts. Appl. Catal A: Gen., 365,
28, 2009.
41. Tropecêlo, A.I., Casimiro, M.H., Fonseca, I.M., Ramos, A.M., Vital, J.,
Castanheiro, J. E., Esterification of free fatty acids to biodiesel over hetero-
polyacids immobilized on mesoporous silica. Appl. Catal A:Gen., 390, 183,
2010.
42. Castanheiro, J.E., Fonseca, I.M., Ramos, A.M., Vital, J., Tungstophosphoric
acid immobilised in SBA-15 as an efficient heterogeneous acid catalyst for
the conversion of terpenes and free fatty acids. Microporous Mesoporous
Mat., 249, 16, 2017.
43. Brahmkhatri, V.; Patel, A., Biodiesel Production by Esterification of Free
Fatty Acids over 12-Tungstophosphoric Acid Anchored to MCM-41. Ind.
Eng. Chem. Res., 50, 6620, 2011.
388 Biodiesel Technology and Applications

44. Srilatha, K., Kumar, C.R., Devi, B.L.A.P., Prasad, R.B.N., Prasada, P.S.S.,
Lingaiah, N., Efficient solid acid catalysts for esterification of free fatty acids
with methanol for the production of biodiesel. Catal. Sci. Technol., 1, 662,
2011.
45. Brahmkhatri, V.; Patel, A., 12-Tungstophosphoric acid anchored to SBA-15:
An efficient, environmentally benign reusable catalysts for biodiesel produc-
tion by esterification of free fatty acids. Appl Catal A: Gen., 403, 161, 2011.
46. Alsalme, A., Kozhevnikova, E.F., Kozhevnikov, I.V., Heteropoly acids as cat-
alysts for liquid-phase esterification and transesterification. Appl. Catal. A:
Gen., 349, 170, 2008.
47. Xu, L., Yang, X., Yu, X., Maynurkader, G.Y., Preparation of mesoporous poly-
oxometalate–tantalum pentoxide composite catalyst for efficient esterifica-
tion of fatty acid. Catal. Commun. 9, 1607, 2008.
48. Santos, P.R.S., Wypych, F., Voll, F.A.P., Hamerski, F., Corazza, M.L., Kinetics
of ethylic esterification of lauric acid on acid activated montmorillonite
(STx1-b) as catalyst. Fuel, 181, 600, 2016.
49. Cirujano, F.G.; Corma, A.; Llabrés, F.X.; Xamena, I., Zirconium-containing
metal organic frameworks as solid acid catalysts for the esterification of free
fatty acids: Synthesis of biodiesel and other compounds of interest. Catal.
Today, 257, 213, 2015.
50. Kanda, L.R.S., Corazza, M.L., Zatta, L., Wypychc, F., Kinetics evaluation of
the ethyl esterification of long chain fatty acids using commercial montmo-
rillonite K10 as catalyst. Fuel, 193, 265, 2017.
51. Rezende, M.J.C., Pinto, A.C., Esterification of fatty acids using acid-activated
Brazilian smectite natural clay as a catalyst. Renew Energ., 92, 171, 2016.
15
Biodiesel Production From Non-Edible
and Waste Lipid Sources
Opeoluwa O. Fasanya1, Aishat A. Osigbesan1 and Onoriode P. Avbenake2,3*
1
Petrochemical and Allieds Department, National Research Institute for Chemical
Technology, Zaria, Nigeria
2
Chemical and Petroleum Engineering Department, Bayero University,
Kano, Nigeria
3
School of Science & Technology, Pan-Atlantic University, Ibeju-Lekki, Nigeria

Abstract
Biodiesel has become a very popular renewable/alternative fuel. The popularity of
biodiesel is driven by the relative ease of synthesis and the wide array of feedstock
from which it can be produced. These include vegetable oils, animal fats, and other
sources of biomass. Vegetable oils are the main feed stock that have been used for
biodiesel synthesis. To the extent that oils meant as food for humans and animals
are being diverted for biodiesel production. To stop this unsustainable practice,
nonedible vegetable feedstocks are constantly identified, researched, and culti-
vated. Much progress has been made over the last few years in this especially in
terms of oils from seeds of plants such as Jatropha curcas, Calophyllum inophyllum,
Azadirachta indica, Hevea brasiliensis, and Ricinus communis, among others. In
addition, waste vegetable oils are also being harnessed for the purpose of biodiesel
production, rather than being discarded once they degrade. Furthermore, the
use of microalgae, a third-generation source of biodiesel, was equally discussed,
owing to their popularity among researchers. The increased interest is a conse-
quence of the speed and ease of cultivation and to a larger extent the quantity of
oil that could be harvested. As a matter of fact, fourth-generation biodiesel is now
being derived from genetically modified algae strains which contain even more
oils. Accordingly, this chapter aims to critically review a portion of the numerous
developments concerning biodiesel from non-edible sources.

*Corresponding author: paulavbenake@gmail.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (389–428) © 2021 Scrivener Publishing LLC

389
390 Biodiesel Technology and Applications

Keywords: Biodiesel, lipids, renewable energy, waste utilization, methyl esters,


fuel, non-edible oils, deacidification

15.1 Introduction
The ever increasing energy requirements in the world today have placed a
large demand on limited quantity of fossil based fuels in the world at large.
The increase in energy demand is due to technological advancement, indus-
trialization, and population growth all over the world. The International
Energy Agency (IEA) estimates that 50% more energy will be required
in 2030 when compared with present energy consumption [1]. There is
no gain saying that copious levels of environmental pollution particularly
crude oil spills and CO2 emissions are associated with extraction and use
of fossil fuels. Coupled with fluctuating crude oil prices, there is pertinent
need of generating alternative sustainable fuels for energy consumption.
Some alternate sources of fuel that have generated a lot of interest include
solar cells [2], fuel cells powered by hydrogen [3], ethanol [4], methanol [5],
and diesel derived from biological materials or biomass. Biomass as a source
of fuel is quite advantageous because it is renewable and also has relatively
low carbon foot print during synthesis of fuels. This guarantees some mea-
sure of environmental stability while energy requirements are being met.
Globally, the United States of America, Brazil, and European Union are
the main consumers of bio-based fuels. As at 2000, annual biodiesel pro-
duction was recorded to be about 0.8 million tons [6]. In 2007, Germany
alone marketed 3.2 million tons of biodiesel [7]. Despite this production
rate, the IEA claims that global biofuel production rate is not able to keep
up with the present demand for the product. As at 2018, renewable energy
met only 3.7% of transportation fuel demand [1]. The IEA have, however,
predicated that between 2019 and 2024, a 25% production increase corre-
sponding to approximately 190 billion litres would be achieved.
Diesel produced from biomass is popularly referred to as biodiesel. It
is a fuel which is made up of monoalkyl esters of long-chain fatty acids
derived from biomass which are sourced from plants or animal fats. Better
combustion efficiency has been observed in biodiesel synthesized by this
route [8]. The International Association for Testing and Materials (ASTM)
has set out criteria in the ASTM D6751 (Table 15.1) which lists out prop-
erties that a fuel must possess to be called biodiesel. Inherent advantages
of biodiesel use include its blending capacity with other energy sources,
lubricity which could invariably help extend engine lifespan, higher flash
point, and lack of toxicity [9, 10].
Biodiesel Production From Non-Edible Sources 391

Table 15.1 Biodiesel specifications and their limits.


Properties EN 14214 ASTM D6751
Composition of biodiesel C12-C22 C12-C22
Esters content Greater than 96.5% (m/m) –
Density at 15°C 860–900 kg m−3 –
Viscosity at 40°C 3.5–5.0 mm2 s−1 1.9–6.0 mm2 s−1
Flash point ≥101°C ≥ 130°C
Sulphur content ≤10 mg kg−1 50 mg kg−1
Carbon residue ≤0.3% (m/m) ≤ 0.05 (m/m)
Cetane number ≥51 ≥ 47
Sulfated ash ≤0.02% (m/m) ≤ 0.02% (m/m)
Water content ≤500 mg kg−1 ≤ 0.05% (v/v)
Corrosion – 3h
Oxidative stability 110°C 4h 3h
Acid value ≤0.50 mg KOH g−1 ≤ 0.50 mg KOH g−1
Iodine value 130 g I2/100 g –
Methanol content ≤0.02% (m/m) –
Monoacylglycerols ≤0.8% (m/m) –
content
Diacylglycerols content ≤0.2 (m/m) –
Triacylglycerols content ≤0.2 (m/m) –
Free glycerine ≤0.02% (m/m) ≤ 0.20% (m/m)
Total glycerrin ≤0.25% (m/m) ≤ 0.25% (m/m)
Pour point – −15 to −16°C
Phosphorous amount ≤4 mg kg−1 ≤ 0.001% (m/m)
Cloud point – −3 to 12°C
Adapted from Fonseca et al. (2019) [11].
392 Biodiesel Technology and Applications

Generally, biodiesel is produced by transesterification of triglycerides


with a short chain alcohol in the presence of a catalyst to form fatty acid
alkyl esters and glycerol as products. The reaction scheme is shown in
Figure 15.1 and in Equations 15.1 to 15.3. Transesterification is a relatively
simple and cost-effective process [12]. Though some challenges with sepa-
ration of the final products from catalyst may occur depending on whether
homogenous or heterogeneous catalysts are employed. Transesterification
catalysts could either be acidic [13], basic [14, 15], or enzymatic [16].
While homogeneous catalysts such as NaOH and KOH are generally more
effective, separation after reaction requires thorough washing with large
amounts of water to ensure complete removal of the catalyst. This chal-
lenge was resolved with use of solid catalysts such as CaO. Though these
are not quite as reactive and are often prone to deactivation or single digit
limited cycles of use. Strides have made to develop stable heterogeneous
catalysts which are resistant to rapid deactivation.

Triglycerides + Methanol = diglycerides + FAME (15.1)

Diglycerides + Methanol = Monoglycerides + FAME (15.2)

Monoglycerides + Methanol = Glycerol + FAME (15.3)

For oils with high free fatty acid (FFA), an initial pre-treatment step is
required where esterification of the oil with methanol in the presence of an
acid catalyst (usually sulphuric acid) is carried out as seen in Equation 15.4

FFA + Methanol = FAME + Water (15.4)

One major point of concern in biodiesel synthesis is the biomass feed-


stock. Globally, oil-bearing plants have been widely explored for biodiesel
production. Typically, plants with seeds that yield high amounts of oil are
usually preferred as starting triglyceride sources, since these would result in
large volumes of biodiesel being produced. Economically, this is the route

H2C COO R1 CH3 H2C OH


Catalyst COO R1
HC COO R2 + 3CH3OH CH3 COO R2 + HC OH
CH3 COO R3 H2C OH
H 2C COO R3

Triglyceride Methanol Fatty acid methyl ester Glycerol

Figure 15.1 Transesterification reaction showing reaction of methanol and triglyceride to


form Fatty acid methyl ester (FAME) and glycerol.
Biodiesel Production From Non-Edible Sources 393

to follow. The challenge, however, is that edible oils which are in large
demand as food sources such as palm oil [17, 18] and soybean oil [19, 20]
are being used to produce the fuel. This has led to unhealthy competition
between use of oil as food or fuel. Inadvertently, resulting in global deple-
tion of edible oils for food and consequently affecting the price of food
and fuel dependent on these oils. It is expected, therefore, that this would
birth major economic crisis in future [12]. Available arable land for agri-
cultural cultivation would reduce and lose nutrients due to large volumes
of plants constantly being cultivated to meet the ever increasing energy
and food requirements. Edible biodiesel feedstocks are generally termed
first-generation feedstock. They were initially explored for large-scale pro-
duction. Currently, it is estimated that over 90% of current biodiesel pro-
duction is from edible vegetable oils [21]. Palm oil, sunflower oil, coconut,
groundnut, and soybean [19] oils all fall in this category [12].
To forestall the impending crisis associated with the use of edible
crops, attention has been focused on utilization of non-edible oils. These
are generally classified as second-generation biodiesel feedstock. Some
of which include but not limited to Jatropha curcas, Hevca brasillenis,
and Azadirachta indica. Biodiesel from waste oil is also considered to be
second-generation feedstock. Using these types of feedstock offers the
advantage of reducing pollution concerns and costs associated with safe
disposal. It is estimated that production costs can be reduced by as much
as 70% when non-edible oils are utilized [22].
There are numerous reviews on the use of non-edible oils for biodiesel
production [6, 23–27]. Atadashi discussed the effect of using high FFA
oils, many of which are not fit for human consumption [28]. The review
by Balat discussed about trends in non-edible feedstock that were gain-
ing traction as of 2011 [29]. Discussions were oils derived from Jatropha
curcas, Pongamia pinnata, Madhucca indica, and microalgae among others
were reviewed. Although there are now more recent reviews highlighting
developments using micro-algae [30], some of which we will discuss here.
The review by Sajjadi and co-workers contained prediction models for
parameters such as viscosity, density, flash point, cetane number, and heat-
ing value of both edible and non-edible based biodiesels [31]. Pereira and
co-workers also reviewed models but their focus was on thermos-physical
models to estimate melting points of fats and cold flow properties of bio-
diesel [32]. While Kirubakaran and Selvan focused their review on using
waste chicken fat [33]. Some other reviews are location specific, highlight-
ing potential feedstock and production strategies [34, 35].
In the end, the choice of non-edible feedstock for biodiesel produc-
tion will be highly location dependent. Such as in the case of Jatropha,
394 Biodiesel Technology and Applications

Callophyllum innophyllum, and neem which are native to India. Another


important property that would determine choice of feed for biodiesel pro-
duction is its ability be grown rapidly and under a wide variety of con-
ditions. This will definitely reduce competition on the use of arable land
for cultivating crops for energy for plant based sources. For animal-based
sources feeds with low FFA or with FFA that can be reduced with ease are
feedstock of choice. This chapter highlights both plant and animal-based
sources feedstock that are non-edible and shows progresses made in utiliz-
ing them for biodiesel production.

15.2 Non-Edible Plant-Based Oils


15.2.1 Jatropha curcas
Jatropha curcas is a drought resistant crop that is capable of growing in fal-
lowed land. It has the ability to withstand a variety climatic zones and soil
conditions. In addition to these, its short gestation period and fast growth
have led to its widespread cultivation. It is a highly suitable crop for bio-
diesel production in developing countries especially in the tropical and
subtropical parts of the world [36]. Commercial growth of the crop has,
however, experienced setbacks all over the world due to poor understand-
ing of the crop [37]. Recently, detailed studies have, however, shown that
optimum growing conditions such as an excess of 700-mm rainfall annu-
ally, soil pH ranging between 6 and 9, and annual temperature between
16°C and 28°C are required for commercial Jatropha plantation [38]. A
lot of attention is given to its cultivation as it is responsible for 36%–44%
of the biodiesel selling price depending on where it is planted [37, 38].
Deliberate J. curcas plantations are currently being grown in Rwanda [39],
China, Ethiopia, India, Nepal, and Tanzania [38].
Its combustion and emission properties closely resemble that of petro-
leum diesel. Properties such as flash point and ignition point are higher
for J. curcas derived fuels [36]. As far back as World War II, J. curcas oil
was used was blended with diesel and used as a fuel alternative [28]. A
number of authors have maintained that it is the best non-edible oil for
biodiesel production [36, 39, 40]. The oil content of the seed appears to be
dependent on location and probably variety as values seen in literature fall
between the range of 25% to 59% [28, 36]. Oil from J. curcas can also be
used for producing soaps and candles.
The use of Jatropha for biodiesel has been widely investigated. Its high
FFA usually entails that a two-step production process is often required.
Table 15.2 Different catalysts, reaction conditions, and yield during synthesis of biodiesel from Jatropha seed oil between 2015
and 2019.
Catalyst Methanol
Yield Time Pressure Temperature loading oil Catalyst
Catalyst (%) (hours) (MPa) (°C) (wt %) ratio Pre-treatment type Ref.
CaSO4/ 94 4 Atm 120 12 9 None Amphoteric [42]
Fe2O2-SiO2
Sr+2 - CaO/ 99.6 2 Atm 65 5 9 Not stated Basic [60]
MgO
Zn8FeC 100 4 4.5 MPa 160 7 40 None Amphoteric [41]
Na2SiO3@Ni/ 97.7 2 Atm 65 7 9 Esterification Basic- [61]
JRC magnetic
KF supported 92.2 1 Atm 75 7 21 Not stated Alkaline [62]
on red
mud
nano CaO 98.54 2.1 Atm 60 2 5.15 Esterification Basic [63]
SO2−4/TiO2 85.3 24 0.2 140 4 9 Acidic [64]
(Continued)
Biodiesel Production From Non-Edible Sources
395
396

Table 15.2 Different catalysts, reaction conditions, and yield during synthesis of biodiesel from Jatropha seed oil between 2015
and 2019. (Continued)
Catalyst Methanol
Yield Time Pressure Temperature loading oil Catalyst
Catalyst (%) (hours) (MPa) (°C) (wt %) ratio Pre-treatment type Ref.
Al-SBA-15 99 24 4 180 6.5 12 None Acidic [43]
Fe-Zn-1 – 4 Atm 160 4 10 Acidic [65]
Na2ZrO3 99 1 Atm 65 30 3 [20]
Cs-Na2ZrO3 90 1 Atm 65 15 3 Basic [20]
Ca-La-Al 96.91 3 Atm 150 15 7 [44]
Biodiesel Technology and Applications

Si-MMT-Ph- 98 2.5 Atm 110 6 5 [66]


SO3H
Calcined 96.1 3 Atm 70 6 9 Basic [67]
animal
bones
Biodiesel Production From Non-Edible Sources 397

The discovery and use of amphoteric catalysts has been suggested as a way
of combining the esterification and transesterification step into one. The
use of Zn-FeC catalysts resulted in as much as 100% biodiesel yield with
extremely high stability [41]. Other studies using amphoteric catalysts
have given yield of 94% after 4 hours of reaction [42]. Table 15.2 contains
data on recent catalyst developments for biodiesel synthesis from Jatropha.
Interestingly, one report has shown that under the right conditions some
transformation can occur without the use of catalysts [43]. Jatropha still
remains one of most widely researched oils for biodiesel production. As
can be seen from Table 15.2, there is still a lot of room for research espe-
cially in terms of catalyst development. Reusability of catalysts for multiple
runs remains a challenge. Though some work has shown reusability of up
to five times with just a 5%–10% drop in FAME yield [44].

15.2.2 Calophyllum inophyllum


Calophyllum inophyllum is a perennial plant that grows predominantly in
Australia, Sri Lanka, India, and Southeast Asia. It is a mangrove plant with
oil containing seeds [45]. The relatively high oil content in its kernels has
also made the plant attractive for biodiesel production in parts of the world
where its growth is prevalent. About 46%–60% of toxic non-edible oil can
be extracted from its kernels [45, 46]. Apart from biodiesel synthesis, the
oils have been found to have properties which can be used to treat skin
diseases, arthritis, and burns, to salve wounds, and many others [47].
Biodiesel from C. Inophyllum generally has high oxidation stability but
it is improved further with the use of additives such as Butylated hydroxy-
tolune, 4-methyl-6-tert-butyphenol and antioxidant extracted from pon-
gamia leaf [48]. As with other oils previously discussed, the biodiesel yield
is dependent on reaction conditions and catalyst type with yields of over
85% reported by different groups [49, 50]. Various reports have also been
written describing its performance in diesel compression engines [51–54].

15.2.3 Mesua ferrea


Mesua ferrea L. also known as Nahar or Ceylon ironwood [55] is a non-­
edible evergreen timber plant native to Sri Lanka and rather abundant in
Northeast India [56, 57]. It also grows freely in Nepal, Indonesia, Vietnam,
Cambodia, and Singapore [58]. Its seeds containing between 55% and 57%
of oil [56]. The timber plant is its unsaturated fatty acid content is between
65.85% and 68.31% [56]. With regard to biodiesel production, this plant
is relatively underutilized but this trend is changing. The reason for this is
398 Biodiesel Technology and Applications

attributed to some shortfalls experienced in Jatropha yield over the last few
years [58]. When tested on compression injection engines, results obtained
were comparable with conventional diesel [58, 59]. Suggesting that the bio-
diesel from this oil is suitable for blending or as a complement to Jatropha
derived biodiesel.

15.2.4 Jojoba Oil


Simmondsia chinensis (linn) Schneider (commonly identified as jojoba but
is also called deer nut, oat nut, wild hazel, and coffee berry) is a prom-
ising oil seed crop for the economic development of the arid and semi-
arid land all over the globe. The jojoba plant is a monogenetic dioecious
gray-green shrub belonging to Simmondsiaceae family. It is native to the
North American deserts, especially those of south western states in the
United States (California, Arizona, and Utah) and Northwestern Mexico
(Baja California and Sonora). Jojoba oil became widely known through
the Spanish missionaries of the 18th and 19th centuries. Native Americans
used the crushed seed oil for skin care and medicinal purposes. The
Spanish missionaries became aware of its uses and introduced it to other
parts of the world. The plant has been cultivated for more than 30 years
in many countries worldwide, such as India, Mexico, Chile, Argentina,
Australia, Tunisia, the Palestinian territories, Saudi Arabia, and Egypt, due
to its promising economic value [3], with the United States considered the
largest jojoba oil-producing country, followed by Mexico. The jojoba oil
has been reported previously as having potential capabilities for the cos-
metics and skincare industry [68].
The oil extracted from Jojoba seeds is golden colored and odorless with
relatively high viscosity which is much higher than petroleum fuels [69,
70]. It has wax-like unsaturated esters, consisting of a straight chain of fatty
acids and higher alcohols. It has been stated severally that wax and not
oils are extracted from Jojoba due to their peculiar nature [71, 72]. The
main constituent of these FFA is gondoic acid (C20:1) with a concentration
of 59.5%. Other components as reported by Shah and co-workers include
oleic acid, erucic acid, and arachidonic acid with concentrations of 10.7%,
12.3%, and 9.1%, respectively [68].

15.2.5 Azadirachta indica


Oil can be extracted from seed kernels of the Azadirachta indica (Neem
tree). Depending on the source, neem seeds could have as low as 30% oil
[73] or between 40% and 45% oil [74]. It has been found to have diverse
Biodiesel Production From Non-Edible Sources 399

applications. One of the main uses of neem seed oil (NSO) has been
widely used as a pesticide due to the presence of Azadirachtin [75, 76].
Due to the presence of this compound and other similar ones, neem oil
is toxic to humans. Children are especially more vulnerable to neem oil
poisoning [77].
NSO has been reported to have high levels of FFA. Like Jatropha and
other high FFA oils, this invariably requires a two-step pre-treatment
during biodiesel synthesis [78]. This involves the use of an acid catalyst
for esterification with methanol to reduce the FFA. This is then followed
by transesterification with a base or acid catalyst to produce the required
methyl ester. H2SO4 has been predominantly used to reduce FFA but solid
catalysts are also being explored. Solid acid catalysts have the advantage
of easy separation after reaction though activity at times may not be
comparable to homogenous catalysts [13]. Recently, solid catalysts with
improved activity are being reported. The use phosphoric mordenite was
reported to result in a 92% reduction on neem oil FFA after just 60 min-
utes of reaction [73].
The performance of biodiesel from neem oil has been evaluated either
as a blend or as a standalone fuel. It was reported that methyl esters from
NSO blended with diesel gave better performance provided the concentra-
tion was below 20% [79]. Higher blends resulted in increased NOx emis-
sions though engine performance was marginally better. Besides, catalytic
converters are currently being employed to reduce emissions from neem
oil derived diesel [80].

15.2.6 Rubber Seed Oil


Rubber tree (Hevea brasiliensis) is one of the abundant natural resources
found in humid tropical countries. Seed yield from rubber plantations
varies from 100 to 150 kg/ha, depending on soil fertility and crop den-
sity. Extensive plantations can be found in Nigeria [81], Malaysia, India
[82], Indonesia, Liberia [83], Thailand [84], and China [85]. As at 2009,
plantations in Malaysia produced more than 1.2 million hectares pro-
ducing roughly 1.2 million tons of seeds per year [86]. It is primarily
grown for latex production as a source of foreign exchange, though its
usefulness is not maximally explored. The tree produces oil-bearing
seeds with oil content far more than that obtainable from the likes of
Jatropha and Karanja seeds. Yields of kernel from rubber seeds range
from 57% to 63% [87]. This leads to yield about 5,500 tons of rubber seed
oil and about 6,000 tons of defatted protein meal for animal and human
consumption [81].
400
Table 15.3 Comparison of FFA composition of selected non-edible oils.
Palmitic Stearic Oleic Linoleic Linolenic Myristic
acid acid acid acid acid Malvalic Palmitoleic Lauric acid
(C16:0) (C18:0) (18:1) (C18:2) (C18:3) acid Ricinoleic Arachidic (C16:1) (C12:0) (C14:0) Ref.

Rubber seed 9.1 5.6 24 46.2 14.2 – 0.9 [84]


oil

Castor oil 1.3 1.2 3.6 4.6 0.4 – 88.9 [21,


91]

Jatropha Oil 13.37 6.1 41.71 36.42 – – – 0.82 [92,


93]

mango seed 15.6 30.2 48.2 6.9 – – – [94]


oil

Neem seed oil 18.1 18.1 44.5 18.3 0.2 – – 0.8 [95]

Yellow 26.4 3.86 39.4 27.03 – – 1.1 [96]


oleander
Biodiesel Technology and Applications

Spent coffee 32.4 7.5 10.1 43.1 0.9 – 3.1 [97]


seed
grounds

Euonymus 14.5 3.1 49.8 29.3 – – 0.07 2 [98]


maackii

Ceiba 22.37 3.8 23.24 33.63 9.14 [99]


pentandra

(Continued)
Table 15.3 Comparison of FFA composition of selected non-edible oils. (Continued)
Palmitic Stearic Oleic Linoleic Linolenic Myristic
acid acid acid acid acid Malvalic Palmitoleic Lauric acid
(C16:0) (C18:0) (18:1) (C18:2) (C18:3) acid Ricinoleic Arachidic (C16:1) (C12:0) (C14:0) Ref.

Attalea 8.5 16.1 1.4 44.0 15.4 [100]


speciose
oil

Mesua Ferrea 13.63 13.65 50.71 19.47 0.316 1.3 0.087 [101]

Prunus [30]
armeniaca 3.31 1.625 71.76 20.19 1.03

Cerbera [102]
odollam 24.86 5.79 52.82 13.65 0.75

Karanja

Tobacco
Biodiesel Production From Non-Edible Sources
401
402 Biodiesel Technology and Applications

The extracted oil from rubber seed is non-edible and rich in unsatu-
rated FFA(roughly 80%) [88], and this is a property essential to making it
a prominent raw material for biodiesel production [89]. This, in turn, gives
more economic value to growth of rubber plantations. It becomes impera-
tive to obtain specific data for sample of rubber seed oil from a particular
area because there is a range of variation in the physicochemical parame-
ters of the oil due to environmental factors such as rain-fall, soil fertility,
agronomic practices, maturation period, and genetic substitution [89, 90].
For biodiesel production, the foremost characterization to consider is
the FFA composition and value of the oil. A freshly extracted rubber seed
oil (RSO) has a total acid value of about 47% [90]; this implies FFA of
23.5% comprising of the saturated and unsaturated acids. The fatty acid
composition of RSO along with other oils is captured in Table 15.3.
Saturated fatty acids are responsible for higher cloud point, cetane num-
ber, and general stability of the oil and diesel synthesized from it [103]. As
with other non-edible oils the high FFA of RSO requires that it under goes
acid esterification before transesterification to FAME. This is a problem
that is encountered even when utilizing continuous flow setup invariably
making a system as proposed by Sai and coworkers a semi batch process
[104]. Some authors have proposed the use of co-solvents during transes-
terification as a way of getting higher ester yield at with relatively milder
transesterification conditions. One such report indicates that using aceto-
nitrile as co-solvent, 99% ester yield was obtained at 40°C after 30 minutes
of reaction [105].

15.2.7 Ricinus communis as Feedstock (Castor Oil)


The castor plant, which has numerous varieties, bears seeds that contain
between 40% and 55% oil. Castor oil has a relatively low cost but it is of
immense value as it can be utilized in a number of different applications.
It is utilized in production of pharmaceuticals [106], paints [107], poly-
mers [91], and for fuel production [108]. As of 2015, the estimated average
castor oil seed worldwide was put down as 1.1 tons ha−1 [21]. This value
is expected to grow exponentially. It is said that the plant is able to thrive
under harsh conditions. It is believed to originate from either Africa or
Asia.
Castor oil is inedible and is composed of mostly ricinoleic acid, between
80% and 90% [108] depending on location and variety of the seeds from
which the oil is extracted. It has a yellow-green hue and the biodiesel
from castor oil is viscous and is usually blended with fossil fuel derived
diesel. Castor oil is rather peculiar in the sense that it possesses high
Biodiesel Production From Non-Edible Sources 403

OH O
18 16 14 10 9 7 5 3
12 1 O
17 15 13 11 8 6 4 2

OH O

OH O

Figure 15.2 Chemical structure of riconelic acid, major component of castor oil [106].

concentration of riconelic acid which has an OH group as seen in the


structure in Figure 15.2.
As a result of riconoleic acid, castor oil is highly viscous and the bio-
diesel produced from it also has high viscosity. Nevertheless, blending with
fossil derived diesels and others with very low viscosity usually results in
overall improved characteristics [109]. Biodiesel from castor oil has been
heavily researched over last two decades [108–114].
Recent focus has however been geared towards process optimization,
intensification and the development of more efficient catalysts. One such
development is the one-step transesterification process which has been
proposed with varying degrees of success. The use of sulphonated phenyl
silane montmorillonite catalyst was reported to have 89% conversion of
castor oil to FAME after 300-minute reaction time [66].
Moradi and Ghanadi [113] evaluated the feasibility of producing bio-
diesel directly from castor seed in a one-step route. Their model developed
using centre composite design predicted a 91% FAME yield.

15.2.8 Other Non-Edible Oils


Oils from plants such as Thevetia peruviana have also been explored as a
potential biodiesel feedstock. The plant is highly toxic in nature [115] and is
commonly known as yellow oleander or milk bush. However, the seeds can
also be used to treat cardiac problems, skin issues, and asthma [96]. The oil
content of the seeds ranges from 60% to 65% and can grow in a variety of
arid areas [116] where it is usually used for decorations. Interestingly, a yield
of 96 wt% biodiesel has been reported by Deka and Basumatary [116] after
404 Biodiesel Technology and Applications

reaction for 3 hours at room temperature. The duo carbonized the trunk
of Musa balbisana plant and used it as catalyst for transesterification with
methanol. Other groups who have used the same plant report 90 wt% yield
after 75 minutes at 70°C using 2.8 (w/v)% of CaO as catalyst [96].
The efficacy of spent coffee grounds (SCG) as potential biodiesel feed-
stock has also been investigated. The drawback with this feedstock is the
high moisture content which the spent grounds possess as a result of brew-
ing. In addition to this, they also are characterized by high FFA content
ranging between 21 and 38 w/w % [97]. Some other authors have proposed
in situ transesterification of the seeds as a way of improving process eco-
nomics of production from SCG [117].
Also, Silybum marianum L., a wild plant native to Iraq, Iran, Syria, and
some parts of China on the Asian continent, produces seeds which have oil
content of between 23% and 45% depending on the variety. Esterification
of high FFA oil from S. Marianum using a sulfonated acid carbon catalyst
resulted in low FFA oil [118]. They utilized 6% w/w of the carbon acid
catalyst, 68°C reaction temperature, and 15:1 methanol-to-oil ratio with a
reaction time of 180 minutes resulted in FFA reduction of over 90% from
an initial value of 20.0 mg KOH/g. Transesterification with KOH at 60°C
gave ester yield of 96.69% after 75 minutes of reaction. Table 15.4 contains
physico-chemical properties of different oils.
There are so many other plants which have not been discussed in this
list. In different regions of the world, scientists are investigating different
oils to see which contain the appropriate esters and which have high oil
concentration [119]. These factors greatly affect the economics of the bio-
diesel synthesis process and are key to ensuring reduced dependence on
fossil derived diesels.

15.3 Waste Animal Fats


Animal fats (tallow) have also been considered as potential biodiesel
sources. There are challenges when dealing with this feedstock which
result in necessary pre-treatment steps before conversion to biodiesel. The
major challenges include (1) high moisture content, (2) high FFA, and (3)
fats that are solid at room temperature. Nevertheless, research is exploring
ways to harness the tallow as an economically viable feedstock. Different
animal fats have been investigated on. Some of which include duck [123],
beef [124, 125], sheep [126], and chicken fats [127].
Similar to other high FFA feedstock, esterification with methanol in the
presence of an acid such as H2SO4 is a method that has been explored but
Biodiesel Production From Non-Edible Sources 405

not as extensively as with vegetable oils. However, water content should be


less than 0.1 wt% before the reaction is carried out as water hampers the
reaction. Laboratory scale studies have been successfully carried out using
microwaves to shorten the esterification time [128]. The cost of utilizing
microwaves at an industrial scale may lead to increased price of biodiesel
due to larger energy requirements.

15.4 Expired and Waste Cooking Oils


Vegetable oils after being used for extensive frying are classified as inedi-
ble. The oils have been said to degrade, due to harmful compounds which
are produced during frying. Initially, the oils were sold as feed to the ani-
mal industry which has now been banned by the European commission
in 2002 [129]. This is to forestall consumption of these chemicals which
could contaminate the meat of animals which consumed. Generally, waste
cooking oils (WCO) are generated in bulk by households, hotels, and
the food industry. To illustrate this, in 2015, 69% of the 77,000 tons of
WCO generated in Portugal came from hotels and food catering services,
while domestic use accounted for 25% [130]. The use of expired and waste
vegetable oils as feedstock for biodiesel is timely as these oils are usually
discarded into sewage systems which invariably become a problem for
wastewater treatment plants [131]. To this end, many developed countries
have established laws prohibiting and penalizing disposal of WCO during
sewage systems [132]. Consequently, utilizing WCOs in biodiesel produc-
tion reduces pollution and gives them added value. The potential for waste
cooking oil (WCO) biodiesel is much higher in European and American
countries where copious amounts of WCOs are generated [133].
The source of WCO may not particularly have an effect on the final
biodiesel produced. Studies have shown that properties such as acid value
and density may have appreciable effect on the biodiesel synthesis process
[134]. Therefore, to obtain higher acid value oils a two-stage reaction pro-
cess is required: first is esterification and then followed by transesterifica-
tion. Some authors have also suggested that FAME yields with WCOs are
marginally lower due to some of these compounds formed as oils degrade
while cooking/frying [11].
Purification of WCO before use in biodiesel synthesis presents a chal-
lenge in some situations. Food debris and water often pose problems as
contaminants, which could have adverse effects in the synthesis process.
Filtration or gravity settling may be employed to remove debris that may
accompany the WCO.
406 Biodiesel Technology and Applications

Also, the level of saturation of the WCO used is apposite. Although deg-
radation is certain, it is not possible to predict the exact level of degrada-
tion that will take place. Generally, degradation of the oil while cooking
results in the oils becoming more saturated due to FFA. As a result, FFA
content of most WCO typically ranges between 2% and 10% [135] which
has a negative effect on the viscosity of the final biodiesel [135]. For bench
scale synthesis, this can be overlooked. It, however, is a major challenge in
plant design, as a flexible system which can cater for oils with a wide range
of FFA would be required.

15.5 Algae/Microalgae
Harnessing algae for biofuels production is gaining widespread attention.
Initially, algae were used for just lipid and protein production. But, it was
since discovered that a good portion of the lipids were suitable for pro-
duction of biodiesel. Fuel applications exploiting algal components include
transesterification of lipids to biodiesel, saccharification of carbohydrates
to ethanol, gasification of biomass to syngas, cracking of hydrocarbons to
gasoline, and biosynthesis of hydrogen gas [136]. This is due to the expand-
ing use of fossil fuels, increasing CO2 emission into the atmosphere, and
consequently contributing to climate change. With depleting fossil fuel
reservoirs coupled with global clampdown on greenhouse gases spur the
motivation for the development of renewable energy sources such as algae/
microalgae [136].
Algae require carbon dioxide, sun light, and water for growth in non-ar-
able land, arid land, and in waste water. They grow very fast in a short
period, producing as much as double their mass in few hours. The lipid
content of algae is higher than that of seed plants, producing not less than
30 times more oil per acre than seed plants [137]. Algae can be subdivided
into microalgae and macroalgae; either of them has different metabolic
behaviors during their growth. These include (1) autotrophic behavior
where algae make use of light solely as a source of energy where the light
energy is converted to chemical energy using CO2 via photosynthesis;
(2) mixotrophic behavior allows algae to perform photosynthesis using
both organic compound and CO2 in the presence of light energy for growth;
(3) heterotrophic behavior where algae utilize only organic compounds as
energy and carbon source; and (4) photoheterotrophic behavior makes
algae use light as well as organic compounds as carbon source [136, 138].
They can change the metabolic pathway according to the changes in
the environmental conditions. The ability of algae to fix CO2 is a method
Biodiesel Production From Non-Edible Sources 407

of removing CO2 from flue gases of power plants, thereby reducing emis-
sions of greenhouse gases [9]. Photoautotrophic means of algae cultivation
is recommended for when growing algae for biodiesel production. This
ensures cost reduction and CO2 sequestration [139]. Figure 15.3 shows
the pathways of biodiesel production from microalgae and by products of
the production process. Large scale growth of algae is conducted either in
open ponds or photobioreactors [140]. The photobioreactors reportedly
have higher algae yield as cultivation conditions can be controlled to max-
imize production.
It has been established that oil yields from microalgae can be as much
as 31 to 72 times that obtainable from Jatropha [30]. The bottleneck in
biodiesel production from algae is to identify the strains with high oil pro-
ductivity and to develop cost-effective growing and harvesting systems
[141]. There is a disagreement on exact quantity of algae strains that have
been identified in the world today. There are possibly between 44,000 and
100,000 different algae strains depending on the source of the information
[140, 142].

Harvest & Drying

Lipid extraction

CO2 + Nutrients

Biodiesel

Animal feeding

Glycerol
Cosmetics

Figure 15.3 Microalgal biodiesel refinery producing multiple products from algal
biomass [136].
Table 15.4 Some selected algae strains and their production data.
408

Total Lipid
Biomass Biomass Lipid extracted
production yield production (wt% of
Strain Habitat Nutrients (g/L/day) (g/L) (g/L/day) biomass) Ref.
Chlorella vulgaris Fresh Water Thin Stillage 2.5 9.8 1.1 43 [145]
Chlorella vulgaris Fresh Water Soy Whey 1.6 6.3 0.2 11 [145]
Chlorella vulgaris Fresh Water Modified 2 8 0.6 27 [145]
basal
Medium
Chlorella pyrenoidosa Fresh Water Bold’s basal 0.106 na 0.019 29.68 [146]
medium
Chlorella vulgaris Fresh Water Bold’s basal 1.65 na 0.44 26 [147]
medium
Biodiesel Technology and Applications

S. obliquus (YSL02) Fresh Water Bold’s basal 1.84 na 0.53 29 [147]


medium
Spirulina sp Fresh Water Zarrouk’s 1.37 na 0.66 20 [148]
medium
Aurantiochytrium sp. Marine Defined 6.69 31.8 Na 38.1 [149]
(KRS101) water medium
Chlamydomonasreinhardtii Fresh water na 2 na 0.505 25.25 [150]
Table 15.5 Some properties of biodiesel synthesizes from microalgae.
Density Iodine Kinematic Oxidation
Acid value Cold filter 15°C value viscosity stability
(mg plugging (kg/ (g I2/100 40°C 110°C Cloud Pour Cetane
KOH/g) (°C) m3) g) (mm2/s) (h) Saponification point point number Ref.

From microalgae 0.29 –13 0.88 4.43 4.52 NA [151]


(Chlorella
protothecoides)

From microalgae NA –4.8 0.88 4.2 95.7 55 [152]


(Nannochloropsis
oculata)

From microalgae na 7.8 0.89 3.74 NA 47.3 [152]


(Phaeodactylum
tricornutum)

From microalgae na –4.6 0.91 4.63 5.6 32.9 [152]


(Scenedesmus
dimorphus)

From microalgae NA –0.99 NA 54.57 [153]


(Chlorella
protothecoides)

(Continued)
Biodiesel Production From Non-Edible Sources
409
410

Table 15.5 Some properties of biodiesel synthesizes from microalgae. (Continued)


Density Iodine Kinematic Oxidation
Acid value Cold filter 15°C value viscosity stability
(mg plugging (kg/ (g I2/100 40°C 110°C Cloud Pour Cetane
KOH/g) (°C) m3) g) (mm2/s) (h) Saponification point point number Ref.

From microalgae 3.55 54.24 [153]


(Chlorella
emersonii)

From microalgae 2.58 49.93 [153]


(Chlorella salina)

From microalgae 4.6 44 [153]


(Chlorella
vulgaris)

From microalgae NA 0.9 88.5 4.5 7.4 233.71 9.2 3.1 50 [154]
Biodiesel Technology and Applications

(Chlorella
vulgaris)

spirulina 0.83 2.9 44.55 [155]

Chlorella sorokiniana –9.66 0.875 48.411 4.9 13.58 59.68 [156]

Micractinium sp. 1.69 0.873 28.33 5.07 17.19 61.48 [156]


Biodiesel Production From Non-Edible Sources 411

The oil content of each algae strain varies with Colonial, Capsoid,
Coccoid, Palmelloid, Filamentous, and Parenchymatou as some of the
various forms of algae. Chlorella zofingiensis, Chlorella protothecoids, and
Schizochytrium limacinum have been identified as strains with large oil
production capacity [143]. Table 15.4 lists some strains with their growth
per day and lipid content, and extensive lists can be found in literature
[144]. Table 15.5 shows the properties of biodiesel synthesized from differ-
ent algal strains.

15.6 Insects as Biodiesel Feedstock


The use of insects as biodiesel feedstock is still a relatively virgin area of
focus. They had hitherto been mainly considered as a source of protein
for animal feed [157]. They have been identified as viable sources of fats
for biodiesel production but there is a lot that is yet to be researched upon
using this feedstock. Insects offer the advantage low cost of rearing, large
insect yield, and relatively small space is required for cultivation [158].
Of which, methods of improving insect yields are constantly being devel-
oped [159]. As with other feedstock, there are criteria to be considered
when selecting an insect as feed for biodiesel production. These include
fat content usually of larva, speed of completion of insect life cycle, space
requirements, and feeding costs. Detailed discussions on this can be stud-
ied in the review by Manzano-Agugliaro et al. [160]. At their larva stage,
metabolic reserves for non-feeding periods are developed. These reserves
are majorly composed of high quantities of fat, in addition to glycogen and
proteins [160]. Insects which have shown high lipid content include Musca
domestica, Hermetia illucens, and Rhynchophorus sp. with lipid contents
of approximately 21%, 29%, and 43%, respectively [161]. Table 15.6 shows
acid content of biodiesel produced from some larva, while Table 15.7 pres-
ents properties of some biodiesel synthesized from insects.
The use of insect larva is also seen as a means of tackling environmental
pollution. The larva can feed on residual food grease, animal waste, sew-
age sludge, and other such contaminants [162], effectively reducing cost of
purchasing feed. This was demonstrated in the work of Zheng et al. [163]
who were able to increase biodiesel yield and reduce food grease consid-
erably by using it as feed for the larva. With regard to process economics,
important factors to consider asides from oil content and cultivation of the
larva include extraction temperature, solvent types, and quantity of solvent
required. The use of microwaves has been studied and recommended as a
means of increasing the rate of lipid extraction [158].
412

Table 15.6 Fatty oil composition of diesel derived from insect larva.
Palmitic acid Stearic acid Oleic Linoleic acid Linolenic acid Myristic Palmitoleic Lauric Capric
(C16:0) (C18:0) acid (C18:2) (C18:3) (C14:1) (C16:1) (C12:0) (C10:0)
Hermetia illucens 18.2 5.1 27.1 7.5 3.7 9.4 23.4 1.8 [163]

Boettcherisca 18.9 0.2 44.5 4.1 7.6 1.4 18.3 1.6 [164]
peregrine

Zophobas morio 32.74 9.36 29.43 22.53 0.85 2.16 [165]

Musca domestica 28.4 2.2 24.1 18.1 1.1 3.3 20.1 [166]
L. (grown
on poultry
manure)

Musca domestica 28.2 3 22.9 18.8 1.4 2.91 20.6 [166]


Biodiesel Technology and Applications

L. (grown on
kitchen waste)

Musca domestic 27.14 5.01 24.54 13.45 0.36 5.96 17.4 1.21 [167]
(food waste
and wheat
bran)
Table 15.7 Fuel properties of biodiesels synthesized from insect larva.
MeOH- Water
to- Density content
oil Acid @ Flash (mg/ Cloud Cetane Kinematic Calorific Oxidative
Catalyst Temperature ratio % Cata value 15°C point kg) point number visco value stability Ref.
Hermetia NaOH 62 10 1.1 0.875 143 387 5.6 49 4.6 [168]
illucens
Boettcherisca 0.62 0.884 130 <.03% 5.6 [164]
peregrine
Musca 0.875 166 <0.03% 56 4.19 39.56 3.2 [166]
domestica
L. (poultry
manure)
Zophobas KOH 50 5 1.25 0.15 0.871 159 <0.03% 57.3 4.22 [165]
morio
Biodiesel Production From Non-Edible Sources
413
414 Biodiesel Technology and Applications

15.7 Deacidification
Majority of the oils discussed above are characterized by high FFA. High
FFA leads to saponification which results in difficulty in obtaining pure
methyl esters. To combat this, copious amounts of water usually required
to wash followed by drying to in order to make it suitable for use. In some
instances, soap emulsions form making washing even more difficult [169].
This is especially the case when alkali catalysts are used. To this end, a
number of methods have been explored to reduce the FFA.
Esterification as mentioned earlier entails the reaction of high FFA
oil with methanol in the presence of a concentrated acid, usually H2SO4.
Neutralization of high FFA oils by an alkali is presently the commercial
means of reducing FFA [170]. The soap formed is washed and subsequently
the oil is dried before transesterification.
Steam deacidification is well utilized in is refining oils for consump-
tion following degumming and bleaching. Operating conditions for
deacidification are largely dependent on the type of oil used. Major
variables that dictate the extent of FFA removal are the temperature
and quantity steam employed in the process. For Jatropha, 3.0 wt % of
steam in relation to oil used at 225°C was reported to achieve 97.33%
reduction in FFA content [169]. Other methods include transesterification
using glycerol [171], microorganisms or enzymes, solvent extraction [172],
or extraction using supercritical fluids. The final choice is inadvertently
decided by the process economics.

15.8 Other Technologies


Asides transesterification, other technologies have been developed for the
production of diesel and other fuel fractions from waste. Most popular
technologies include pyrolysis [173] and thermal cracking [174–176].
Pyrolysis is advantageous in that a wide variety of materials can be used
as feedstock. Vegetable oils [177], animal fat [173], or even plant husks can
be used.
The pyrolytic process, however, requires temperatures as high as 400°C.
Despite the higher energy requirements, pyrolysis offers the advantage of
handling feedstock with high FFA. The complexity of product washing due
to saponification is removed via this process. It, however, would require the
use of catalysts which are selective to diesel fractions. Apart from hydro-
carbons, typical pyrolysis products include solid char and gases such CO
and CO2.
Biodiesel Production From Non-Edible Sources 415

15.9 Conclusion
In conclusion, the use of non-edible oils as biodiesel feedstock is undoubt-
edly the preferred pathway for future energy generation. It will reduce
pressure on edible feedstock such as palm oil, soybean oil, and rapeseed oil
which have been widely used for large-scale biodiesel production.
On the one hand, transesterification is still the more preferred route,
although optimization of other methods is still being developed. Processes
will employ either heterogeneous catalysts with high methyl ester yield or
one-step processes with better economics of production. Algae has been
found to be rather promising non-edible alternative, and it appears that
it just may be the preferred feedstock of choice for a very long while to
come. Though technological and scientific research are already heralding
fourth-generation sources.

References
1. IEA, Renewables 2019. IEA: Paris. 2019.
2. Ye, Y., Jo, C., Jeong, I., et al., Functional mesoporous materials for energy
applications: solar cells, fuel cells, and batteries. Nanoscale. 5(11). 4584.2013.
3. Fasanya, O.O., Al-Hajri, R., Ahmed, O.U., et al., Copper zinc oxide nanocata-
lysts grown on cordierite substrate for hydrogen production using methanol
steam reforming. International Journal of Hydrogen Energy. 2019.
4. Manochio, C., Andrade, B.R., Rodriguez, R.P., et al., Ethanol from biomass:
A comparative overview. Renewable and Sustainable Energy Reviews. 80.
743.2017.
5. Larson, E.D. and Katofsky, R.E., Production of Hydrogen and Methanol via
Biomass Gasification, in Advances in Thermochemical Biomass Conversion,
A.V. Bridgwater, Editor. 1993, Springer Netherlands: Dordrecht. p. 495.
6. Mardhiah, H.H., Ong, H.C., Masjuki, H.H., et al., A review on latest devel-
opments and future prospects of heterogeneous catalyst in biodiesel produc-
tion from non-edible oils. Renewable and Sustainable Energy Reviews. 67.
1225.2017.
7. Bockey, D., The significance and perspective of biodiesel production – A
European and global view☆. Journal of Oilseeds and fats Crops and Lipids.
26. 40.2019.
8. Tang, Z.-E., Lim, S., Pang, Y.-L., et al., Synthesis of biomass as heterogeneous
catalyst for application in biodiesel production: State of the art and funda-
mental review. Renewable and Sustainable Energy Reviews. 92. 235.2018.
9. Azeem, M.W., Hanif, M.A., Al-Sabahi, J.N., et al., Production of biodiesel
from low priced, renewable and abundant date seed oil. Renewable Energy.
86. 124.2016.
416 Biodiesel Technology and Applications

10. Devarajan, Y., Munuswamy, D.B., and Mahalingam, A., Influence of nano-ad-
ditive on performance and emission characteristics of a diesel engine running
on neat neem oil biodiesel. Environmental Science and Pollution Research.
25(26). 26167.2018.
11. Fonseca, J.M., Teleken, J.G., de Cinque Almeida, V., et al., Biodiesel from
waste frying oils: Methods of production and purification. Energy Conversion
and Management. 184. 205.2019.
12. Mishra, V.K. and Goswami, R., A review of production, properties and
advantages of biodiesel. Biofuels. 9(2). 273.2018.
13. Sani, Y.M., Daud, W.M.A.W., and Abdul Aziz, A.R., Activity of solid acid
catalysts for biodiesel production: A critical review. Applied Catalysis A:
General. 470. 140.2014.
14. Banković–Ilić, I.B., Miladinović, M.R., Stamenković, O.S., et al., Application
of nano CaO–based catalysts in biodiesel synthesis. Renewable and
Sustainable Energy Reviews. 72. 746.2017.
15. Kouzu, M., Fujimori, A., Suzuki, T., et al., Industrial feasibility of powdery
CaO catalyst for production of biodiesel. Fuel Processing Technology. 165.
94.2017.
16. Tan, T., Lu, J., Nie, K., et al., Biodiesel production with immobilized lipase: A
review. Biotechnology Advances. 28(5). 628.2010.
17. Mekhilef, S., Siga, S., and Saidur, R., A review on palm oil biodiesel as a
source of renewable fuel. Renewable and Sustainable Energy Reviews. 15(4).
1937.2011.
18. Mat Yasin, M.H., Mamat, R., Najafi, G., et al., Potentials of palm oil as new
feedstock oil for a global alternative fuel: A review. Renewable and Sustainable
Energy Reviews. 79. 1034.2017.
19. Kouzu, M., Kasuno, T., Tajika, M., et al., Calcium oxide as a solid base cata-
lyst for transesterification of soybean oil and its application to biodiesel pro-
duction. Fuel. 87(12). 2798.2008.
20. Torres-Rodríguez, D.A., Romero-Ibarra, I.C., Ibarra, I.A., et al., Biodiesel
production from soybean and Jatropha oils using cesium impregnated
sodium zirconate as a heterogeneous base catalyst. Renewable Energy. 93.
323.2016.
21. Sánchez, N., Sánchez, R., Encinar, J.M., et al., Complete analysis of castor oil
methanolysis to obtain biodiesel. Fuel. 147. 95.2015.
22. Samsudeen, N., Dammalapati, S., Mondal, S., et al. Production of Biodiesel
from Neem Oil Feedstock Using Bifunctional Catalyst. in Materials, Energy
and Environment Engineering. 2017. Singapore: Springer Singapore.
23. Rezania, S., Oryani, B., Park, J., et al., Review on transesterification of non-­
edible sources for biodiesel production with a focus on economic aspects, fuel
properties and by-product applications. Energy Conversion and Management.
201.112155.2019.
24. Demirbas, A., Bafail, A., Ahmad, W., et al., Biodiesel production from
non-edible plant oils. 34(2). 290.2016.
Biodiesel Production From Non-Edible Sources 417

25. Borugadda, V.B. and Goud, V.V., Biodiesel production from renewable feed-
stocks: Status and opportunities. Renewable and Sustainable Energy Reviews.
16(7). 4763.2012.
26. Atabani, A.E., Silitonga, A.S., Ong, H.C., et al., Non-edible vegetable oils: A
critical evaluation of oil extraction, fatty acid compositions, biodiesel pro-
duction, characteristics, engine performance and emissions production.
Renewable and Sustainable Energy Reviews. 18. 211.2013.
27. khan, T.M.Y., Atabani, A.E., Badruddin, I.A., et al., Recent scenario and tech-
nologies to utilize non-edible oils for biodiesel production. Renewable and
Sustainable Energy Reviews. 37. 840.2014.
28. Atadashi, I.M., Aroua, M.K., Abdul Aziz, A.R., et al., Production of biodiesel
using high free fatty acid feedstocks. Renewable and Sustainable Energy
Reviews. 16(5). 3275.2012.
29. Balat, M., Potential alternatives to edible oils for biodiesel production –
A review of current work. Energy Conversion and Management. 52(2).
1479.2011.
30. Singh, K., Kaloni, D., Gaur, S., et al., Current research and perspectives on
microalgae-derived biodiesel. Biofuels. 11(1). 1.2017.
31. Sajjadi, B., Raman, A.A.A., and Arandiyan, H., 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. 62.2016.
32. Pereira, E., Meirelles, A.J.A., and Maximo, G.J., Predictive models for phys-
ical properties of fats, oils, and biodiesel fuels. Fluid Phase Equilibria. 508.
112440.2020.
33. Kirubakaran, M. and Arul Mozhi Selvan, V., A comprehensive review of low
cost biodiesel production from waste chicken fat. Renewable and Sustainable
Energy Reviews. 82. 390.2018.
34. Đurišić-Mladenović, N., Kiss, F., Škrbić, B., et al., Current state of the biodiesel
production and the indigenous feedstock potential in Serbia. Renewable and
Sustainable Energy Reviews. 81. 280.2018.
35. Yang, L., Takase, M., Zhang, M., et al., Potential non-edible oil feedstock for
biodiesel production in Africa: A survey. Renewable and Sustainable Energy
Reviews. 38. 461.2014.
36. Thapa, S., Indrawan, N., and Bhoi, P.R., An overview on fuel properties and
prospects of Jatropha biodiesel as fuel for engines. Environmental Technology
& Innovation. 9. 210.2018.
37. Edrisi, S.A., Dubey, R.K., Tripathi, V., et al., Jatropha curcas L.: A crucified
plant waiting for resurgence. Renewable and Sustainable Energy Reviews. 41.
855.2015.
38. Baral, N.R., Neupane, P., Ale, B.B., et al., Stochastic economic and environ-
mental footprints of biodiesel production from Jatropha curcas Linnaeus
in the different federal states of Nepal. Renewable and Sustainable Energy
Reviews. 120. 109619.2020.
418 Biodiesel Technology and Applications

39. Ntaribi, T. and Paul, D.I., Status of Jatropha plants farming for biodiesel pro-
duction in Rwanda. Energy for Sustainable Development. 47. 133.2018.
40. Chukwuezie, O.C., Nwaigwe, K.N., Asoegwu, S.N., et al., Diesel engine per-
formance of jatropha biodiesel: a review. Biofuels. 5(4). 415.2014.
41. Wang, Y.-T., Fang, Z., Yang, X.-X., et al., One-step production of biodiesel
from Jatropha oils with high acid value at low temperature by magnetic acid-
base amphoteric nanoparticles. Chemical Engineering Journal. 348. 929.2018.
42. Teo, S.H., Islam, A., Chan, E.S., et al., Efficient biodiesel production from
Jatropha curcus using CaSO4/Fe2O3-SiO2 core-shell magnetic nanoparti-
cles. Journal of Cleaner Production. 208. 816.2019.
43. Meloni, D., Perra, D., Monaci, R., et al., Transesterification of Jatropha
curcas oil and soybean oil on Al-SBA-15 catalysts. Applied Catalysis B:
Environmental. 184. 163.2016.
44. Syamsuddin, Y., Murat, M.N., and Hameed, B.H., Transesterification of
Jatropha oil with dimethyl carbonate to produce fatty acid methyl ester
over reusable Ca–La–Al mixed-oxide catalyst. Energy Conversion and
Management. 106. 1356.2015.
45. Hapsari, S., Susanto, D.F., Aparamarta, H.W., et al., Separation and
Purification of Wax from Nyamplung (Calophyllum inophyllum) Seed Oil.
Materials Science Forum. 964. 1.2019.
46. Azad, A.K., Rasul, M.G., Khan, M.M.K., et al., Prospects, feedstocks and chal-
lenges of biodiesel production from beauty leaf oil and castor oil: A noned-
ible oil sources in Australia. Renewable and Sustainable Energy Reviews. 61.
302.2016.
47. Susanto, D.F., Aparamarta, H.W., Widjaja, A., et al., Identification of phy-
tochemical compounds in Calophyllum inophyllum leaves. Asian Pacific
Journal of Tropical Biomedicine. 7(9). 773.2017.
48. Mohamed Shameer, P. and Ramesh, K., FTIR evaluation on the fuel stability
of calophyllum inophyllum biodiesel: Influence of tert-butyl hydroquinone
(TBHQ) antioxidant. Journal of Mechanical Science and Technology. 31(7).
3611.2017.
49. Arumugam, A. and Ponnusami, V., Biodiesel production from Calophyllum
inophyllum oil a potential non-edible feedstock: An overview. Renewable
Energy. 131. 459.2019.
50. Naveenkumar, R. and Baskar, G., Biodiesel production from Calophyllum
inophyllum oil using zinc doped calcium oxide (Plaster of Paris) nanocata-
lyst. Bioresource Technology. 280. 493.2019.
51. Ong, H.C., Masjuki, H.H., Mahlia, T.M.I., et al., Optimization of Biodiesel
Production and Engine Performance from High Free Fatty Acid Calophyllum
inophyllum Oil in CI Diesel Engine. Journal Energy Conversion and
Management. 81. 30.2014.
52. Ong, H.C., Masjuki, H.H., Mahlia, T.M.I., et al., Engine performance and
emissions using Jatropha curcas, Ceiba pentandra and Calophyllum inophyl-
lum biodiesel in a CI diesel engine. Energy. 69. 427.2014.
Biodiesel Production From Non-Edible Sources 419

53. How, H.G., Masjuki, H.H., Kalam, M.A., et al., Effect of Calophyllum
Inophyllum biodiesel-diesel blends on combustion, performance, exhaust
particulate matter and gaseous emissions in a multi-cylinder diesel engine.
Fuel. 227. 154.2018.
54. Ashok, B., Nanthagopal, K., and Sakthi Vignesh, D., Calophyllum inophyl-
lum methyl ester biodiesel blend as an alternate fuel for diesel engine appli-
cations. Alexandria Engineering Journal. 57(3). 1239.2018.
55. Jaikumar, S., Bhatti, S.K., Srinivas, V., et al., Combustion, vibration, and noise
characteristics of direct injection VCR diesel engine fuelled with Mesua fer-
rea oil methyl ester blends. International Journal of Ambient Energy. 1.2020.
56. Bora, P., Boro, J., Konwar, L.J., et al., A comparative study of Mesua ferrea
L. based hybrid fuel with diesel fuel and biodiesel. Energy Sources, Part A:
Recovery, Utilization, and Environmental Effects. 38(9). 1279.2016.
57. Bora, A.P., Dhawane, S.H., Anupam, K., et al., Biodiesel synthesis from
Mesua ferrea oil using waste shell derived carbon catalyst. Renewable Energy.
121. 195.2018.
58. Dash, S.K., Lingfa, P., and Chavan, S.B., An experimental investigation on the
application potential of heterogeneous catalyzed Nahar biodiesel and its die-
sel blends as diesel engine fuels. Energy Sources, Part A: Recovery, Utilization,
and Environmental Effects. 40(24). 2923.2018.
59. Singh, N., Kumar, H., Jha, M.K., et al., Complete heat balance, performance,
and emission evaluation of a CI engine fueled with Mesua ferrea methyl
and ethyl ester’s blends with petrodiesel. Journal of Thermal Analysis and
Calorimetry. 122(2). 907.2015.
60. Sudsakorn, K., Saiwuttikul, S., Palitsakun, S., et al., Biodiesel production
from Jatropha Curcas oil using strontium-doped CaO/MgO catalyst. Journal
of Environmental Chemical Engineering. 5(3). 2845.2017.
61. Zhang, F., Tian, X.-F., Fang, Z., et al., Catalytic production of Jatropha bio-
diesel and hydrogen with magnetic carbonaceous acid and base synthesized
from Jatropha hulls. Energy Conversion and Management. 142. 107.2017.
62. Zhang, L.Y., Wang, Y.Z., Wei, G.T., et al., Biodiesel preparation from Jatropha
oil catalyzed by KF/Red mud catalyst. Energy Sources, Part A: Recovery,
Utilization, and Environmental Effects. 38(12). 1713.2016.
63. Anr, R., Saleh, A.A., Islam, M.S., et al., Biodiesel Production from Crude
Jatropha Oil using a Highly Active Heterogeneous Nanocatalyst by
Optimizing Transesterification Reaction Parameters. Energy & Fuels. 30(1).
334.2016.
64. Chen, C., Cai, L., Shangguan, X., et al., Heterogeneous and efficient trans-
esterification of Jatropha curcas L. seed oil to produce biodiesel catalysed by
nano-sized SO42; TiO2. 5(11). 181331.2018.
65. Kumar, G., Singh, V., and Kumar, D., Ultrasonic-assisted continuous metha-
nolysis of Jatropha curcas oil in the appearance of biodiesel used as an inter-
mediate solvent. Ultrasonics Sonochemistry. 39. 384.2017.
420 Biodiesel Technology and Applications

66. Negm, N.A., Sayed, G.H., Yehia, F.Z., et al., Biodiesel production from
one-step heterogeneous catalyzed process of Castor oil and Jatropha oil
using novel sulphonated phenyl silane montmorillonite catalyst. Journal of
Molecular Liquids. 234. 157.2017.
67. Nisar, J., Razaq, R., Farooq, M., et al., Enhanced biodiesel production from
Jatropha oil using calcined waste animal bones as catalyst. Renewable Energy.
101. 111.2017.
68. Shah, S.N., Sharma, B.K., Moser, B.R., et al., Preparation and Evaluation of
Jojoba Oil Methyl Esters as Biodiesel and as a Blend Component in Ultra-
Low Sulfur Diesel Fuel. BioEnergy Research. 3(2). 214.2010.
69. Al-Obaidi, J.R., Halabi, M.F., AlKhalifah, N.S., et al., A review on plant
importance, biotechnological aspects, and cultivation challenges of jojoba
plant. Biological Research. 50(1). 25.2017.
70. Shah, S.N., Sharma, B.K., and Moser, B.R., Preparation of Biofuel Using
Acetylatation of Jojoba Fatty Alcohols and Assessment as a Blend Component
in Ultralow Sulfur Diesel Fuel. Energy & Fuels. 24(5). 3189.2010.
71. Harry-O’kuru, R.E., Biresaw, G., Gordon, S., et al., Physical Characteristics
of Tetrahydroxy and Acylated Derivatives of Jojoba Liquid Wax in Lubricant
Applications. Journal of Analytical Methods in Chemistry. 2018.2018.
72. Abdelmoez, W., Tayeb Aghareed, M., Mustafa, A., et al., Green Approach
for Biodiesel Production from Jojoba Oil Supported by Process Modeling and
Simulation, in International Journal of Chemical Reactor Engineering. p. 185.
2016.
73. SathyaSelvabala, V., Varathachary, T.K., Selvaraj, D.K., et al., Removal of free
fatty acid in Azadirachta indica (Neem) seed oil using phosphoric acid mod-
ified mordenite for biodiesel production. Bioresource Technology. 101(15).
5897.2010.
74. Chhabra, M., Saini, B.S., and Dwivedi, G., Impact assessment of biofuel
from waste neem oil. Energy Sources, Part A: Recovery, Utilization, and
Environmental Effects. 1.2019.
75. Zanuncio, J.C., Mourão, S.A., Martínez, L.C., et al., Toxic effects of the neem
oil (Azadirachta indica) formulation on the stink bug predator, Podisus
nigrispinus (Heteroptera: Pentatomidae). Scientific Reports. 6(1). 30261.2016.
76. Naumann, K. and Isman, M.B., Evaluation of neem Azadirachta indica seed
extracts and oils as oviposition deterrents to noctuid moths.76(2). 115.1995
77. Mishra, A. and Dave, N., Neem oil poisoning: Case report of an adult with
toxic encephalopathy. Indian J Crit Care Med. 17(5). 321.2013.
78. Awolu, O.O. and Layokun, S.K., Optimization of two-step transesterification
production of biodiesel from neem (Azadirachta indica) oil. International
Journal of Energy and Environmental Engineering. 4(1). 39.2013.
79. Sivalakshmi, S. and Balusamy, T., The Performance, Combustion, and
Emission Characteristics of Neem Oil Methyl Ester and Its Diesel Blends
in a Diesel Engine. Energy Sources, Part A: Recovery, Utilization, and
Environmental Effects. 36(2). 142.2014.
Biodiesel Production From Non-Edible Sources 421

80. Viswanathan, K., Experimental investigation on emission reduction in neem


oil biodiesel using selective catalytic reduction and catalytic converter tech-
niques. Environmental Science and Pollution Research. 25(14). 13548.2018.
81. Nwokolo, E., Rubber (Hevea brasiliensis L.) seed, oil and meal, in Food and
Feed from Legumes and Oilseeds, E. Nwokolo and J. Smartt, Editors. 1996,
Springer US: Boston, MA. p. 333.
82. Gandhi, V.M., Cherian, K.M., and Mulky, M.J., Nutritional and toxicological
evaluation of rubber seed oil. Journal of the American Oil Chemists’ Society.
67(11). 883.1990.
83. Olugasa, B.O., Dogba, J.B., Ogunro, B., et al., The rubber plantation environ-
ment and Lassa fever epidemics in Liberia, 2008–2012: A spatial regression.
Spatial and Spatio-temporal Epidemiology. 11. 163.2014.
84. Roschat, W., Siritanon, T., Yoosuk, B., et al., Rubber seed oil as potential
non-edible feedstock for biodiesel production using heterogeneous catalyst
in Thailand. Renewable Energy. 101. 937.2017.
85. Zhu, Y., Xu, J., and Mortimer, P.E., The influence of seed and oil storage on
the acid levels of rubber seed oil, derived from Hevea brasiliensis grown in
Xishuangbanna, China. Energy. 36(8). 5403.2011.
86. Singh, H.K.A.P.G., Yusup, S., and Wai, C.K., Physicochemical Properties of
Crude Rubber Seed Oil for Biogasoline Production. Procedia Engineering.
148. 426.2016.
87. Nwokolo, E. and Akpapunam, M., Content and availability of nutrients in
rubber seed meal. J. Tropical Science.1986.
88. Onoji, S.E., Iyuke, S.E., Igbafe, A.I., et al., Rubber seed oil: A potential renew-
able source of biodiesel for sustainable development in sub-Saharan Africa.
Energy Conversion and Management. 110. 125.2016.
89. Abubakar, B., Study on Extraction and Characterization of Rubber Seeds Oil.
Australian Journal of Basic and Applied Sciences. 8. 7.2014.
90. Aigbodion, A.I. and Bakare, I.O., Rubber seed oil quality assessment and
authentication. Journal of the American Oil Chemists’ Society. 82(7). 465.2005
91. Nekhavhambe, E., Mukaya, H.E., and Nkazi, D.B., Development of castor
oil–based polymers: A review.1(4). e10030.2019.
92. Díaz, L. and Borges, M.E., Low-Quality Vegetable Oils as Feedstock for
Biodiesel Production Using K-Pumice as Solid Catalyst. Tolerance of Water
and Free Fatty Acids Contents. Journal of Agricultural and Food Chemistry.
60(32). 7928.2012.
93. Raia, R.Z., da Silva, L.S., Marcucci, S.M.P., et al., Biodiesel production from
Jatropha curcas L. oil by simultaneous esterification and transesterification
using sulphated zirconia.Catalysis Today. 289. 105.2017.
94. Nde, D.B., Astete, C., and Boldor, D., ESolvent-free, enzyme-catalyzed bio-
diesel production from mango, neem, and shea oils via response surface
methodology. AMB Express. 5(1). 83.2015.
95. Modi, A.J., Gosai, D.C., and Solanki, C.M., Experimental Study of Effect of
EGR Rates on NOx and Smoke Emission of LHR Diesel Engine Fueled with
422 Biodiesel Technology and Applications

Blends of Diesel and Neem Biodiesel. Journal of The Institution of Engineers


(India): Series C. 99(2). 181.2018.
96. Arun, S.B., Suresh, R., and Avinash, E., Optimization of Biodiesel Production
from Yellow Oleander (Thevetia Peruviana) using Response Surface
Methodology. Materials Today: Proceedings. 4(8). 7293.2017.
97. Efthymiopoulos, I., Hellier, P., Ladommatos, N., et al., Effect of Solvent
Extraction Parameters on the Recovery of Oil From Spent Coffee Grounds
for Biofuel Production.Waste and Biomass Valorization. 10(2). 253.2019.
98. Liu, J.-Z., Cui, Q., Kang, Y.-F., et al., Euonymus maackii Rupr. Seed oil as
a new potential non-edible feedstock for biodiesel. Renewable Energy. 133.
261.2019.
99. Anwar, F., Rashid, U., Shahid, S.A., et al., Physicochemical and Antioxidant
Characteristics of Kapok (Ceiba pentandra Gaertn.) Seed Oil.91(6). 1047.2014.
100. Solis, J.L., Alejo, L., and Kiros, Y., Calcium and tin oxides for heterogeneous
transesterification of Babasssu oil (Attalea speciosa). Journal of Environmental
Chemical Engineering. 4(4, Part B). 4870.2016.
101. Aslam, M., Saxena, P., and Sarma, A., Green Technology for Biodiesel Production
from Mesua Ferra L. Seed Oil. Energy and Enviroment Research. 4.2014.
102. Ang, G.T., Ooi, S.N., Tan, K.T., et al., Optimization and kinetic studies of
sea mango (Cerbera odollam) oil for biodiesel production via supercritical
reaction. Energy Conversion and Management. 99. 242.2015.
103. Ramadhas, A.S., Jayaraj, S., and Muraleedharan, C., Biodiesel production
from high FFA rubber seed oil. Fuel. 84(4). 335.2005.
104. A.V.S.L.Sai, B., Subramaniapillai, N., Khadhar Mohamed, M.S.B., et al.,
Optimization of continuous biodiesel production from rubber seed oil (RSO)
using calcined eggshells as heterogeneous catalyst. Journal of Environmental
Chemical Engineering. 8(1). 103603.2020.
105. Le, H.N.T., Imamura, K., Watanabe, N., et al., Biodiesel Production from
Rubber Seed Oil by Transesterification Using a Co-solvent of Fatty Acid
Methyl Esters.41(5). 1013.2018.
106. Patel, V.R., Dumancas, G.G., Kasi Viswanath, L.C., et al., Castor Oil:
Properties, Uses, and Optimization of Processing Parameters in Commercial
Production. Lipid Insights. 9. 1.2016.
107. Mubofu, E.B., Castor oil as a potential renewable resource for the production
of functional materials. Sustainable Chemical Processes. 4(1). 11.2016.
108. Elango, R.K., Sathiasivan, K., Muthukumaran, C., et al., Transesterification
of castor oil for biodiesel production: Process optimization and characteriza-
tion. Microchemical Journal. 145. 1162.2019.
109. Sánchez, N., Encinar, J.M., Nogales, S., et al., Biodiesel Production from
Castor oil by Two-Step Catalytic Transesterification: Optimization of the
Process and Economic Assessment.9(10). 864.2019.
110. Meneghetti, S.M.P., Meneghetti, M.R., Wolf, C.R., et al., Biodiesel from
Castor Oil: A Comparison of Ethanolysis versus Methanolysis. Energy &
Fuels. 20(5). 2262.2006.
Biodiesel Production From Non-Edible Sources 423

111. Benavides, A., Benjumea, P., and Pashova, V., Castor Oil Biodiesel as an
Alternative Fuel for Diesel Engines. Vol. 74. 141. 2007.
112. Ramezani, K., Rowshanzamir, S., and Eikani, M.H., Castor oil transesteri-
fication reaction: A kinetic study and optimization of parameters. Energy.
35(10). 4142.2010.
113. Moradi, G. and Ghanadi, T., An experimental investigation of direct bio-
diesel production from castor seed using waste resource as economical cata-
lyst.38(5). 13180.2019.
114. Du, L., Li, Z., Ding, S., et al., Synthesis and characterization of carbon-based
MgO catalysts for biodiesel production from castor oil. Fuel. 258. 116122.2019.
115. Kohls, S., Scholz-Böttcher, B.M., Teske, J., et al., Cardiac glycosides from
Yellow Oleander (Thevetia peruviana) seeds. Phytochemistry. 75. 114.2012.
116. Deka, D.C. and Basumatary, S., High quality biodiesel from yellow oleander
(Thevetia peruviana) seed oil. Biomass and Bioenergy. 35(5). 1797.2011.
117. Najdanovic-Visak, V., Lee, F.Y.-L., Tavares, M.T., et al., Kinetics of extraction
and in situ transesterification of oils from spent coffee grounds. Journal of
Environmental Chemical Engineering. 5(3). 2611.2017.
118. Fadhil, A.B., Aziz, A.M., and Al-Tamer, M.H., Biodiesel production from
Silybum marianum L. seed oil with high FFA content using sulfonated car-
bon catalyst for esterification and base catalyst for transesterification. Energy
Conversion and Management. 108. 255.2016.
119. Ibrahim, H., Fasanya, O.O., Aminu, H., et al., Fatty Acid Composition of
Mahogany Seed Oil and its Suitability for Biodiesel Production. Nigerian
Journal of Technological Research. 13(1). 45.2018.
120. Orozco, L.M., Echeverri, D.A., Sánchez, L., et al., Second-generation green
diesel from castor oil: Development of a new and efficient continuous-­
production process. Chemical Engineering Journal. 322. 149.2017.
121. Borah, M.J., Devi, A., Borah, R., et al., Synthesis and application of Co doped
ZnO as heterogeneous nanocatalyst for biodiesel production from non-­
edible oil. Renewable Energy. 133. 512.2019.
122. Sut, D., Chutia, R.S., Bordoloi, N., et al., Complete utilization of non-edible
oil seeds of Cascabela thevetia through a cascade of approaches for biofuel
and by-products. Bioresource Technology. 213. 111.2016.
123. Chung, K.-H., Kim, J., and Lee, K.-Y., Biodiesel production by transesterifica-
tion of duck tallow with methanol on alkali catalysts. Biomass and Bioenergy.
33(1). 155.2009.
124. Liu, S., Wang, Y., Oh, J.-H., et al., Fast biodiesel production from beef tallow
with radio frequency heating. Renewable Energy. 36(3). 1003.2011.
125. Okwundu, O.S., El-Shazly, A.H., and Elkady, M., Comparative effect of reac-
tion time on biodiesel production from low free fatty acid beef tallow: a defi-
nition of product yield. SN Applied Sciences. 1(2). 140.2019.
126. Booramurthy, V.K., Kasimani, R., Pandian, S., et al., Nano-sulfated zirconia
catalyzed biodiesel production from tannery waste sheep fat. Environmental
Science and Pollution Research.2020.
424 Biodiesel Technology and Applications

127. Hariprasath, P., Vijayakumar, V., Selvamani, S.T., et al., Some Studies on Waste
Animal Tallow Biodiesel Produced by Modified Transesterification Method
Using Heterogeneous Catalyst. Materials Today: Proceedings. 16. 1271.2019.
128. Idowu, I., Pedrola, M.O., Wylie, S., et al., Improving biodiesel yield of animal
waste fats by combination of a pre-treatment technique and microwave tech-
nology. Renewable Energy. 142. 535.2019.
129. Vlahopoulou, G., Petretto, G.L., Garroni, S., et al., Variation of density and
flash point in acid degummed waste cooking oil.42(3). e13533.2018.
130. Lopes, M., Miranda, S.M., Alves, J.M., et al., Waste Cooking Oils as Feedstock
for Lipase and Lipid-Rich Biomass Production.121(1). 1800188.2019.
131. Tsoutsos, T.D., Tournaki, S., Paraíba, O., et al., The Used Cooking Oil-to-
biodiesel chain in Europe assessment of best practices and environmental
performance. Renewable and Sustainable Energy Reviews. 54. 74.2016.
132. Kulkarni, M.G. and Dalai, A.K., Waste Cooking OilAn Economical Source
for Biodiesel: A Review. Industrial & Engineering Chemistry Research. 45(9).
2901.2006.
133. Math, M.C., Kumar, S.P., and Chetty, S.V., Technologies for biodiesel produc-
tion from used cooking oil — A review. Energy for Sustainable Development.
14(4). 339.2010.
134. Cordero-Ravelo, V. and Schallenberg-Rodriguez, J., Biodiesel production as
a solution to waste cooking oil (WCO) disposal. Will any type of WCO do for
a transesterification process? A quality assessment. Journal of Environmental
Management. 228. 117.2018.
135. Dimian, A.C. and Kiss, A.A., Eco-efficient processes for biodiesel production
from waste lipids. Journal of Cleaner Production. 239. 118073.2019.
136. El-Sheekh, M. and Abomohra, A.E.-F., Biodiesel Production from Microalgae.
2016.
137. Abomohra, A.E.-F., El-Naggar, A.H., and Baeshen, A.A., Potential of mac-
roalgae for biodiesel production: Screening and evaluation studies. Journal
of Bioscience and Bioengineering. 125(2). 231.2018.
138. Chojnacka, K. and Marquez-Rocha, F.-J., Kinetic and Stoichiometric
Relationships of the Energy and Carbon Metabolism in the Culture of
Microalgae. Biotechnology. 3. 21.2004.
139. Avagyan, A.B. and Singh, B., Biodiesel from Algae, in Biodiesel: Feedstocks,
Technologies, Economics and Barriers: Assessment of Environmental Impact
in Producing and Using Chains, A.B. Avagyan and B. Singh, Editors. 2019,
Springer Singapore: Singapore. p. 77.
140. Kröger, M. and Müller-Langer, F., Review on possible algal-biofuel produc-
tion processes. Biofuels. 3(3). 333.2012.
141. Khan, I., Shah, S., Ayaz, M., et al., Production of Biodiesel from Algae.
Journal of Pure and Applied Microbiology. 9. 79.2015.
142. Kumar, D. and Singh, B., Algal biorefinery: An integrated approach for sus-
tainable biodiesel production. Biomass and Bioenergy. 131. 105398.2019.
Biodiesel Production From Non-Edible Sources 425

143. Chen, J., Li, J., Dong, W., et al., The potential of microalgae in biodiesel pro-
duction. Renewable and Sustainable Energy Reviews. 90. 336.2018.
144. Enamala, M.K., Enamala, S., Chavali, M., et al., Production of biofuels
from microalgae - A review on cultivation, harvesting, lipid extraction, and
numerous applications of microalgae. Renewable and Sustainable Energy
Reviews. 94. 49.2018.
145. Mitra, D., van Leeuwen, J., and Lamsal, B., Heterotrophic/mixotrophic cul-
tivation of oleaginous Chlorella vulgaris on industrial co-products. Algal
Research. 1(1). 40.2012.
146. Shekh, A.Y., Shrivastava, P., Krishnamurthi, K., et al., Stress-induced lipids
are unsuitable as a direct biodiesel feedstock: A case study with Chlorella
pyrenoidosa. Bioresource Technology. 138. 382.2013.
147. Abou-Shanab, R.A.I., Matter, I.A., Kim, S.-N., et al., Characterization and
identification of lipid-producing microalgae species isolated from a fresh­
water lake. Biomass and Bioenergy. 35(7). 3079.2011.
148. Anitha, S. and Narayanan, J.S., Isolation and identification of microalgal
strains and evaluation of their fatty acid profiles for biodiesel production. Int
J Pharm Biol Arch. 3. 939.2012.
149. Ryu, B.-G., Kim, K., Kim, J., et al., Use of organic waste from the brewery
industry for high-density cultivation of the docosahexaenoic acid-rich
microalga, Aurantiochytrium sp. KRS101. Bioresource Technology. 129.
351.2013.
150. Pittman, J.K., Dean, A.P., and Osundeko, O., The potential of sustainable
algal biofuel production using wastewater resources. Bioresource Technology.
102(1). 17.2011.
151. Chen, Y.-H., Huang, B.-Y., Chiang, T.-H., et al., Fuel properties of microalgae
(Chlorella protothecoides) oil biodiesel and its blends with petroleum diesel.
Fuel. 94. 270.2012.
152. Islam, M.A., Magnusson, M., Brown, R.J., et al., Microalgal Species Selection
for Biodiesel Production Based on Fuel Properties Derived from Fatty Acid
Profiles.6(11). 5676.2013.
153. Talebi, A.F., Mohtashami, S.K., Tabatabaei, M., et al., Fatty acids profiling: A
selective criterion for screening microalgae strains for biodiesel production.
Algal Research. 2(3). 258.2013.
154. Pandit, P.R. and Fulekar, M.H., Biodiesel production from microalgal bio-
mass using CaO catalyst synthesized from natural waste material. Renewable
Energy. 136. 837.2019.
155. Rajak, U., Nashine, P., and Verma, T.N., Effect of spirulina microalgae bio-
diesel enriched with diesel fuel on performance and emission characteristics
of CI engine. Fuel. 268. 117305.2020.
156. Selvarajan, R., Felföldi, T., Tauber, T., et al., Screening and Evaluation of
Some Green Algal Strains (Chlorophyceae) Isolated from Freshwater and
Soda Lakes for Biofuel Production.8(7). 7502.2015.
426 Biodiesel Technology and Applications

157. Shumo, M., Osuga, I.M., Khamis, F.M., et al., The nutritive value of black sol-
dier fly larvae reared on common organic waste streams in Kenya. Scientific
Reports. 9(1). 10110.2019.
158. Feng, W., Qian, L., Wang, W., et al., Exploring the potential of lipids from
black soldier fly: New paradigm for biodiesel production (II)—Extraction
kinetics and thermodynamic. Renewable Energy. 119. 12.2018.
159. Liu, Z., Najar-Rodriguez, A.J., Minor, M.A., et al., Mating success of the black
soldier fly, Hermetia illucens (Diptera: Stratiomyidae), under four artificial
light sources. Journal of Photochemistry and Photobiology B: Biology. 205.
111815.2020.
160. Manzano-Agugliaro, F., Sanchez-Muros, M.J., Barroso, F.G., et al., Insects
for biodiesel production. Renewable and Sustainable Energy Reviews. 16(6).
3744.2012.
161. Wong, C.-Y., Lim, J.-W., Uemura, Y., et al., Insect-based lipid for biodiesel
production.2016(1). 020150.2018.
162. Liu, T., Awasthi, M.K., Awasthi, S.K., et al., Effects of black soldier fly lar-
vae (Diptera: Stratiomyidae) on food waste and sewage sludge composting.
Journal of Environmental Management. 256. 109967.2020.
163. Zheng, L., Li, Q., Zhang, J., et al., Double the biodiesel yield: Rearing black
soldier fly larvae, Hermetia illucens, on solid residual fraction of restaurant
waste after grease extraction for biodiesel production. Renewable Energy. 41.
75.2012.
164. Yang, S., Li, Q., Zeng, Q., et al., Conversion of Solid Organic Wastes into Oil via
Boettcherisca peregrine (Diptera: Sarcophagidae) Larvae and Optimization
of Parameters for Biodiesel Production. PLOS ONE. 7(9). e45940.2012.
165. Leung, D., Yang, D., Li, Z., et al., Biodiesel from Zophobas morio Larva Oil:
Process Optimization and FAME Characterization. Industrial & Engineering
Chemistry Research. 51(2). 1036.2012.
166. Zi-zhe, C., De-po, Y., Sheng-qing, W., et al., Conversion of poultry manure to
biodiesel, a practical method of producing fatty acid methyl esters via house-
fly (Musca domestica L.) larval lipid.Fuel. 210. 463.2017.
167. Niu, Y., Zheng, D., Yao, B., et al., A novel bioconversion for value-added prod-
ucts from food waste using Musca domestica. Waste Manag. 61. 455.2017.
168. Ishak, S. and Kamari, A., Biodiesel from black soldier fly larvae grown on
restaurant kitchen waste. Environmental Chemistry Letters. 17(2). 1143.2019.
169. Kombe, G.G. and Temu, A.K., Steam Deacidification of High Free Fatty Acid
in Jatropha Oil for Biodiesel Production. Energy & Fuels. 31(6). 6206.2017.
170. Habaki, H., Hayashi, T., and Egashira, R., Deacidification process of crude
inedible plant oil by esterification for biodiesel production. Journal of
Environmental Chemical Engineering. 6(2). 3054.2018.
171. Yeom, S.H. and Go, Y.W., Optimization of a Novel Two-step Process
Comprising Re-esterification and Transesterification in a Single Reactor for
Biodiesel Production Using Waste Cooking Oil. Biotechnology and Bioprocess
Engineering. 23(4). 432.2018.
Biodiesel Production From Non-Edible Sources 427

172. Muanruksa, P., Winterburn, J., and Kaewkannetra, P., A novel process for
biodiesel production from sludge palm oil. MethodsX. 6. 2838.2019.
173. Ito, T., Sakurai, Y., Kakuta, Y., et al., Biodiesel production from waste animal
fats using pyrolysis method. Fuel Processing Technology. 94(1). 47.2012.
174. Laksmono, N., Paraschiv, M., Loubar, K., et al., Biodiesel production from
biomass gasification tar via thermal/catalytic cracking. Fuel Processing
Technology. 106. 776.2013.
175. Ramkumar, S. and Kirubakaran, V., Biodiesel from vegetable oil as alternate
fuel for C.I engine and feasibility study of thermal cracking: A critical review.
Energy Conversion and Management. 118. 155.2016.
176. Santos, A.L.F., Martins, D.U., Iha, O.K., et al., Agro-industrial residues as
low-price feedstock for diesel-like fuel production by thermal cracking.
Bioresource Technology. 101(15). 6157.2010.
177. Abdelfattah, M.S.H., Abu-Elyazeed, O.S.M., Abd El mawla, E., et al., On bio-
diesels from castor raw oil using catalytic pyrolysis. Energy. 143. 950.2018.
16
Microalgae for Biodiesel Production
Charles Oluwaseun Adetunji1*, Victoria Olaide Adenigba2,
Devarajan Thangadurai3 and Mohd Imran Ahamed4

Applied Microbiology, Biotechnology and Nanotechnology Laboratory,


1

Department of Microbiology, Edo State University Uzairue, PMB 04,


Auchi, Edo State, Nigeria
2
Department of Science Laboratory Technology, Ladoke Akintola University of
Technology, Ogbomosho, Nigeria
3
Department of Botany, Karnatak University, Dharwad, India
4
Department of Chemistry, AMU, Aligarh, India

Abstract
It has been observed that the application of petroleum fuel is not sustainable and
environment-friendly due to the release of high level of contaminants as well as
the high amount of money involved in their production. Hence, there is a need to
search for carbon neutralization, transport fuels, and renewable, cost-effective, and
eco-friendly energy with little amount of NOx and less CO2 emissions. The utiliza-
tion of algae has been identified as a great source of biodiesel generation without
any adverse effect. Therefore, this chapter intends to provide detailed information
on the fabrication of biodiesel from microalgae. Specific information on the phys-
ical properties, amount of biodiesel production, and level of transesterification of
biodiesel were discussed. The application of photobioreactors for the production
of biodiesel with the special consideration of several factors, such as flow rate,
temperature, light intensity, CO2 concentration, and time, was highlighted. Several
techniques for the extraction of biodiesel such as supercritical CO2, physicochem-
ical, direct transesterification, chemical solvents, and biochemical, respectively,
were highlighted. The structural elucidation of the biodiesel produced by micro­
algae using high pressure liquid chromatography, thin-layer chromatography, gas
chromatography, Fourier transform infrared spectroscopy, and nuclear magnetic
resonance was highlighted.

*Corresponding author: charliguitar@yahoo.com; adetunjicharles@gmail.com;


adetunji.charles@edouniversity.edu.ng

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (429–446) © 2021 Scrivener Publishing LLC

429
430 Biodiesel Technology and Applications

Keywords: Transesterification, biodiesel, microalgae, petroleum fuel, cost


effective, eco-friendly

16.1 Introduction
The total dependence on fossil fuel as a source of energy has become a
major problem in the world’s economy with potential challenge in energy
depletion and environmental pollution [13]. The increase in the world’s
population has a direct impact on the energy reserve, thereby resulting
in depletion in the energy reserve as a result of overdependence on a
non-renewable energy source. In addition to the explosion in popula-
tion, another challenge from dependence on fossil fuel is the production
of toxic gases during refinery process. The presence of these gases in the
environment results in global warming which is another environmental
concern [13].
The need for energy sources which are renewable and at the same time
environmental compliant cannot be overemphasized [32]. Microalgae
possess the potential as a feedstock for production of renewable energy
in place of fossil fuel [2]. Microalgae are considered as third-generation
biofuel and they have several advantages over both first-generation biofuel
(feedstock such as cash crops) and second-generation biofuel (lignocel-
lulosic feedstock) [36]. The merit of microalgae as feedstock for biofuels
includes rapid growth in a variety of aquatic environment, ease of cultiva-
tion without dependence on large farm land and food industry, ability to
absorb atmospheric CO2 for carbon fixation, and maximum biomass pro-
duction within a short time and high lipid invention [24, 29]. Microalgae
are able to biodegrade inorganic matter present in wastewater, and this
alleviates the change of supplying nutrients to the culture and these nutri-
ents attract high cost.
Algae could produce renewable energy due extraordinary lipid and bio-
mass production, photosynthetic nature, and rapid growth rate [4]. Algal
cultivation can play an extensive role in carbon emission reduction as it
captures large amount of CO2 (up to 2 kg of CO2 per 1 kg of dry biomass)
due to photosynthesis [34]. Algae can be produced through autotrophic
or chemoheterotrophic conditions by using solar energy and an artificial
light source. Furthermost of their indispensable nutrient can be provided
by wastewater and CO2 from the atmosphere principal to high efficiency
and an connected extraordinary lipid content creation them a very attrac-
tive alternative [1].
Microalgae for Biodiesel Production 431

Microalgae could be used in the generation of methanol [9], bio-oil [18],


hydrogen [14], and methane [19]. There are several ways in which micro­
algae can be used in the production of energy in different forms which
are sustainable and can be used to alleviate the impending challenges of
depletion of fossil fuel.
Therefore, this chapter has provided detailed information on the sus-
tainable production of biodiesel from microalgae. Specific information
on the physical properties, amount of biodiesel production, and level of
transesterification of biodiesel were discussed. The application of photo-
bioreactors for the production of biodiesel with the special consideration
of several factors such as flow rate, temperature, light intensity, CO2 con-
centration, and time was highlighted. Several techniques for the extraction
of biodiesel such as supercritical CO2, physicochemical, direct transesteri-
fication, chemical solvents, and biochemical respectively were highlighted.
The structural elucidation of the biodiesel produced by microalgae using
high pressure liquid chromatography, gas chromatography, Fourier trans-
form infrared spectroscopy, thin-layer chromatography, and many more
was highlighted.

16.2 Physicochemical Properties of Biodiesel


From Microalgae
The properties of biodiesel depend on the strain of microalgae from which
the oil was extracted. There are several techniques which can be adopted
in the determination of both quantitative and qualitative characteristics
of biodiesel; some of these techniques are with flame ionization detector
which is specifically used in the analysis of feedstock for the fabrication of
biodiesel [20]. It is expedient to first determine the composition of lipid in
an alga of choice for biodiesel invention. The presence of poly unsaturated
fatty acids which has at least four double bonds is a characteristic of oils
obtained from algae [29].
The properties that any fuel must possess in order to fit in as a replace-
ment for fossil fuel include stability to oxidation, the extent of lubricat-
ing, viscosity, ignition, and combustion potential [37]. The length of chain,
quantity of fatty esters, and quantity of double bonds also regulate the
chemical and physical characteristics of the triglycerides and fatty acids
and obtained in microalgae-based biodiesel [27]. There are more physical
and chemical properties that define the suitability of biodiesel produced
by microalgae and other petroleum diesel. The standards that determine
432 Biodiesel Technology and Applications

the quality of biodiesel are controlled by the European Committee for


Standardization in the European Union and many other relevant organi-
zations [6].
The chemical and physical features of biodiesel from microalgae differ
from one strain to another. The following are some of the parameters that
measure the suitability of microalgae biodiesel as a dependable and sus-
tainable biofuel: kinetic viscosity and the chain length of fatty acid along
with the degree of unsaturation that determines how viscous a microalgae
biodiesel will be. A high degree of unsaturation can result in modifica-
tion of fatty acid composition at the different phases which are transpor-
tation and fluctuation in temperature and storage [17]. Soares et al. [35]
described that the development condition and extraction of lipid have a
tendency of affecting the fatty acid profile and the final features of the bio-
diesel manufactured.
Another important property is acid value known as the quantity of free
fatty acids available in the biodiesel and this is determined by the quantity
of potassium hydroxide which will be needed for neutralization. Also a
biodiesel can be defined by its cetane number which shows the effect of
biodiesel on engine ignition [5]. The lowest temperature at which a known
quantity of diesel will flow through a standard filter in 1 minute is referred
to as cold filter plugging point [16] while the cloud point is the lowest tem-
perature when formation of crystal in diesel can be visualized as a suspen-
sion which appears cloudy [31]. These two properties are also crucial in
determining the status of biodiesel produced by microalgae. Knothe [16]
further explain that the content of saturated fatty acids has an impact on
both cold filter plugging point and cloud point.
Flash point is another temperature-related property of biodiesel and it is
the lowest temperature required for the oil to give the vapor which activates
ignition and subsequent production of flame. Biodiesel from microalgae
has a lower flash point compared to that from some other sources [31]. The
purity of the biodiesel produced depends on the content of ester and lino-
lenic acid. Microalgae possess unique abilities to regulate their metabolic
activities based on the growth conditions [6].

16.3 Genetic Engineering/Techniques Enhancing


Biodiesel Production
Microalgae multiply rapidly through cell division and this accounts for
their fast growth rate which has presented them as a more preferable
feedstock for biodiesel production. Their ability to divide rapidly has
Microalgae for Biodiesel Production 433

given access for genetic manipulation to be able to produce strains with


improved potential for biofuel production [7]. Apart from rapid repro-
ducibility, a large number of algae (micro and macro) are known for their
ability to accumulate fatty acids which have up to 16–20 carbon in terms
of length [10]. In order to obtain maximum biodiesel production from any
strain of algae, the knowledge of the liquid content in relation to length of
carbon chain is required. This knowledge will be used for manipulation
which is expected to yield trait improvement through bioengineering of
algae strain with biofuel production that can contribute greatly to yield an
improved output [38].
A large number of algae (micro and macro) are known for their ability
to accumulate fatty acids with the carbon length as much as 20. In order
to obtain maximum biodiesel production from any strain of algae, the
knowledge of the lipid content in relation to the length of carbon chain is
required. This will be a template for manipulation with the expectation of
a better yield [7]. The production of biodiesel by micro­algae can only be
a replacement for fossil fuel if the adequate techniques are put in place to
enhance the production process in terms of time, quality, and quantity of
fuel produced.
The economic viability of biodiesel from algae is still a concern since the
cost of operation and corresponding output is lower than what is required
to meet the energy demand (Gouveia et al., 2017).
In algae, the chloroplast is the center of activity where lipids are syn-
thesized with fixed atmospheric CO2 as a source of acetyl-CoA, and it is
responsible for the production of the different carbon atoms available in
the chain of the fatty acid [60]. Genetic engineering targeting the chloro-
plast of the algae could result help in improving the biodiesel produced.
Zienkiewicz et al. [63] also discussed the importance of genetic tools
in characterizing the exact gene responsible for biosynthesis of lipids in
microalgae. They further highlighted some microalgae which have the
potential as research tool in the synthesis of both triacylglycerol and lipid
synthesis; these are Phaeodactylum tricornutum, Chlamydomonas rein-
hardtii, Ostreococus tauri, and Chlorella spp.
The production of biomass and high lipid are two important parame-
ters that will favor biodiesel production by any microalgae strain. There
are some enzymes which are directly involved in the biosynthesis of bio-
diesel from microalgae. Overexpression of the target enzyme responsible
for the synthesis of fatty acids is very crucial. Davidson et al. [3] pointed
out the fact that one of the ways by which biofuel can replace fossil fuel is
by genetically engineering key enzymes to promote large-scale production
of biofuel.
434 Biodiesel Technology and Applications

16.4 Nanotechnology in Microalgae Biodiesel


Production
Apart from genetic engineering, another technology which can enhance
the quality and quantity of biodiesel production from microalgae is nan-
otechnology through the addition of nanomaterial in the form of nano-
droplets, nanocrystals, nanomagnet, and nanofiber [21, 23]. Application of
nanotechnology in biofuel production can be employed at different point
in the production process. Reports have been documented on the contri-
bution of nanomaterial as catalyst in the production of biofuel with a con-
sequential improvement in the fuel produced [8, 11]. Biomass production
and cell harvesting are crucial to algae cell culture for biodiesel produc-
tion, and there are various factors that contribute to biomass production
and one of these important factors is light intensity. If the light available to
the algae culture is not adequately and uniformly distributed in the pho-
toreactor, then it can have a negative impact on biomass production and
cell harvesting. Addition of nanoparticles to the cell culture of microalgae
has been employed in order to improve light distribution in the reactor
and this resulted in improvement in yield and ease in cell harvesting [30].
Biofuel doping is another area where nanoparticles have been found use-
ful, and the addition of nanoparticles to fuel enhances fuel efficiency of
engines [12, 26].

16.5 Specific Examples of Biodiesel Production


From Microalgae
Microalgae over the last few decades till date have been considered as
a major source of biofuel, their rapid growth, friendliness in handling,
and ultimately the absence of threat to the food industry in terms of
cultivation accounts for global choice as feedstock for biofuel produc-
tion especially biodiesel [13, 25, 28]. Some microalgae with high lipid
production have been investigated for their biodiesel producing poten-
tial. Neochloris oleoabundans, Spirulina maxima, Chlorella vulgaris,
Scenedesmus obliquus, Dunaniella tertiolecta, and Nannochloropsis
sp. were screened as feedstock for biofuel production. Scenedesmus
obliquus had the most suitable fatty acid profile (polyunsaturated fatty
acids and linolenic) required to be used in the production of biodiesel
while both Nannochloropsis sp. and Neochloris oleoabundans had high
Microalgae for Biodiesel Production 435

oil content when the culture was put under nitrogen stress [6]. The
potential of Nannochloropsis sp. as feedstock for bioduesel production
was also documented by Nobre et al. [22]. Both fatty acids and carot-
enoids were targeted and an appreciable quantity of lipids (45 g/100
g biomass) and pigments (70%) was obtained. Euglena sanguinea was
equally screened for biodiesel production and both saturated and unsat-
urated fatty acids produced were confirmed with the possible poten-
tial of being used as fuel in the present-day automobiles [15]. Sharif
et al. [33] also reported the potential of Oedogonium spp. and Spirogyra
spp. for biodiesel production. Biomass production was greater in
Spirogyra while Oedogonium sp. had a biodiesel production. Both algal
strains were confirmed as biodiesel producers.
The application of biodegradable fuel, with little amount of NOx and
less CO2 emissions. It has been stated that the utilization of petroleum fuel
has been recognized not to be sustainable which might be linked to a high
reduction in the amount of supplies and the buildup of the carbon dioxide
in the environment is not eco-friendly. Therefore, the utilization of carbon
neutralization, transport fuels, and renewable energy has been recognized
as an alternative option because it is eco-friendly and cost effective. The
utilization of algae has been identified as a great source of biodiesel genera-
tion. It has been discovered that brown algae cultivated in CO2-enriched air
could be transformed into oily substances. It has been identified that such
a techniques could play a crucial role in the resolution of all the significant
problems regarding the release of pollutant generated from CO2 evolution
as well as the forthcoming predicament due to decrease of energy source.
Hossain and Salleh [46] performed a research study on the physical prop-
erties, amount of biodiesel production, and level of transesterification of
biodiesel. The authors utilized common species Spirogyra and Oedogonium
so as to evaluate the level of biodiesel fabrication. It was discovered that the
biodiesel and algal oil generation was in higher quantity in Oedogonium
spp. when compared to Spirogyra spp. Moreover, it was discovered that the
biomass obtained immediately after the extraction was in a higher qual-
ity in Spirogyra than Oedogonium spp. Also, the sediment obtained from
Spirogyra spp. than Oedogonium spp. in terms of water, pigments, and glyc-
erine was in a higher quantity, while there was no observable variation
in their pH. The result obtained indicated that the biodiesel derived from
the Oedogonium spp. is showed more quality when compared to that of
Spirogyra spp.
Fossil fuel has been recognized as a non-renewable, and the volume of
the required quantities has diminished over period of time due the high
436 Biodiesel Technology and Applications

rate of pumping. The high rate in the global population has led to increase
in traffic which always necessitate more fuel to meet their demand as well
as increase in the prices of crude oil and gas. Diesel could be derived from
numerous feedstock such as algae, sunflower, palm, oil rape, waste animal
fats, and soybean respectively. Bošnjaković [45] evaluates the possibilities
of utilizing biodiesel derived from algae and the potential of utilizing such
techniques for generation of biodiesel. The biodiesel derived from algae
was evaluated and the system that could support their mass production
was also highlighted. The authors stated that algae are very grow anywhere
in the desert most especially in fresh and salt water even in the presence
of waste water and with resistance in nature. It has been realized that they
multiply quickly and they possess the capability to utilize CO2 for photo-
synthesis and require minimal amount of water when compared to other
crops. Also, the biofuel derived from algae is biodegradable without the
presence of sulfur without the presence of any contaminants.
Biodiesel has been recognized as a natural biofuel with several attri-
butes which includes lower net carbon cycle, eco-friendly nature, bio-
degradability, and non-toxic characteristics, respectively. In view of the
aforementioned, Ihsanullah et al. [44] utilize the effectiveness of algal spe-
cies of spirogyra derived from numerous region of Khyber Pakhtunkhwa,
Pakistan, and the algae were engaged as a feedstock for biodiesel fabrication.
Initially, the oil derived from algae species was obtained utilizing di-ethyl
ether and n-hexane, while the second stage which involves the process of
oil extraction was obtained using the dig out oil converted into biodiesel
through the process of transestrification rejoinder. Furthermore, the level
of biodiesel generated was quantified based on the following parameters:
temperature, reaction time, amount of catalyst (sodium hydroxide), and
influence of molar ratio during the period of transestrification reaction.
Beetul et al. [43] performed an investigation that evaluates the amount
of lipid from endosymbiotic dinoflagellates and the cyanobacterial mats
were determined using genetically and morphology utilizing RFLP. The
level of samples was evaluated utilizing 1H and 13C NMR spectroscopy and
quantified gravimetrically, respectively.
Microalgae have been identified to possess that capability to generate
proteins, lipids, and carbohydrates as significant fractions. Typical example
of biofuel includes biogas, biohydrogen, biodiesel, bio-oil, and bioethanol
through numerous processes and conditions. Moreover, the application of
microalgae has been shown to possess the capability to serve as a replace-
ment for CO2 emissions fixation because it has been observed that the gas
could be utilized for their metabolism. Moreover, microalgae have been
considered as a feedstock that could be incorporated a bio-refinery, the
Microalgae for Biodiesel Production 437

process by which numerous product can derived from biofuel to food.


Cuellar-Bermudez et al. [42] carried out an extensive review on the appli-
cation of biofuel derived from microalgae, and the disadvantage and lim-
itation that could prevent increase in the production of microalgae were
also elaborated.
It has been stated that the recent fossil fuel reserve might not be enough
to meet the daily increase in request for the biofuel. The high increase in
price of oils, pollution, and global warming has led to increase in request in
renewable energy sources. Therefore, the utilization of biofuel derived from
algae might represent a typical example of renewable energy sources. The
biofuel obtained from algae feedstock that could be utilized for production
of bioethanol and biodiesel generation due to the fact that they could be
easily be mass produced with interesting biomass yield, high level of car-
bohydrate, and lipids contents. This shows that the quantity and quality of
biodiesel derived from algae might surpass those derived from second- and
first-generation feedstock. It has been stated that only few algae species
have been examined as a probable biofuel sources. Therefore, Saad et al.
[41] wrote a comprehensive review on the application algae for sustainable
production of biofuel, which significantly needs to be produced, and some
of the problem encountered during their commercial production.
The increase in the level of the carbon dioxide (CO2) in the atmosphere
could be linked to the numerous dangerous activities generating numer-
ous changes in the world carbon cycle which draw the attention of sev-
eral people around the globe. The availability of the carbon dioxide in the
atmosphere could be linked to several processes such as volcanic eruption,
forest fires, power plants, decomposition of organic matters, and automo-
biles. All these CO2 could be utilized through the help of microalgae which
could be subsequently be converted into the lipids for the production of
biofuel as a substitute petroleum-product transport fuel starved of having
any adverse effect on the supply of food and crop. Therefore, in view of
these aforementioned, Mondal et al. [40] wrote a comprehensive review
on the recent advances on the application of microalgae for the utilization
of the carbon capture for the production of biodiesel. Numerous features
such as selection of efficient strains, production inside closed and open
system, metabolism of microalgae, and production of biomass together
with international and natural biodiesel specification. The application of
photobioreactors for the production of biodiesel with the consideration of
several factors such as flow rate, temperature, light intensity, CO2 concen-
tration, and time was highlighted. Moreover, the economic overview and
the future outlook as well as necessary strategies that could enhance the
production of biodiesel from microalgae were also highlighted.
438 Biodiesel Technology and Applications

The utilization of energy sources has been discovered to be reducing


gradually because of the increase in the demand for a safe biofuel. The
application of microalgae has been identified for the production of a sus-
tainable biofuel because the application of food and crop producing plants
will not be enough to meet the demand required for the production of bio-
diesel. It has been realized that numerous types of algae possess the capa-
bility to produce biodiesel. Almost 20%–80% oil contents are produce from
the algae which can be transformed into numerous varieties of oil such as
biodiesel and kerosene. This might be linked to the fact that the biodiesel
derived from algae is easier and cost effective. Numerous types of algae
have been realized to possess the capability to produce biodiesel. Typical
examples of these algae strain include euglena, ulothrix, and tribonema,
respectively. The application of gene technology could be utilized for the
generation of biodiesel and oil contents and for improving the stability of
algae. This can go a long way to increase the genetic expression required for
increasing the total amount of biofuel available in different strains. Khan
et al. [13] perform a comprehensive review on the application of algae as a
permanent replacement to fossil fuel.
Aljabarin and Al Jarrah [39] performed an experiment that involves
the production of biodiesel from algae. The algae were obtained from
Almansora stream at Tafila Governorate, Jordan, after which it was sub-
jected to drying for a period of 12 hours at 80°C in an oven. The dried
form of the algae was later grind into powder. The powdered algae were
later subjected into the process of distillation through the incorporation
of hydrate solution of FeSO4 to the algae utilizing a distillation column so
as to obtain the oil. Methanol was also incorporated into the oil which was
derived through the process of distillation utilizing KOH as a catalyst. The
product was removed using funnel for effective separation of the two layers
of the solution. The top layer which contains the organic layer denotes the
biodiesel. The biodiesel qualities are comparable to the diesel oil, excluding
that it decreases the production of CO2 without any generation of sulfur
dioxide.

16.6 Methodology Involved in the Extraction of Algae


It has been recognized that biodiesel could be produced from microalgae
through the aids of lipid extraction methodology using supercritical CO2,
physicochemical, direct transesterification, chemical solvents, and bio-
chemical, respectively.
Microalgae for Biodiesel Production 439

16.6.1 Chemical Solvents Extraction


This is one of the most adopted methodology that is mostly utilized but
with little efficiency.
Most especially when the microalgae are in a wet stage [47]. The applica-
tion of freeze-drying techniques is utilized for the extraction of lipids most
especially in the laboratory [48], oven-drying [50], while the other tech-
niques used for drying of microalgae includes vacuum evaporation [49]. It
has been stated that the process involved in the drying of the microalgae
involved the preliminary drying which might necessitate 2.5 greater energy
than a process in the absence of drying [51]. It has been observed that sol-
vent such as hexane is preferable when compared to chloroform-methanol
[52], because they possess lesser affinity for non-lipid pollutants and their
minimal level of toxicity [53]. Moreover, at industrial level, the application
of Soxhlet extraction is not preferable because there is a need for a great
rate of energy [53]. Also, the application of other toxic solvents contain-
ing 8-diazabicyclo-[5.4.0]-undec-7-ene and alcohols (octanol, ethanol)
has been evaluated but five times of the amount of fatty acid methyl ester
was in a lower amount when compared to the extraction carried out sing
n-hexane [47].

16.6.2 Extraction by Supercritical Carbon Dioxide


The supercritical CO2 have several merits which include nontoxic and the
high rate of recovery most especially at a lower temperature which is lesser
than 40°C [54]. This technique has been established to require a very high
amount of equipment with high level of energy which is required during
the process extent great pressures [55]. It has been observed that super-
critical CO2 extraction could be utilized for the recovery of microalgae
lipids and their eventual transformation into biodiesel [53] while some
level of lipid containing n26% (g lipid/g dry weight) from Nannocloropsis
spp. [54].

16.6.3 Extraction Using Biochemical Techniques


It has been stated that little studies have been carried out that involves that
application of biochemical extraction for effective extraction of lipids from
microalgae. Fu et al. [57] utilized a 72-h cellulase break down pretreatment
using microalgae most especially from Chlorella sp. The author discovered
that break down yield of sugars of 70%. The following parameters were
440 Biodiesel Technology and Applications

tested: total sugar concentration and concentration reducing sugar. It was


later observed that there was significant upsurge in the lipid content which
varies from 52% to 54% (g lipid/g dry weight).

16.6.4 Extraction Involving Direct Transesterification


Direct transesterification involves that combination of microalgae together
with an alcohol in the presence of a catalyst. This methodology did not
involve the prior stage of extraction [61].
The quantity of acid catalysts which was examined mainly for the het-
erotopic microalgae biomass entails sulfuric acid (H2SO4) and hydrochloric
(HCl) but the application of acetyl chloride (CH3COCl) could lead to more
enhanced biodiesel [50]. Moreover, the addition of a less polar solvent such
as chloroform or hexane could be added to improve the yield and quality
of biodiesel obtained [62]. Furthermore, the application of heterogeneous
catalyst could be used during direct transesterification in the presence of
microwaves heating.

16.6.5 Extraction Using Transesterification Techniques


It has been observed that the application of basic vegetable oil in engines
utilizing diesels has been established but it might led to several practical
challenges which includes struggle to vaporize in cold circumstances, high
viscosity, and high density. All this could be linked to the deposit most
especially in the ignition chamber without the presence of polluting and an
upsurge in the emission level [59]. Therefore, the application of biodiesel
derived from algae might be an alternative to those derived from the veg-
etative oil [58]. Therefore, in order to overcome all these highlights, there
is a need to base more emphasis on the conversion of microalgae lipids in
equivalent in necessary.

16.7 Conclusion and Future Recommendation


to Knowledge
Microalgae are energy-rich organisms because they are reservoir of lipids
and fatty acids, and this is one of the many reasons they have been called
third-generation biofuel. Out of the various biofuels available, biodiesel
has drawn intense research attention in a bid to overcome the challenge of
Microalgae for Biodiesel Production 441

energy depletion and environmental pollution from the overdependence


and refinery process of fossil fuel, respectively. Even though microalgae
have been appraised as a biofuel producer, there is still need to improve the
invention of biodiesel from microalgae in order to meet up with energy
demand. Different techniques majorly, genetic engineering and nanotech-
nology were discussed in this review as tools that could deliver the required
microalgae biodiesel into the system in terms of both quality and quantity.

References
1. Agwa, O. K., Ibe, S. N., and Abu, G.O. Biomass and lipid production of a
fresh water algae Chlorella sp. using locally formulated media. International
Journal of Microbiology., 3(9):288–295, 2012.
2. Ahmed E.M. Abdelaziz, Gustavo B. Leite & Patrick C. Hallenbeck, Addressing
the challenges for sustainable production of algal biofuels: I. Algal strains
and nutrient supply, Environmental Technology, 34:13–14, 1783–1805, 2014.
3. Davidson S. Sustainable bioenergy: genomics and biofuels development. Nat
Educ 1(1):175, 2008.
4. Feng, Y., Li, C., and Zhang, D. Lipid production of Chlorella vulgaris cultured in
artificial wastewater medium. Bioresource Technology, 102(1): 101–105, 2011.
5. Francisco, E.C., Neves, D.B., Jacob-Lopes, E., Franco, T.T., Microalgae as
feedstock for biodiesel production: carbon dioxide sequestration, lipid pro-
duction and biofuel quality. J. Chem. Technol. Biotechnol. 85:395–403, 2010.
6. Gouveia, Oliveira LAC. Micro algae as a raw material for biofuels produc-
tion. J. Ind. Microbial Biotech 36: 269–274, 2009.
7. Hannon M., Javier Gimpel, Miller Tran, Beth Rasala, and Stephen Mayfield
Biofuels from algae: challenges and potential, Biofuels, 1(5): 763–784, 2010.
8. Hasannuddin AK, Yahya WJ, Sarah S, Ithnin AM, Syahrullail S, Sidik NAC,
et al. Nano-additives incorporated water in diesel emulsion fuel: fuel prop-
erties, performance and emission characteristics assessment. Energy Convers
Manag., 169:291–314, 2018.
9. Hirano, A., Hon-Nami, K.,Kunito, S, Hada, M., and Ogushi, Y. Temperature
effect on continuous gasification of microalgal biomass: Theoretical yield
of methanol production and its energy balance. Catal Today, 45(1–4):
399–404, 1998.
10. Jagadevan, S., Banerjee, A., Banerjee, C., Guria, C., Tiwari, R., Baweja, M.,
and Shukia, P., Recents developments in synthetic biology and metabolic
engineering in microalgae towards biofuel production. Biotechnol. Biofuels,
11:185–196, 2018.
11. Karthikeyan S, Prathima A. Microalgae biofuel with CeO2 nano additives as
an eco-friendly fuel for CI engine. Energy Source Part A., 39:1332–8, 2017.
442 Biodiesel Technology and Applications

12. Karthikeyan S, Prathima A. Environmental effect of CI engine using microal-


gae methyl ester with doped nano additives. Transp Res D Transp Environ.,
50:385–96, 2017.
13. Khan, S., Siddique, R., Sajjad, W., Nabi, G., Hayat, K.M., Duan, P., and Yao,
L., HAYATI. Journal of Biosciences, 24: 163–167, 2017.
14. Kim, J. R., Jung, S. H., Regan, J. M., and Logan, B. E. Electricity generation
and microbial community analysis of alcohol powered microbial fuel cells.
Bioresource Technology, 98:2568–2577, 2006.
15. Kings AJ, Raj RE, Miriam LRM, Visvanathan M.A. Cultivation, extraction
and optimization of biodiesel production from potential microalgae Euglena
sanguinea using eco-friendly natural catalyst. Energy Convers Manag.,
141:224–35, 2017.
16. Knothe, G., Dependence of biodiesel fuel properties on the structure of fatty
acid alkylesters. Fuel Process. Technol. 86: 1059–1070, 2005.
17. Knothe, G., Steidley, K.R., Kinematic viscosity of biodiesel components
(fatty acid alkylesters) and related compounds at low temperatures. Fuel, 86:
2560–2567.
18. Li, H. B, Cheng, K. W, Wong, C. C., Fan, K.W, Chen F., and Jiang, Y. Evaluation
of antioxidant capacity and total phenolic content of different fractions of
selected microalgae. Food Chemistry, 102:771–776, 2007.
19. Minowa, T., and Sawayama, S. A novel microalgal system for energy produc-
tion with nitrogen cycling. Fuel., 78(10):1213–1215, 1999.
20. Mutanda, T., Ramesh, D., Karthikeyan et al., Kumari, S.Anandraj, A.,
and Bux, F., Bioprospecting for hyper-lipid producing microalgal strains
for sustainable biofuel production. Bioresource Technology, 102:57–70,
2011.
21. Nizami A, Rehan M. Towards nanotechnology-based biofuel industry.
Biofuel Res J., 18:798–9, 2018.
22. Nobre BP, Villalobos F, Barragán BE, Oliveira AC, Batista AP, Marques
PASS, Mendes RL, Sovová H, Palavra AF, Gouveia L. A biorefinery from
Nanno­chloropsis sp. microalga-extraction of oils and pigments. Production
of biohydrogen from the left over biomass. Bioresour Technol.,135:128–36,
2013.
23. Palaniappan K. An overview of applications of nanotechnology in biofuel
production. World Appl Sci J., 35:1305–11, 2017.
24. Parker MS, Mock T, Armbrust EV. Genomic insights into marine micro­
algae. Annu Rev Genet., 42:619–645, 2008.
25. Patil V, Tran KQ, Giselrød HR. Towards sustainable production of biofuels
from microalgae. International Journal of Molecular Sciences, 9(7):1188e95,
2008.
26. Prabu A. Nanoparticles as additive in biodiesel on the working characteris-
tics of a DI diesel engine. Ain Shams Eng J., 9:2343–9, 2018.
Microalgae for Biodiesel Production 443

27. Ramos, M.J., Fernández, C.M., Casas, A., Rodríguez, L., Pérez, A., Influence
of fatty acid composition of raw materials on biodiesel properties. Bioresour.
Technol. 100, 261–268, 2009.
28. Roberts GW, Fortier MOP, Sturm BS, Stagg-Williams SM. Promising path-
wayfor algal biofuels through wastewater cultivation and hydrothermal con-
version. Energy & Fuels, 27(2):857–67, 2013.
29. Rodolfi L, Chini Zittelli G, Bassi N, et al. Microalgae for oil: strain selection,
induction of lipid synthesis and outdoor mass cultivation in a low-cost pho-
tobioreactor. Biotechnol Bioeng. 102(1):100–112, 2009.
30. Safarik I, Prochazkova G, Pospiskova K, Branyik T. Magnetically modified
micro­algae and their applications. Crit Rev Biotechnol. 36:931–41, 2016.
31. Sanford, S.D., White, J.M., Shah, P.S., Wee, C., Valverde, M.A., Meier, G.R.,
Feedstock and biodiesel characteristics report. Renewable Energy Group,
Inc., www.regfuel.com (In Press), 2009.
32. Schenk, P. M., Thomas-Hall, S. R., Stephens, E., Marx, U. C., Mussgnug, J. H.,
Posten, C., Kruse, O., and Hankamer, B. Second generation biofuels: high-­
efficiency microalgae for biodiesel production. Bioenergy Resource, 1:20–43,
2008.
33. Sharif HABM, Salleh A, Boyce AN, Chowdhury P, Naqiuddin M. Biodiesel
fuel production from algae as renewable energy. Am J Biochem Biotech,
4(3):250e4, 2008.
34. Singh, A., Nigam, P. S., and Murphy, J. D. Renewable fuels from algae: An answer
to debatable land based fuels. Bioresource Technology, 102(1): 10–16,
2011.
35. Soares, A.T., Costa, D.C., Silva, B.F., Lopes, R.G., Derner, R.B., Antoniosi
Filho, N.R., Comparative analysis of the fatty acid composition of microalgae
obtained by different oil extraction methods and direct biomass transesterifi-
cation. BioEnerg. Res. 73: 1035–1044, 2014.
36. Fu, W., Nelson, D.R., Mystikou, A., Daakour, S., and Salehi-Ashtiani, K.,
Advances in microalgal research and engineering development. Current
Opinion in Biotechnology, 59:157–164, 2019.
37. Xiaoling, M., and Qingyu, W., Biodiesel production from heterotrophic
microalgal oil. Bioresour. Technol. 97, 841–846, 2006.
38. Zaslavskaia LA, Lippmeier JC, Shih C, Ehrhardt D, Grossman AR, Apt KE.
Trophic conversion of an obligate photoautotrophic organism through met-
abolic engineering. Science. 292(5524):2073–2075, 2001.
39. Aljabarin N, Al Jarrah A. Production of biodiesel from local available algae
in Jordan. Journal of Ecological Engineering. 18(6): 8–12, 2017.
40. Mondal M, Goswami S, Ghosh A, Oinam G, Tiwari ON, Das P, Gayen K,
Mandal MK, Halder HN. Production of biodiesel from microalgae through
biological carbon capture: a review. 3 Biotech volume 7, Article number: 99,
2017.
444 Biodiesel Technology and Applications

41. Saad GM, Dosoky NS, Zoromba MS, Shafik HM. Algal Biofuels: Current
Status and Key Challenges. Energies, 12, 1920, 2019.
42. Cuellar-Bermudez SP, Romero-Ogawa MA, Rittmann BE, Parra-Saldivar R.
Algae Biofuels Production Processes, Carbon Dioxide Fixation and Biorefinery
Concept. J Pet Environ Biotechnol, 5: 185, 2014.
43. Beetul K, Sadally SB, Taleb-Hossenkhan N, Bhagooli R, Puchooa D. An
investigation of biodiesel production from microalgae found in Mauritian
waters. Biofuel Research Journal, 2 58–64, 2014.
44. Ihsanullah, Shah S, Ayaz M, Ahmed I, Ali M, Ahmad M, Ahmad I. Production
of Biodiesel from Algae. Journal of pure and Applied Microbiology, March
2015. Vol. 9(1), p. 79–85, 2015.
45. Bošnjaković M. Biodiesel from Algae Journal of Mechanics Engineering and
Automation 3, 179–188, 2013.
46. Hossain S A.B.M., Salleh A. 2008. Biodiesel Fuel Production from Algae as
Renewable Energy. American Journal of Biochemistry and Biotechnology, 4
(3):250–254, 2008.
47. Samorì, C.; Torri, C.; Samorì, G.; Fabbri, D.; Galletti, P.; Guerrini, F.; Pistocchi,
R. & Tagliavini, E. Extraction of hydrocarbons from microalga Botryococcus
braunii with switchable solvents. Bioresource Technology, Vol.101, No.9,
pp. 3274–3279, 2010.
48. Lee, J.; Yoo, C.; Jun, S.; Ahn, C. & Oh, H. Comparison of several methods for
effective lipid extraction from microalgae. Bioresource Technology, Vol.101,
pp. 575–577, 2010.
49. Umdu, E.S.; Tuncer, M. & Seker, E. Transesterification of Nannochloropsis
oculata microalga’s lipid to biodiesel on Al2O3 supported CaO and MgO
catalysts. Bioresource Technology, Vol.100, No.11, (6), pp. 2828-2831, 2009.
50. Cooney, M.; Young, G. & Nagle, N. Extraction of bio-oils from microalgae.
Separation & Purification Reviews, Vol.38, pp. 291–325, 2009.
51. Lardon, L.; Hélias, A.; Sialve, B.; Steyer, J. & Bernard, O. Life-cycle assess-
ment of biodiesel production from microalgae. Environmental Science and
Technology, Vol.43, No.17, pp. 6475–6481, 2009.
52. Yaguchi, T.; Tanaka, S.; Yokochi, T.; Nakahara, T. & Higashihara, T.
Production of high yields of docosahexaenoic acid by Schizochytrium sp.
strain SR21. JAOCS, Journal of the American Oil Chemists’ Society, Vol.74,
No.11, pp. 1431–1434, 1997.
53. Halim, R.; Gladman, B.; Danquah, M.K. & Webley, P.A. Oil extraction from
microalgae for biodiesel production. Bioresource Technology, Vol.102, No.1,
pp. 178–185, 2010.
54. Andrich, G.; Nesti, U.; Venturi, F.; Zinnai, A. & Fiorentini, R. Supercritical
fluid extraction of bioactive lipids from the microalga Nannochloropsis sp.
European Journal of Lipid Science and Technology, Vol.107, No.6, pp. 381–
386, 2005.
Microalgae for Biodiesel Production 445

55. Perrut, M. Supercritical fluid applications: Industrial developments and eco-


nomic issues. Industrial and Engineering Chemistry Research, Vol.39, No.12,
pp. 4531–4535, 2000.
56. Cooney, M.; Young, G. & Nagle, N. Extraction of bio-oils from microalgae.
Separation & Purification Reviews, Vol.38, pp. 291–325, 2009.
57. Fu, C.; Hung, T.; Chen, J.; Su, C. & Wu, W. Hydrolysis of microalgae cell
walls for production of reducing sugar and lipid extraction. Bioresource
Technology, Vol.101, No.22, (November 2010), pp. 8750–8754, 2010.
58. Altin, R., Cetinkaya, S., Yucesu, H.S. The potential of using vegetable oil fuels
as fuel for Diesel engines. Energy Convers. Manage., 42, 529–538, 2001.
59. Basha, S.A.,Gopal, K.R., Jebaraj, S. A review on biodiesel production, com-
bustion, emissions and performance. Renewable and Sustainable Energy
Reviews 13,1628–1634, 2009.
60. De Bhowmick, G., Koduru, L., Sen, R. Metabolic pathway engineering
towards enhancing microalgal lipid biosynthesis for biofuel application—a
review. Renew Sust Energ Rev, 50:1239–1253, 2015.
61. Gouveia, L., Marques, A.E., Sausa, J.M., Bandarra, N.M. Microalgae—source
of natural bioactive molecules as functional ingredients. Food Sci Technol
Bull: Funct Foods, 7(2):21–37, 2010.
62. Johnson, M.B. and Wen, Z. Production of Biodiesel Fuel from the Microalgal
Schizochytrium limacinum by Direct Transesterification of Algal-Biomass.
Energy Fuels, 23,10: 5179–5183, 2009.
63. Zienkiewicz,K. Du,Z., Ma, W.,Vollheyde,K., Benning, C. Stress-induced neu-
tral lipid biosynthesis in microalgae – Molecular, cellular and physiological
insights, BBA - Molecular and Cell Biology of Lipids, 2016.
17
Biodiesel Production Methods
and Feedstocks
Setareh Heidari1* and David A. Wood2
1
Faculty of Chemical Engineering, Sahand University of Technology,
Tabriz, Iran
2
DWA Energy Limited, Bassingham, Lincoln, United Kingdom

Abstract
Biodiesel is an attractive alternative transportation fuel to crude oil–derived diesel
because it is sustainable and can be produced in almost carbon-neutral supply
chains that avoid the accumulation of greenhouse gases in the atmosphere. There
are three generations of biofuel production options: involving food crops on agri-
cultural-quality land, involving non-food crops and low-grade land, and involving
algae (naturally occurring). A fourth-generation option is to genetically engineer
algal to generate higher lipid yields. Also, waste oils, such as cooking oils, are also
a common feedstock for biodiesel production. There are several methods used to
produce biofuels including the transformation of lipids into microemulsions by
mixing with alcohols, pyrolysis of solid biomass, including woody materials and
transesterification. The latter is the most widely used for large-scale biodiesel pro-
duction. Catalyst selection, plant operating conditions, and the alcohol:plant lipid
ratio influence the conversion efficiency of transesterification. Overcoming metal
corrosion issues and achieving commercial price parity with crude oil–derived
diesel are two key challenges the biodiesel industry currently faces.

Keywords: Transportation fuel, pollution control, biodiesel, waste oils,


transesterification, enzymatic catalysis, nanocatalysts, alcohol:lipid ratio

*Corresponding author: setarehaidari@yahoo.com

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (447–464) © 2021 Scrivener Publishing LLC

447
448 Biodiesel Technology and Applications

17.1 Introduction
In recent decades, there have been some concerns about fuel supply and
fuel security of supply globally. A realistic expectation is to increase con-
sumer confidence in emerging biofuel technologies to meet growing
demands for sustainable fuel supplies in the next decades. Unfortunately,
around the world, the primary energy and fuel sources are mostly provided
for by fossil fuels, with their negative consequences for the environment
and violation of sustainability development obligations. Crude oil–refined
products such as gasoline, diesel, and aviation fuel dominate transporta-
tion fuels. However, there is a small contribution from biofuels, mainly
ethanol, with some biodiesel produced and/or consumed in many coun-
tries. Biofuels provide a substantial opportunity to deliver a supply of road
transportation fuels that is sustainable into the far future and limits the
greenhouse gases emitted to the atmosphere and reduces pollution more
generally during its consumption. Biodiesel shows many attractive charac-
teristics that are considered in detail in this chapter.
Today, there is a broad range of fluids, like water [1], solvents, fossil
energy [2–5], supercritical carbon dioxide [6–9], and biofuels that can be
utilized to promote the world’s economy effectively. The exploitation of
these substantial sources must be administered efficiently and their sup-
ply chains developed in ways that are efficient and sustainable in the long
term. Biodiesel is generated from waste oils, vegetable oils, and/or fatty
compounds extracted from animal tissue, more generally referred to as
lipids. It is valued as a biofuel because it can be produced from a broad
range of feedstocks. It produces diesel-fuel specifications including a blend
dominated by straight-chain alkanes with number carbon of 16 to 20 other
hydrocarbon molecules such as the functional alkyl group and organic
esters [10, 11]. These components of biodiesel are mostly generated as a
consequence of the transformation reactions imposed on the lipids to form
alcohols and esters.Throughout the 20th century, vegetable oils were used
in small quantities to form biodiesel for road vehicle utilization. In the
1920s, vehicle engines were adapted to consume cheaper, crude oil–based
diesel fuels provided from refineries [10]. Biodiesel received significant
attention on a mercantile scale in South Africa before World War II to sup-
ply a road vehicle fuel for heavy-duty vehicles; however, it was rarely used
at that time in other countries. Years later, in the 1970s, the oil tensions
caused a fuel deficiency and a commercial motive to produce at least some
biodiesel, and bioethanol, as alternatives to fossil-fuel-based diesel. There
are various factors, such as political policy, commercial viability, subsidies
and incentives for sustainable products, and environmental factors that
Biodiesel Production Methods and Feedstocks 449

affect governments’ and industries’ motivations and attraction to biodiesel.


Typically, it has had to be justified, either on commercial or sustainability
grounds, to compete as a substitute for traditional diesel in miscellaneous
countries for the past two to three decades [12]. Today, there is a wide
range of countries that are keen to generate biodiesel in Asia, Europe, and
North America and are striving to reduce its cost of supply and improve
the efficiencies and capacities of the processes used to generate it. As a
rule of thumb, a typical plant is able to produce some 6 million liters (~1.5
million gallons), or more, specification-grade biodiesel fuel annually [13].

17.2 Biofuel Classification in Terms of Origin


and Technological Conversion of Raw Materials
Biofuels refer to the fuel types derived from biological/organic feedstocks
[14–16]. When the feedstocks are processed, they are transformed into
renewable and, in most cases, sustainable, carbon-based energy fuels. These
fuels can be used commercially, including ethanol and biodiesel for numer-
ous applications, but most of them end up as blendstocks for transporta-
tion usage [17]. Although ethanol, specifically that blended into refinery
gasoline, dominates the global biofuel markets, biodiesel is a substantial
and growing contributor. Generally, biofuels can be subdivided into first to
third generations, depending on the type of feedstocks involved.
The first-generation biofuel production method is the most widely used
technique to directly produce fuel from food crops (e.g., soybean, palm oil,
sugarcane, sugarcane, and sunflower) in the form of ethanol and biodiesel.
First-generation biofuels have clear benefits including producing valuable
fuel supply, burning cleaner with less emissions, little or no toxicity, and
sustainability compared to fossil fuels. However, when comparing the
impact of such production on the prices of the raw materials, particularly
cereal crops, and the plant operating costs and costs of supply, it is hard to
justify some biofuel supply chains to divert invaluable and comparatively
high-cost food crops to the biofuel industry [17–21]. Moreover, using palm
oil as a feedstock for biodiesel is not justifiable on environmental grounds,
if pristine rain forest is destroyed to give way to oil farm plantations, even
though palm oil delivers greater biodiesel yield than most other first-
generation feedstocks.
Second-generation biofuels, also known as an advanced biofuel, are
made from non-food crops or crops grown on very low-grade non-
agricultural land by generally more complex and costly methods [22, 23].
There are three main sources of second-generation biofuels production
450 Biodiesel Technology and Applications

including non-edible vegetable oils (e.g., cotton seed, milkweed, linseed,


rubber seed, and mahua), waste oils, and animal fat feedstocks. The main
advantages of second-generation biofuels are their positive sustainability
and environmental credentials, and potential economic profitability and
competitiveness at lower crude oil prices. Feedstocks used in producing
second-generation biofuels tend to be considerably cheaper than food
crop–derived vegetable oils. Also, they are beneficial as they significantly
diminish the cost of waste disposal and treatment. However, yields and
conversion efficiency from fast-growing lower-grade plants, e.g., Jatropha
curcas (with a natural oil content of about 37%, need to be improved to
further justify their development on low-grade semi-arid land.
Nevertheless, second-generation biofuels offer substantial potential to
reduce the planet’s dependence on crude oil–derived transportation fuels.
However, it is unlikely to grow medium-term demand for transportation
fuels around the world as they are still produced in relatively small quan-
tities, particularly biodiesel [20, 24–27]. Besides, when it comes to using
animal fat feedstock, the major barrier is the existence of a high amount
of free fatty acids (FFA) in conjunction with higher water content, which
tends to diminish the biofuel production yield [28–34].
Some countries, such as the United Kingdom, process a lot of waste oil,
particularly used cooking oils, in some cases blended with waste engine
oils (of fossil fuel origin), to produce biodiesel in a cost effective way. The
acquisition costs for waste oils are very low as most producers of large-
scale waste oils avoid disposal costs by “gifting” them to biofuels producers.
Most of the supply chain costs for the biodiesel producers are in the gath-
ering, transportation, and storage costs (handling logistics) collecting indi-
vidual small quantities of oil from many suppliers located in distal parts
of a catchment area. Additionally, because waste oils tend to be of vari-
ous compositions with variable contaminants within them, biodiesel pro-
ducers from waste oils have additional operating costs related to ensuring
specification grade fuel products and disposal costs of some contaminants
removed from the feedstocks.
The third-generation biofuels use various forms of algal biomass to gen-
erate biofuels. Although algal species may need relatively large volumes
of water, nitrogen, and phosphorus to grow especially at industrial scale,
certain species possess outstanding oil yields. Also, they have the poten-
tial to improve biodiesel production capacity and efficiency and reduce
the environmental footprint of individual biodiesel plants. Such plants
could be located on low-quality land and close to carbon dioxide emitters
(e.g., power plants and sewage processing plants) that could, in the future,
provide algal plants a regular supply of captured carbon dioxide. High
Biodiesel Production Methods and Feedstocks 451

rapeseed oil
1% 1% 1%

13%
sunflower seed oil

soybean oil

84%
palm oil

other oils

Figure 17.1 Primary feedstocks used to produce biodiesel [37]. Other oils include those
derived from waste cooking oils, jatropha and tallow.

hydrolysis and/or fermentation efficiency, no need for herbicides or pes-


ticides in algal cultivation, relatively high resistance even in harsh condi-
tions (e.g., saline and brackish water) are some of the exceptional features
make algal biomass an attractive candidate for a new generation of biofuel
production. Even more appealing, though, is the possibility of genetically
modifying algae, also known as fourth-generation biofuels (FGBs), to fur-
ther increase the productivity of customized algal species. It seems very
likely that FGB will help to open new more-commercially viable options
for biodiesel production in the future [17, 35].
Despite introducing a few commercial-scale second-generation bio-
diesel production techniques, today, the majority of the world’s total bio-
diesel (above 95%) is synthesized from edible oils mostly from rape seed,
accounting for approximately 85% [36] (see Figure 17.1).

17.3 Techniques Capable of Producing Biodiesel


on Commercial Scales
Biodiesel may be produced via various techniques [38–41]. These include
the following:

• direct use of plant-derived and waste oils;


• blending of plant-derived oils and oil wastes with other fuels
(e.g., refinery diesel);
• microemulsion with excess methanol and ethanol to
decrease viscosity;
452 Biodiesel Technology and Applications

• thermal cracking to break down longer chain molecules;


• pyrolysis of oxygen-rich organic feedstock of algal, plant, or
animal origin; and,
• transesterification.

The most commonly used method in commercial biodiesel plants is


transesterification [40].

17.3.1 Direct and Blending Methods With the Aim


of Biodiesel Generation
There are some harmful materials in crude oil–derived diesel fuel, par-
ticularly the aromatic molecules xylenes, toluene, and benzene, and some
particulates. Diesel from refineries also consists of contaminants like sul-
fur and particles of heavy oil residues which are harming people’s health
and the atmosphere [42, 43]. On the other hand, the chemical components
of biodiesel are much cleaner and purer than those found in crude oil–
derived diesel due to the lack of such contaminants in the biological feed-
stocks [37, 44]. The hydrocarbon molecules in biodiesel are mostly straight
chains with carbon numbers ranging from 16 to 20, not aromatics. The
lack of sulfur in biodiesel decreases the accumulation of sulfuric acid in an
engine’s crankcase oil and subsequently, makes it less corrosive to engines
compared to the crude oil–derived diesel [45]. On the negative side, the
emissions of nitrogen oxides (NOx) can be higher with biodiesel than
refinery diesel [46, 47]. By weight, there is about 10% oxygen in biodiesel.
However, during the consumption of biodiesel in an engine, fewer hydro-
carbons, particles, carbon monoxide, and sulfur oxides (SOx) are released
than with crude oil–derived diesel [37].

17.3.2 Microemulsion Methods


Oil derivatives of vegetable tend to result in products with higher vis-
cosities, unless treated, leading to some biodiesel processing issues. To
overcome this obstacle, microemulsions with solvents [1, 48] such as eth-
anol and methanol can be utilized to effectively reduce product viscosities.
Microemulsions are defined as isotropic colloidal fluid dispersions utiliz-
ing particles with diameter of 1 to 150 nm. They are commonly generated
from two liquids which in standard conditions are mostly immiscible, i.e.,
a) ionic and b) non-ionic amphiphiles. Microemulsions typically promote the
Biodiesel Production Methods and Feedstocks 453

spray characteristics of the oils because they vaporize immediately because


they contain components with lower boiling points within their micelles.

17.3.3 Pyrolysis Methods


Pyrolysis is the reaction in which primarily solid biomass is degraded by
applying high temperatures at different heating rates. This is mostly attained
through enhancing the temperature of the vegetable oils in the absence of
oxygen. This generates bio-oils (liquids) along with co-generation products
of charcoal (solid) and fuel gas. Biomass pyrolysis methods mostly produce
and recover biofuels with medium or low calorific amounts/energy con-
tents [49]. Based on the process conditions, biofuel pyrolysis methods are
divided into the “ordinary”, “quick”, and “flash” methods as follows:
a) Ordinary pyrolysis involves reactions which are stimulated by rela-
tively slow warming rates ranging from 0.1°K/s to 1°K/s over a residence-
time interval of 45 to 550 seconds. Initially, a pre-pyrolysis step is required.
This is typically conducted in the temperature range of 550–950°K, stim-
ulating biomass decomposition to occur. In this step, molecular restruc-
turing arises, such as water elimination, cracking of some chemical bonds,
free-radical liberation, creation of carbonyl functional groups, and, in
some cases, hydroperoxide functional groups [50].
b) Quick pyrolysis is implemented in a temperature range of 850°K to
1250°K exerting heating rates from 10°K/s to 200°K/s over a residence-time
interval of 0.5–10 s. Feedstocks that are fragmented into the very small
particle, i.e., less than 1 mm, are commonly utilized. This involves pre-
treatment preparations of most solid biomass feedstocks.
c) Flash pyrolysis is commonly applied to ground fragments of relatively
large chunks of woody materials as its biomass feedstock. It is accomplished
in a temperature range from 1,050°K to 1,300°K at high rates of heating
attaining more than 1,000°K/s. Residence-time intervals of less than 0.5 s
and particle sizes of <0.2 mm are typically involved.

17.3.4 Transesterification Methods


The other name for the transesterification method is “alcoholysis”.
Vegetable oils which are formed principally of lipids are mixed with alco-
hols, in which they are soluble, and react to generate esters and glycerin.
This is commonly directed in the presence of a catalyst to amend the effi-
ciencies and reaction rates [51]. Methanol or ethanol are typically chosen
from among a broad range of potential alcohols that could be utilized,
454 Biodiesel Technology and Applications

because they possess highly suitable chemical and physical properties and
are of low cost to procure [52].

17.4 Influential Parameters on Biodiesel Production


Fundamentally, the most influential parameters affecting biodiesel produc-
tion are the nature of feedstocks or raw materials. Additionally, the catalyst
type and the concentration of that catalyst, the alcohol:lipid ratio, reaction
temperature, and time are also key influencing factors [53].

17.4.1 The Choice of Transesterification Catalysts


The process of biodiesel synthesis is widely performed using the transes-
terification reaction, catalyzed by homogeneous and/or heterogeneous
catalysts [54]. Generally, homogeneous catalysts can be divided into two
types, namely, alkaline catalysts (e.g., NaOH, KOH, and CH3OK) and acid
catalysts (e.g., H2SO4, HCl, and Fe2(SO4)3). The main advantage of homo-
geneous catalysis is that they are capable of dissolving in the reaction mix-
ture, facilitating high interactions between catalyst and reactants. Despite
the wide application of homogeneous catalysts, in transesterifing feedstocks,
they suffer from process difficulties and very high costs, associated with the
requirement to pretreat low-quality feedstocks [54–60]. Thus, the develop-
ment of novel and more efficient catalysts to overcome the aforesaid issues is
of great importance. In this regard, heterogeneous catalyst, including solid
acid (e.g., SiO2/Al2O3 and ion-exchange resins), solid alkaline (e.g., CaO,
LiOH/g-Al2O3, ZnO, BaO, and MgO), enzymatic catalysis (e.g., Candida
Antarctica, a key source of lipases, and Pseudomonas cepacia), and nano-
material-based catalysts (e.g., nanostructured mixed-metal oxides of CaO-
MgO) have drawn considerable attention owing to their excellent results
from a wide range of potential biomass conversion applications tested [54,
59, 61–77]. Of particular appeal is the possibility of tuning the structure of
nanocatalysts, by initiating modification to their functional groups, in order
to maximize activity, selectivity, and stability. Such approaches result in opti-
mizing the catalyst reactions and, consequently, boosting biofuel yield [61].

17.4.2 Effects of Catalyst Characteristics on Biodiesel


Production Efficiency
Variations of catalyst type, changes in catalyst concentrations, reac-
tion times, and the ratios of alcohol/plant oil (lipids) all influence the
Biodiesel Production Methods and Feedstocks 455

(a) 100 120 (b)

95 100
80

Yield (%)
Yield (%)

90
60
85 40

80 20
0
75 0 200 400 600
NaOH KOH CH30Na CH3OK Time (min)
(c) (d)
100 100
98
80
96
Yield (%)
Yield (%)

60 94
40 92
20 90
88
0 4:01 5:01 6:01 7:01 8:01
0.2 0.4 0.6 0.8 1 1.2 115 Methanol/Oil (molar ratio)
Catalyst Concentration

Figure 17.2 (a) Impact of catalyst type on esters yield for four different alkaline catalysts
to produce biodiesel product from sunflower and rape seed lipids (T = 65°C; methanol:oil
molar ratio is 6:1; catalyst 1 wt.% of plant lipids). (b) The effect of SnCl2/Nb2O5 catalyst
extraction time on the ethyl oleate yield (T= 60°C; 50 mg of 50% wt. SnCl2/Nb2O5) [78].
(c) the effect of catalyst concentration on biodiesel yield from coconut oil (T = 60°C; 350
rpm; 60-min reaction time; NaOH catalyst in methanol) [79]. (d) The effect of methanol
in the range of 4:1–8:1 molar ratio (T = 65 ± 1°C; reaction time = 1 (h) in the analyses of
Jatropha oil [80].

performance of biodiesel production techniques. These impacts are illus-


trated in Figure 17.2. Accordingly, substantial testing efforts are required to
determine the optimum reaction time, catalyst concentration, and relative
concentration of alcohol to plant oil to maximize the biodiesel yield.

17.5 Biodiesel Markets and Economic Considerations


Currently, the expansion of novel biodiesel technologies is affected by eco-
logical and economic obstacles which are attracting a high level of interna-
tional concern and attention. Carbon dioxide can be exploited beneficially
by some recent technologies [81, 82]; however, when emitted to the atmo-
sphere, it has harmful consequences for the environment. The “green-
house effect” is causing the atmosphere to warm due to the anthropogenic
456 Biodiesel Technology and Applications

emissions of carbon dioxide and other gases, thereby degrading the Earth’s
atmosphere, ecosystems, and surface environment. Various multi-national
accords and organizations are analyzing the impacts and seeking ways to
mitigate their detrimental effects on the atmosphere and the surface envi-
ronment. The further substitution of crude oil–derived transportation
fuels with biodiesel is one possible beneficial solution to address the cur-
rent climate emergency. Several directives issued by the European Union
support the increased uptake of biodiesel and encourage the more wide-
spread production and consumption of biodiesel as an alternative road
vehicle fuel, although its original targets of 20% biofuel contribution by
2020 [83] have not been achieved. Nevertheless, most of Europe’s large
road vehicles are diesel-powered and, consequently, there is a substantial
opportunity for biodiesel generated from plant lipid, waste oil, and algae
to expand. Ultimately, it will be government policies, fuel tax strategies,
and the cost of producing biodiesel compared to the cost of importing and
refining crude oil–derived diesel in Europe that will determine the uptake
of this environmentally friendly transportation fuel [10, 84].

17.6 Challenges Confronting Biodiesel Uptake


About 15 years ago, the world was faced with an energy shortage due, at
that time, to the decline of developed crude oil production sources. That
situation has been reversed over the past 15 years by the rapid expansion of
the United States shale industry leading to a massive oversupply of crude oil
and products globally in 2019 and 2020. However, the past 25 years has also
seen a growing awareness of the pending global climate catastrophe. With
such awareness, regardless of crude oil prices and abundant short-term
supply, countries continue to look for renewable fuel substitutes for crude
oil, which would be environment-friendly and sustainable. Rapid oil price
falls in 2014 (oversupply due to shale oil production in the United States),
2019 (exports of oil from United States flooding global markets), and 2020
(COVID-19 pandemic dramatically reducing global energy demand) have
made fossil fuels in the short-term substantially more cost effective than
biofuels. This situation, most likely, will adversely slow down the uptake of
biofuels around the world for the next few years. However, the prevailing
situation should also stimulate more research to improve biodiesel pro-
duction processes and make them cost competitive with crude oil–derived
diesel at oil prices of less than $40/barrel.
There are several energy resources of biological origin such as alco-
hols (butanol, ethanol, and methanol), plant and algal oils/lipids, woody
Biodiesel Production Methods and Feedstocks 457

biomass, agricultural organic wastes and biogases, other waste oils, and
various synthetic fuels, which could all be usefully exploited to achieve
the objectives of substituting for fossil fuels. Biodiesel produced from such
sources, in particular, offers the opportunity to be blended with refinery-
derived diesel. Blending means that it can be used in existing vehicles, at
specifications close to refinery grade diesels, with no engine modifications
required [85]. Today, there are several countries like the USA, the EU
nations, and Brazil where government mandates exist to ensure that road
fuels contain minimum quantities of biofuels, mainly ethanol in gasoline,
but also biodiesel into refinery diesel. Such mandates have underpinned
the economics of the biofuels industry for more than 15 years and continue
to do so.

17.7 Corrosion and Quality Monitoring Issues


for Biodiesel
Corrosion of metals is of concern for the long-term storage of refinery
diesel and, particularly biodiesel, mainly due to its acidity. This impacts
storage facilities, including vehicle fuel tanks, and the quality of the fuel
itself with traces of metal potentially damaging vehicle engines [86–89].
This means that diesel vehicle engines that use large quantities of bio-
diesel require continuous checks for corrosion impacts. Further studies
are required to investigate traces of metals in biofuel and their impacts on
the oxidation stability of biodiesel. Developing effective antioxidants and
corrosion inhibitors for biodiesel tanks and engines is also required to mit-
igate or avoid metal corrosion.
Precise, fast, and readily available on-site analysis to verify blended die-
sel fuel specification is being met. These are typically required at vehicle
refueling stations and fuel storage terminals which handle large volumes
of biofuels [90]. This is particularly so for biodiesel, because it is manufac-
tured from a diverse range of feedstocks of quite disparate compositions.
This means that it is prone to contain small traces of unwanted contami-
nants if quality control during its production, blending, and storage is not
rigorously conducted.

17.8 Conclusions
Biodiesel is a proven, effective, and sustainable transportation fuel that can,
and does, substitute for refinery diesel in many supply chains. Biofuels can
458 Biodiesel Technology and Applications

be subdivided into first to third generations depending on the type of feed-


stocks involved; biofuels produced from food crops are first generation;
biofuels produced from non-food crops and/or on low grade land are sec-
ond generation; biofuels produced from algal feedstocks are third genera-
tion; and some refer to biofuels derived from genetically engineered algae
as fourth generation. Biofuels generated from waste oils (e.g., cooking oil)
are generally considered as second-generation products. This means that
there is a diverse range of feedstocks available for biofuel production.
Some plant-derived oils and waste oils can be used directly or in blended
form as vehicle fuels. Others, with little processing and low-cost treatments
and blending, convert viscous oils into low viscosity microemulsions with
methanol or ethanol render other biolipids ready to consume more effi-
ciently in vehicle engines. For solid biomass, including woody materials,
various pyrolysis methods, sometimes combined with gasification are
required to produce biofuels without leaving substantial solid residues.
The most common and efficient techniques that operate in larger-scale
commercial biodiesel production plants involve transesterification. These
produce biodiesel by mixing plant lipids with alcohols and catalysts.
There are some potential metal corrosion issues in engines and storage
facilities and fuel specification limits that need to be carefully adhered to.
These require careful monitoring, and in the case of corrosion mitigation
by corrosion inhibitors. Expansion of global biodiesel demand, consump-
tion, and production capacities in the long-term still depends on some
technology breakthroughs to increase product yields and reduce operating
costs. In the short term, the COVID-19 pandemic and the global recession
it has induced in 2020 is expected to negatively impact transportation fuel
demand around the world for the next year or so. This is also likely to
hinder investment and growth in this promising sector over the next few
years.

References
1. Sadatshojaei; E., Wood; D. A., Water, The Most Accessible Green Solvent,
Extraction and Separation Agent, in: Green Sustainable Process for Chemical
and Environmental Engineering and Science: Environmental Remediation
Elsevier, 2020.
2. E. Sadatshojaei, M. Jamialahmadi, F. Esmaeilzadeh, M.H. Ghazanfari, Effects
of low-salinity water coupled with silica nanoparticles on wettability alter-
ation of dolomite at reservoir temperature, Petroleum Science and Technology,
34, 1345–1351, 2016.
Biodiesel Production Methods and Feedstocks 459

3. E. Sadatshojaei, M. Jamialahmadi, F. Esmaeilzadeh, D.A. Wood, M.H.


Ghazanfari, The impacts of silica nanoparticles coupled with low-salinity
water on wettability and interfacial tension: Experiments on a carbonate
core, Journal of Dispersion Science and Technology, 1–15, 2019.
4. A. Choubineh, H. Ghorbani, D.A. Wood, S.R. Moosavi, E. Khalafi, E.
Sadatshojaei, Improved predictions of wellhead choke liquid critical-flow
rates: modelling based on hybrid neural network training learning based
optimization, Fuel, 207, 547–560, 2017.
5. M.S. Mirqasemi, M. Homayoonfal, M. Rezakazemi, Zeolitic imidazolate
framework membranes for gas and water purification, Environmental
Chemistry Letters, 18, 1–52, 2020.
6. M. Rezakazemi, O. Tavakoli, Hydrothermal Decomposition of Strongly
Acidic Cation-Exchange Resin to Valuable Compounds Using Subcritical
Water in Alkaline Media, ChemistrySelect, 5, 3257–3265, 2020.
7. M. Rezakazemi, M. Sadrzadeh, T. Matsuura, Thermally stable polymers for
advanced high-performance gas separation membranes, Progress in Energy
and Combustion Science, 66, 1–41, 2018.
8. N. Ramezani, F. Raji, M. Rezakazemi, M. Younas, Juglone extraction from
walnut (Juglans regia L.) green husk by supercritical CO2: Process optimiza-
tion using Taguchi method, Journal of Environmental Chemical Engineering,
8, 103776, 2020.
9. M. Younas, M. Rezakazemi, M. Daud, M.B. Wazir, S. Ahmad, N. Ullah,
Inamuddin, S. Ramakrishna4, Recent progress and remaining challenges in
post-combustion CO2 capture using Metal-Organic Frameworks (MOFs),
Progress in Energy and Combustion Science, 2020.
10. M. Ahmad, M.A. Khan, M. Zafar, S. Sultana, Practical handbook on bio-
diesel production and properties, CRC Press, 2012.
11. N. Hajilary, M. Rezakazemi, S. Shirazian, Biofuel types and membrane sepa-
ration, Environmental Chemistry Letters, 17, 1–18, 2019.
12. K.P. McDonnell, M. W. Shane, P.B. Mcnulty., Result of engine and vehicle
testing of semi-refied rapeseed oil, in: 10th International Rapeseed Congress,
Canberra, Australia., 1999.
13. J.H. Van Gerpen, C.L. Peterson, C.E. Goering, Biodiesel: an alternative
fuel for compression ignition engines, American Society of Agricultural and
Biological Engineers, 2007.
14. Sadatshojaei; E., Wood; D. A., Mowla; D., Third generation of biofuels
exploiting microalgae, in: Sustainable Green Chemical Process and their
Allied Applications, Springer Nature, Switzerland AG, 2020.
15. Sadatshojaei; E., Samiei Nezhad; M., Wood; D. A., Rahimpour; M.R.,
Progress towards the sustainable generation of biofuels exploiting microal-
gae, in: Biodiesel Technology and Applications, Wiley-Scrivener, USA., 2020.
16. Sadatshojaei; E., Mowla; D., Wood; D. A., Review of Progress in Microalgal
Biotechnology Applied to Wastewater Treatment, in: Sustainable Green
Chemical Process and their Allied Applications, Springer Nature, 2020.
460 Biodiesel Technology and Applications

17. R.A. Lee, J.-M. Lavoie, From first-to third-generation biofuels: Challenges
of producing a commodity from a biomass of increasing complexity, Animal
Frontiers, 3, 6–11, 2013.
18. G. Pérez-Lechuga, S.L. Aguilar-Velázquez, M.A. Cisneros-López, F.V.
Martínez, A model for the location and scheduling of the operation of
­second-generation ethanol biorefineries, Journal of Mathematics in Industry,
9, 3, 2019.
19. M. Kannahi, R. Arulmozhi, Production of biodiesel from edible and non-­
edible oils using Rhizopus oryzae and Aspergillus niger, Asian Journal of
Plant Science and Research, 3, 60–64, 2013.
20. A. Demirbas, A. Bafail, W. Ahmad, M. Sheikh, Biodiesel production from
non-edible plant oils, Energy Exploration & Exploitation, 34, 290–318, 2016.
21. R. Luque, J.A. Melero, Advances in biodiesel production: Processes and
Technologies, Elsevier, 2012.
22. A. Dahiya, Bioenergy: biomass to biofuels, Academic Press, 2014.
23. J. Bundschuh, G. Chen, Sustainable energy solutions in agriculture, CRC
Press, 2014.
24. M.M. Gui, K. Lee, S. Bhatia, Feasibility of edible oil vs. non-edible oil vs.
waste edible oil as biodiesel feedstock, Energy, 33, 1646–1653, 2008.
25. A. Talebian-Kiakalaieh, N.A.S. Amin, H. Mazaheri, A review on novel pro-
cesses of biodiesel production from waste cooking oil, Applied Energy, 104,
683–710, 2013.
26. C.D.M. de Araújo, C.C. de Andrade, E.d.S. e Silva, F.A. Dupas, Biodiesel pro-
duction from used cooking oil: A review, Renewable and Sustainable Energy
Reviews, 27, 445–452, 2013.
27. A. Gnanaprakasam, V.M. Sivakumar, A. Surendhar, M. Thirumarimurugan,
T. Kannadasan, Recent strategy of biodiesel production from waste cook-
ing oil and process influencing parameters: a review, Journal of Energy, 2013
2013.
28. C. Öner, Ş. Altun, Biodiesel production from inedible animal tallow and an
experimental investigation of its use as alternative fuel in a direct injection
diesel engine, Applied Energy, 86, 2114–2120, 2009.
29. J. Lu, K. Nie, F. Xie, F. Wang, T. Tan, Enzymatic synthesis of fatty acid methyl
esters from lard with immobilized Candida sp. 99–125, Process Biochemistry,
42, 1367–1370, 2007.
30. H. Zhao, Z. Lu, X. Bie, F. Lu, Z. Liu, Lipase catalyzed acidolysis of lard with
capric acid in organic solvent, Journal of Food Engineering, 78, 41–46, 2007.
31. M. Gürü, A. Koca, Ö. Can, C. Çınar, F. Şahin, Biodiesel production from
waste chicken fat based sources and evaluation with Mg based additive in a
diesel engine, Renewable Energy, 35, 637–643, 2010.
32. M. Balat, Potential alternatives to edible oils for biodiesel production–A
review of current work, Energy Conversion and Management, 52, 1479–1492,
2011.
Biodiesel Production Methods and Feedstocks 461

33. S. Pinzi, Feedstocks for advanced biodiesel production, in: Advances in


Biodiesel Production, Elsevier, pp. 69–90, 2012.
34. V. Feddern, A. Cunha Junior, M. Pra, P.G. de Abreu, J.I. dos Santos Filho,
M.M. Higarashi, M. Sulenta, A. Coldebella, Animal fat wastes for biodiesel
production, Embrapa Suínos e Aves-Capítulo em livro científico (ALICE),
2011.
35. L. Brennan, P. Owende, Biofuels from microalgae—a review of technologies
for production, processing, and extractions of biofuels and co-products,
Renewable and Sustainable Energy Reviews, 14, 557–577, 2010.
36. V.K. Mishra, R. Goswami, A review of production, properties and advan-
tages of biodiesel, Biofuels, 9, 273–289, 2018.
37. P. Arkoudeas, S. Kalligeros, F. Zannikos, G. Anastopoulos, D. Karonis, D.
Korres, E. Lois, Study of using JP-8 aviation fuel and biodiesel in CI engines,
Energy Conversion and Management, 44, 1013–1025, 2003.
38. A. Dimitriadis, I. Natsios, A. Dimaratos, D. Katsaounis, Z. Samaras, S.
Bezergianni, K. Lehto, Evaluation of a hydrotreated vegetable oil (HVO) and
effects on emissions of a passenger car diesel engine, Frontiers in Mechanical
Engineering, 4, 7, 2018.
39. Y. Sharma, B. Singh, Development of biodiesel: current scenario, Renewable
and Sustainable Energy Reviews, 13, 1646–1651, 2009.
40. M.N. Varma, G. Madras, Synthesis of biodiesel from castor oil and linseed oil
in supercritical fluids, Industrial & Engineering Chemistry Research, 46, 1–6,
2007.
41. S.N. Gebremariam, J.M. Marchetti, Biodiesel production technologies, AIMS
Energy, 5, 425–457, 2017.
42. D.L. Klass, Biomass for renewable energy, fuels, and chemicals, Elsevier,
1998.
43. A. Keskin, D. Altıparmak, M. Gürü, Usability of sunflower oil methyl ester
as an alternative fuel in diesel engines, in: Proceedings of the eighth interna-
tional combustion symposium, Ankara, Turkey, pp. 647–652, 2004.
44. S. Kalligeros, F. Zannikos, S. Stournas, E. Lois, G. Anastopoulos, C. Teas, F.
Sakellaropoulos, An investigation of using biodiesel/marine diesel blends on
the performance of a stationary diesel engine, Biomass and Bioenergy, 24,
141–149, 2003.
45. K.S. Tyson, R. McCormick, Biodiesel Handling and Use Guidelines, National
Renewable Energy Laboratory, NREL Report, (2001).
46. J. Van Gerpen, Biodiesel processing and production, Fuel Processing Tech­
nology, 86, 1097–1107, 2005.
47. M. Cetinkaya, Y. Ulusoy, Y. Tekìn, F. Karaosmanoğlu, Engine and winter
road test performances of used cooking oil originated biodiesel, Energy
Conversion and Management, 46, 1279–1291, 2005.
48. Heidari; S., Sadatshojaei; E., Wood; D. A., Large-scale Molecular Solvents
for Environmentally Sustainable Applications, in: Green Sustainable Process
462 Biodiesel Technology and Applications

for Chemical and Environmental Engineering and Science: Environmental


Remediation, Elsevier, 2020.
49. A. Demirbas, Current technologies for the thermo-conversion of biomass
into fuels and chemicals, Energy Sources, 26, 715–730, 2004.
50. F. Shafizadeh, Introduction to pyrolysis of biomass, in: Specialists workshop
on fast pyrolysis of biomass, pp. 79–103, 1980.
51. A. Ramadhas, S. Jayaraj, K. Lakshmi Narayana Rao, Experimental investiga-
tion on non edible vegetable oil operation in diesel engine for improved per-
formance, in: National conference on advances in mechanical engineering,
JNTU, Anantapur, India, pp. 2, 2002.
52. A. Ramadhas, S. Jayaraj, C. Muraleedharan, Use of vegetable oils as IC engine
fuels—a review, Renewable Energy, 29, 727–742, 2004.
53. P. Verma, M. Sharma, Review of process parameters for biodiesel produc-
tion from different feedstocks, Renewable and Sustainable Energy Reviews,
62, 1063–1071, 2016.
54. I. Atadashi, M. Aroua, A.A. Aziz, N. Sulaiman, The effects of catalysts in bio-
diesel production: A review, Journal of Industrial and Engineering Chemistry,
19, 14–26, 2013.
55. A.S. Ramadhas, S. Jayaraj, C. Muraleedharan, Biodiesel production from
high FFA rubber seed oil, Fuel, 84, 335–340, 2005.
56. Y. Sharma, B. Singh, S. Upadhyay, Advancements in development and char-
acterization of biodiesel: a review, Fuel, 87, 2355–2373, 2008.
57. S. Chongkhong, C. Tongurai, P. Chetpattananondh, Continuous esterifica-
tion for biodiesel production from palm fatty acid distillate using economi-
cal process, Renewable Energy, 34, 1059–1063, 2009.
58. S. Induri, S. Sengupta, J.K. Basu, A kinetic approach to the esterification of
maleic anhydride with methanol on HY zeolite, Journal of Industrial and
Engineering Chemistry, 16, 467–473, 2010.
59. H. Fukuda, A. Kondo, H. Noda, Biodiesel fuel production by transesterifica-
tion of oils, Journal of Bioscience and Bioengineering, 92, 405–416, 2001.
60. A. Demirbas, Biofuels securing the planet’s future energy needs, Energy
Conversion and Management, 50, 2239–2249, 2009.
61. J. Jitputti, B. Kitiyanan, P. Rangsunvigit, K. Bunyakiat, L. Attanatho, P.
Jenvanitpanjakul, Transesterification of crude palm kernel oil and crude
coconut oil by different solid catalysts, Chemical Engineering Journal, 116,
61–66, 2006.
62. J.-S. Lee, S. Saka, Biodiesel production by heterogeneous catalysts and super-
critical technologies, Bioresource Technology, 101, 7191–7200, 2010.
63. F. Chai, F. Cao, F. Zhai, Y. Chen, X. Wang, Z. Su, Transesterification of vege-
table oil to biodiesel using a heteropolyacid solid catalyst, Advanced Synthesis
& Catalysis, 349, 1057–1065, 2007.
64. S. Praserthdam, P. Wongmaneenil, B. Jongsomjit, Investigation of different
modifiers for nanocrystal zirconia on W/ZrO2 catalysts via esterification,
Journal of Industrial and Engineering Chemistry, 16, 935–940, 2010.
Biodiesel Production Methods and Feedstocks 463

65. Z. Helwani, M. Othman, N. Aziz, W. Fernando, J. Kim, Technologies for pro-


duction of biodiesel focusing on green catalytic techniques: a review, Fuel
Processing Technology, 90, 1502–1514, 2009.
66. A.A. Kiss, A.C. Dimian, G. Rothenberg, Solid Acid Catalysts for Biodiesel
Production–‐Towards Sustainable Energy, Advanced Synthesis & Catalysis,
348, 75–81, 2006.
67. O. İlgen, A.N. Akin, Development of alumina supported alkaline catalysts
used for biodiesel production, Turkish Journal of Chemistry, 33, 281–287,
2009.
68. W.M. Antunes, C. de Oliveira Veloso, C.A. Henriques, Transesterification
of soybean oil with methanol catalyzed by basic solids, Catalysis Today, 133,
548–554, 2008.
69. X. Liu, H. He, Y. Wang, S. Zhu, X. Piao, Transesterification of soybean oil to
biodiesel using CaO as a solid base catalyst, Fuel, 87, 216–221, 2008.
70. A. Salis, M. Pinna, M. Monduzzi, V. Solinas, Comparison among immo-
bilised lipases on macroporous polypropylene toward biodiesel synthesis,
Journal of Molecular Catalysis B: Enzymatic, 54, 19–26, 2008.
71. T. Tan, J. Lu, K. Nie, L. Deng, F. Wang, Biodiesel production with immobi-
lized lipase: a review, Biotechnology Advances, 28, 628–634, 2010.
72. K. Bozbas, Biodiesel as an alternative motor fuel: Production and policies in
the European Union, Renewable and Sustainable Energy Reviews, 12, 542–
552, 2008.
73. A.V.L. Pizarro, E.Y. Park, Lipase-catalyzed production of biodiesel fuel
from vegetable oils contained in waste activated bleaching earth, Process
Biochemistry, 38, 1077–1082, 2003.
74. S.E. Lumor, K.C. Jones, R. Ashby, G.D. Strahan, B.H. Kim, G.-C. Lee, J.-F.
Shaw, S.E. Kays, S.-W. Chang, T.A. Foglia, Synthesis and characterization of
canola oil− stearic acid-based trans-free structured lipids for possible mar-
garine application, Journal of Agricultural and Food Chemistry, 55, 10692–
10702, 2007.
75. D. Casanave, J.-L. Duplan, E. Freund, Diesel fuels from biomass, Pure and
Applied Chemistry, 79, 2071–2081, 2007.
76. I. Atadashi, M.K. Aroua, A.A. Aziz, High quality biodiesel and its diesel
engine application: a review, Renewable and Sustainable Energy Reviews, 14,
1999–2008, 2010.
77. D.Y. Leung, X. Wu, M. Leung, A review on biodiesel production using cata-
lyzed transesterification, Applied energy, 87, 1083–1095, 2010.
78. M. Silva, A. Cardoso, Assessing the Activity of Solid-Suported SnCl2 Catalysts
on the Oleic Acid Esterification for Biodiesel Production, J Thermodyn Catal,
7, 2, 2016.
79. Nahadi; Janet John, Atadashi; Musa Idris, The Effects of Variation of Catalyst
Concentration on Biodiesel Production, in: International Journal of Research
- Granthaalayah, pp. 487–496, 2018.
464 Biodiesel Technology and Applications

80. G.K. Ayetor, A. Sunnu, J. Parbey, Effect of biodiesel production parameters


on viscosity and yield of methyl esters: Jatropha curcas, Elaeis guineensis and
Cocos nucifera, Alexandria Engineering Journal, 54, 1285–1290, 2015.
81. E. Sadatshojaei, F. Esmaeilzadeh, J. Fathikaljahi, S.E.H. Barzi, D.A. Wood,
Regeneration of the midrex reformer catalysts using supercritical carbon
dioxide, Chemical Engineering Journal, 343, 748–758, 2018.
82. E. Sadatshojaei, D.A. Wood, Applications Of Supercritical Fluids In
Environmental Remediation, in: Green Sustainable Process for Chemical
and Environmental Engineering and Science: Environmental Remediation,
Elsevier, 2020.
83. S. Lebedevas, A. Vaicekauskas, Research into the application of biodiesel in
the transport sector of Lithuania, Transport, 21, 80–87, 2006.
84. D. Borgman, Agriculture, bio-fuels and striving for greater energy indepen-
dence, John Deere White Paper, 4 2007.
85. B. Barnwal, M. Sharma, Prospects of biodiesel production from vegetable
oils in India, Renewable and Sustainable Energy Reviews, 9, 363–378, 2005.
86. S.K. Thangavelu, A.S. Ahmed, F.N. Ani, Impact of metals on corrosive behav-
ior of biodiesel–Diesel–Ethanol (BDE) alternative fuel, Renewable Energy,
94, 1–9, 2016.
87. E. Torsner, Solving corrosion problems in biofuels industry, Corrosion engi-
neering, Science and Technology, 45, 42–48, 2010.
88. A.T. Hoang, M. Tabatabaei, M. Aghbashlo, A review of the effect of biodiesel
on the corrosion behavior of metals/alloys in diesel engines, Energy Sources,
Part A: Recovery, Utilization, and Environmental Effects, 1–21, 2019.
89. M. Meira, P.M.B. Santana, A.S. Araújo, C.L. Silva, J.R. Leal Filho, H.T.
Ferreira, Oxidative degradation and corrosiveness of biodiesel, Corrosion
Reviews, 32, 143–161, 2014.
90. Z. Fang, Biodiesel: Feedstocks, Production and Applications, BoD–Books on
Demand, 2012.
18
Application of Nanoparticles for the
Enhanced Production of Biodiesel
Muhammad Hilman Mustapha, Akhsan Kamil Azizi, Wan Nur Aini Wan
Mokhtar and Mohamad Azuwa Mohamed*

Department of Chemical Sciences, Faculty of Science and Technology, Universiti


Kebangsaan Malaysia, UKM Bangi, Selangor, Malaysia

Abstract
Nanocatalysis has spotted as a promptly growing field in nanotechnological
advancement. Involving the application of nanoscale materials (nanoparticles,
NPs) or nanomaterials (NMs) as catalysts, they can be found in a variety char-
acter of homogeneous and heterogeneous catalysis applications. Their exqui-
site physical, chemical, and biological properties are gaining interest in diverse
fields, including in the production of biodiesel by transesterification reaction. To
improve the biodiesel production rate, nanocatalysts and NMs are bound with a
metal oxide or microbial enzymes to enhance the modification of triglycerides
into fatty acid methyl ester (FAME) yield. Thus, this chapter provides an insight
of NPs in the synthesis of biodiesel, beginning with a general introduction, then
recent progress of NPs in transesterification application, and, finally, conclusion.

Keywords: Catalyst, nanoparticles, nanocatalyst, biodiesel

18.1 Introduction
The typical transesterification reaction was intensively relied on hetero-
geneous catalysts to increase the production of biodiesel. Heterogeneous
catalysts were usually emphasized metal oxide-based, carbon-based, boron
group-based, waste material-based, or enzyme-based as agents to be per-
formed in the transesterification reaction. However, these non-nanoscale

*Corresponding author: mazuwa@ukm.edu.my

Inamuddin, Mohd Imran Ahamed, Rajender Boddula, and Mashallah Rezakazemi (eds.) Biodiesel
Technology and Applications, (465–480) © 2021 Scrivener Publishing LLC

465
466 Biodiesel Technology and Applications

Nanoparticles/Nanomaterial catalysts

Characteristics Types of Nanomaterial


• Nanoscale size
• High surface area
• High catalytic activity Carbon-based Inorganic-based Organic-based Composite-based
• Natural and synthetic Fullerenes • Metal Organic matter Combination of
• Prone to agglomeration Carbon nanotubes • Metal oxides Dendrimers, carbon-based, metal-
Carbon nanofibers • semiconductors micelles, based, organic-based
Carbon black liposomes, with any form of
Graphene polymers metal, ceramic or
Carbon onions polymer bulk
materials

Figure 18.1 Characteristics and type of nanoparticles catalysts.

catalysts showed some downside as they were ineffective, mass transfer-


resistant, and leaching effect. Thus, nanocatalysts have been introduced to
resolve the difficulties mentioned above. The nanoscale of particle diam-
eter in the range of 1 to 100 nm allows a large surface area–to-volume
ratio and thus allows plenty active sites on the catalyst surface to be partic-
ipated in the reaction. Furthermore, different morphologies of nanoparti-
cles (NPs) vary their applications to be matched with other precursors and
fasten the transesterification reaction rate compared to the larger size of
heterogeneous catalysts.
In general, nanocatalysts can be organized into four categories, which
are carbon-based nanomaterials (NMs), inorganic-based NMs, organ-
ic-based NMs, and composite-based NMs (Figure 18.1). Typically, NPs
can be derived in many forms, such as powder, crystal, and cluster. NP
in powder usually forms from mixtures of a fine powder, while nano-
crystals described as mixtures of ultrafine particles [1]. Then, the cluster
form of NP is categorized as they have a narrow size distribution with a
range of 1–10 nm and one dimension only. Recently, NP including solid
NP, nanobioparticles, and magnetic NP is usually become favor candidates
for numerous researches in catalysis applications due to superhigh surface
area compared to other materials. Commonly, nanocatalyst describes as
heterogeneous nanocatalyst in the transesterification process for biodiesel
production.

18.2 Solid Nanoparticles


Metal NP (MNP) is one of the solid NPs used as nanocatalyst by many
researchers in enhancing the yield of biodiesel. The processes of MNP are
NPs for the Enhanced Production of Biodiesel 467

energy-dependent, performed mostly in laboratories by using either phys-


ical or chemical method. Both methods give different condition where
physical methods usually require huge capital costs but less impact to
environment, while chemical approach are cost effective due to less energy
requirement. The synthesis method of MNP consists of five steps [2].
Firstly, the clustering and complexation steps, the formation of nuclei with
uniform or non-uniform size distribution from the nucleation process.
The rise of nuclei from the NP and then the diffusional growth process
will take place. Moreover, there are two properties that can be governed by
using the nanotechnology that is size of NP and also molecular structure.
In recent years, the unique physical and chemical properties of heter-
ogenous nanocatalyst has risen the interest for its application in catalyst
for biodiesel production. MNPs in single catalyst such as nano-K2O, nano-
CaO, nano-MgO, nano-TiO2, and nano-ZnO are conventional nanocat-
alyst used that shown potential in produce high fatty acid methyl ester
(FAME) yield. Among others, CaO is the most metal oxide used in the
transesterification process of triglyceride due to high availability, cheap,
highly active, non-corrosive, and environmentally friendly [3, 4]. The pro-
duction of Nano-CaO can be from various materials, either natural such
as eggshells and animal bone or through a synthetic process with calcium
nitrate, calcium sulfate, etc. The synthesis of nano-CaO can be obtained
using two main steps, which are preparation and also activation of NP
[5]. Various preparation steps include (dry or wet) impregnation, mixing,
thermal decomposition, and precipitation, whereas for another step, which
is the most critical step to activate the active site of the catalyst by calci-
nation of catalysts with suitable temperature. By performing this critical
step, the specific surface area of the nano-CaO catalyst can be enhanced
or increased.
Various types of nanocatalyst can be produced by neat, doped, and
loaded. By neat-catalyst, pure metal is used. Nano-CaO derived from
calcination an eggshell assist by sonochemical can convert the algae oil
efficiently to biodiesel for 96.3% yield with optimum conditions 9:1 ratio
methanol/oil, 1.25% amount of catalyst, and 55°C reaction temperature
[6]. Another report shows that nano-CaO nanopowder was efficiently con-
verted soybean oil to biodiesel for 96.6% with methanol/oil molar ratio of
7:1, 3.0 wt.% amount of catalyst used, reaction temperature of 338 K, and
reaction time of 60 minutes [7].
Based on the research by Reddy et al. (2006), a very high conversion
of 99% biodiesel are obtained via the usage of nanocrystalline CaO as the
catalyst in the transesterification of soybean oil and poultry fat [8]. This
468 Biodiesel Technology and Applications

high conversion are due to high activity and surface reactivity of nanocrys-
talline CaO caused by high surface or volume ratio of nano-sized catalyst
particles.
Other than nano-CaO, nano-TiO2, and nano-ZnO are the most exten-
sively employed in NMs application. There are plenty synthetic approaches
available to produce TiO2 and ZnO NP. The synthetic approaches are
such as sol-gel processing, reverse microemulsion, dialysis hydrolysis,
microwave-assisted emulsion polymerization, alcohol-thermal method,
hydrothermal, chemical vapor, combustion, and gas-phase methods.
Between current available approaches, hydrothermal process is recognized
as promising method to prepare TiO2 and ZnO, where at a lower tempera-
ture condition (100°C), the reaction can be executed other than another
available approaches that requires higher temperature condition.
Bimetallic NP has shown dominant technological applications in het-
erogeneous catalysis. They possessed special properties, which lead better
reactivity, this is due to the core metal particle could adjust the lattice
strain of the metal, which emerge in a shift of the electronic band struc-
ture of the metal. Nano–CaO-K2CO3 catalyst has produced to achieved
97.67% ± 1.7 yield by using methanol to oil ratio of 9/1 and performed
in reaction time of 8 hours in a liquid batch reactor at 64.85°C [9]. The
NP of CaO-K2CO3 with approximately high surface-to-volume ratios in
which particle sizes range between 20 and 160 nm acts as catalyst sup-
ports. Moreover, high-temperature decomposition products of K2CO3
are deployed to functionalize and enhance the basicity of nanocatalyst.
Calcium methoxide was recorded develop at the surface of CaO through
the K2O phase or Ca–O–K groups and attacks the carbonyl group of the
triglycerides. A power law kinetics equation was used to calculate the
transesterification reaction rate for this reaction. In addition, this catalyst
can be reusable for five runs.
The underlying strength of CaO-K2CO3 reduced slightly upon reuse
compare to the fresh catalyst. As for basicity of the catalyst, it did not
decline further with the increasing number of catalyst reusability. The
transesterification reaction activation energy was determined as 25.34 ±
2.48 kJ/mol [9]. Therefore, the calculated activation energy of the reaction
was low and shows that this reaction was mass transfer controlled, which
lead on how to enhance the reaction rate. More than that, Chinese tallow
oil was converting to biodiesel with a 96.8% yield by KF/CaO nanocatalyst.
The formation of KCaF3 was formed and enhanced catalytic ability [10].
The doped Li toward MgO had enhanced biodiesel yield from soybean
oil which used Li/Mg catalyst with molar ratio of 0.08 and undergo calci-
nation process at 549.85°C, methanol/oil of 1:12, 9 wt% amount of catalyst,
NPs for the Enhanced Production of Biodiesel 469

and yield of 93.9% [11]. The production of strong base sites due to Li doped
influences the biodiesel yield. Nano-λ-Al2O3 as supported with doped KF
by wetness-impregnation method for the transesterification of canola oil
achieved for the yield of 97.7% with calcination temperature of 773 K, 15
wt% loading catalyst, 8-hour reaction time, 338-K reaction temperature,
3 wt% catalyst, and 15:1 molar ratio of methanol/oil [12]. The TiO2-ZnO
nanocatalyst shows an excellent catalytic activity compared to the single
ZnO nanocatalyst with 92.2% conversion of FAME from palm oil with 6:1
methanol to oil molar ratio, 200-mg catalyst loading, 5-hour reaction time,
and 60°C reaction temperature [13].
The ferric manganese–doped sulfated zirconia NP solid catalysts were
developed by impregnation reaction, which are purposely for the trans-
esterification of waste cooking oil. The optimum conditions for the nano-
catalyst for transesterification reaction were 160°C of temperature, 1:10 M
ratio of oil:alcohol, and 3 wt% catalyst dosage for 97.2% yield of biodiesel
[14]. The high conversion of biodiesel was caused by high acidity, large sur-
face area, and pore diameter provided from zirconia as nanocatalyst, which
results rapid diffusion of reactant within the internal pore of catalyst and
contact to the active sites. The bimetallic gold-silver core-shell NP, Au @
Ag NPs, was used as nanocatalyst to convert sunflower oil to biodiesel with
reaction condition of 65°C for temperature and 2 hours for reaction time,
while 5% for catalyst concentration and 5:1 for methanol or oil molar ratio
[15]. The highest yield of biodiesel of 86.9% with three cycles of transester-
ification reaction maintained as a results of greater concentration of highly
reactive sites affected by increased surface-to-volume ratio.
The sulfonated acid on single-walled carbon nanohorn (SWCNH) as a
solid acid catalyst was dissolved with Fe/Fe2O3-mixed NP was performed
for esterification of palmitic acid and transesterification of tributyrin to
produce methyl butyrate [16]. Tributyrin and palmitic acid are an exam-
ple of triglycerides and free fatty acid (FFA) to generate transesterifica-
tion and esterification process. As a results of SWCNH properties such as
high surface area, chemical resistivity, thermal stability, and its nature of
hydrophobic, made it is used as catalyst support. Other than that, for ester-
ification of palmitic acid, it was stated that sulfonated acid on SWCNH
have greater performance of catalytic on the reaction than another types
of sulfonated carbon-based catalyst developed from other sources. More
than that, the use of Fe of Fe2O3 was compared to other metal because it
was found that a catalyst that consists of kind of iron oxide (magnetite)
can simulate immense activity toward transesterification and esterification
reaction. In addition, iron oxide also possesses a good trait for carboxylic
acids and FFA high adsorption.
470 Biodiesel Technology and Applications

The catalytic reaction mechanism for the transesterification and esteri-


fication by using the acid catalyst and the reaction pathway in this mech-
anism is shown in Figures 18.2 and 18.3 [16]. Other than that, by one of
Lewis acid or Brønsted acid, the carbocation formation was stimulates by
the reaction of FFA or triglycerides with the acidic site, and results tetra-
hedral intermediate produces from nucleophilic attack of methanol. After
H2O or glycerin is eliminated, FAME can be produced while the acidic side
can revert to the original structure. This mechanism is genuine to −SO3H
group and a few types of solid acid catalysts [17–19]. Moreover, it can be
expected from this reaction, on the surface of Fe/F2O3 NP, the concentra-
tion of FFA molecules would be higher due to its interaction with the car-
boxyl group of FFAs, thus leads to increasing reaction catalyzed at –SO3H
due to the diffusion of concentrated FFAs molecules to –SO3H.
Other than that, the reaction will also lead to the adsorption of palmitic
acid on Fe/Fe2O3 NP on the catalyst surface due to the existence of hydro-
philic Fe/Fe2O3 surfaces that would attract the carboxylic part of palmitic
acid. While, concurrent with reactions of palmitic acid and tributyrin,
the hydrophobic alkyl parts of the adsorbed palmitic acid will attract the
hydrophobic tributyrin. There is a lot of research regarding production of
biodiesel that uses ferric mixture which is ferric magnetite (Fe3O4), Fe-Ti
mixed oxide, and ferric alginate as a solid catalyst [19, 20, 21]. Fe/Fe2O3
could have the same possibility to acts the same catalytic effect even no

SO3H OR’
O SO3H
Fe Fe2O3
+ ö C+
SO3H
SO3H

SO3H R OR’ R
SO3H

O+
R’ = alkyl group for +O CH3OH
triglyceride or H of FFA
R OR’
R OR’ OR’ H
ö C O+
R CH3

SO3H +
OHR’
Fe Fe2O3 +
O C O ö C O
SO3H

SO3H R CH3 R CH3


+ R’ OH
O O di-, mono glyceride
C Methyl ester or water from FFA
CH3
R

Figure 18.2 General reaction for possible mechanism of esterification and


transesterification with methanol on SO3H-Fe/Fe2O3-SWCNH. Reproduced with
permission [16]. Copyright 2015 Elsevier.
NPs for the Enhanced Production of Biodiesel 471

Tributyrin Alkyl group


molecule
Hydrophilic part
Alkyl group
Palmtic acid
molecule Hydrophilic part Diffusion of reactants
From bulk-liquid into pores on catalyst
methanol

acidic site (-SO3H) SWCNH walls

Diffusion of palmitic acid


from adsorbing zones
on Fe/Fe2O3 particles Diffusion of tributyrin
to acidic site (-SO3H) from concentrating zones
on palmitic acid layer to
acidic site (-SO3H)

Fe/Fe2O3 particles

Figure 18.3 Schematic image of diffusion of reactants in pores in SO3H-Fe/Fe2O3-


SWCNH. Reproduced with permission [16]. Copyright 2015 Elsevier.

researcher report on the catalyst. A magnetic field can recover the mag-
netic properties on SO3H-Fe/Fe2O3- SWCNH. In addition, based on the
research conducted, the esterification process to produce methyl palmitate
can achieve 90% yield or higher in a period of 3 to 7 hour [16].

18.3 Nanobioparticles/Nanobiocatalyst
Over the conventional process, the enzymatic process has several advan-
tages on separating byproduct of glycerol without any complicated oper-
ation steps [22]. There are many research has been reported using the
enzymatic process to enhanced biodiesel production. The use of micro-
organisms such as bacteria, yeast, and fungi as known as a biocatalyst
to improve the biodiesel yield with low-cost material compared to other
materials in biocatalyst such as lipase [23–25]. Lipase catalyst with immo-
bilized material shows the excellent conversion rate of biodiesel, easy sepa-
ration, and reusable catalyst. However, there are limitations in using lipase
due to its high cost that inhibited for wide applications in industrial [22].
In nanotechnology, the nanobiocatalyst (NBC) had not been explored
widely by researchers. NBC is known as promising innovation that
472 Biodiesel Technology and Applications

deployed advance nanotechnology with biotechnology and provides


attracting benefits such as enhancing enzyme activity, capability and sta-
bility, and also engineering performances in bioprocessing applications
[26]. It is also constructed by using functional NMs as enzyme carriers or
storage for immobilizing enzymes. In addition, for biocatalytic reactions,
nanoscale components are identified as ideal substrate materials for the
immobilization of enzymes used for the reactions.
Many researchers have modified the enzyme with the nanosupport or
known as enzyme immobilization, to enhance the biodiesel yield. The
nanostructured scaffolds are used for enzyme immobilization and it can
necessarily increase biocatalyst life cycles, thus decreasing cost needed
for biocatalytic process. This comprises nanofibers, NPs, nanotubes, and
nanocomposites [27–30]. In enzyme immobilization, the NMs possess
several advantages, which are the consistently high surface area for assists
in enzyme-loading capacity, lower-resistance against mass transfer, high
mechanical resistance, easy separation, activity enhancement, and enzyme
stability.
Most of the biocatalyst use immobilization enzymes, such as magnetic
NP. Other than that, immobilization of protein, peptide, and enzyme also
extensively used nanosized magnetic particles for recent development.
The magnetic NP as supports for enzyme immobilization has several ben-
efits such as for the binding efficiency, its offers high specific surface area.
Other than that, less contaminate and low mass transfer resistance, easy
segregation of immobilized enzymes by an external magnetic force, then
reduce operational expense, and a continuous biocatalysts system applica-
tion also provided by magnetic NP as support for enzyme immobilization
[31]. Moreover, the covalent coupling technique was used for immobi-
lization of enzyme onto different supports. The activated magnetic car-
ries and lipase could efficiently bound by a covalent bond and produce
better recovery activity. More than that, nanoparticles, like Fe2O3, Fe3O4,
polymeric, and silica NP, have been introduced for the immobilization
of lipases.
Moreover, Dang and Chin et al. (2012) reported lipase yielded from
an isolated strain Burkholderia sp. C20 was immobilized on magnetic NP
(Fe3O4) for biodiesel production [32]. The immobilize lipase used in pro-
duction of biodiesel from olive oil with 90% yield within 30 hours of reac-
tion time in batch operation with 11 wt% of immobilized lipases. These
immobilized lipases can be reused up to ten cycles without significant loss
in the transesterification reaction. Another research reported the use of
immobilized lipase on Fe3O4 nanocatalyst for synthesis of biodiesel from
soybean oil [33]. For preparation of magnetic Fe3O4 NP, it was done by
NPs for the Enhanced Production of Biodiesel 473

coprecipitating Fe2+ (FeSO4·7H2O) and Fe3+ (FeCl3·6H2O) ions under hydro-


thermal conditions in ammonia solution. The biodiesel yield reached over
90% by three-step addition of methanol when 60% immobilized lipase.
The reusability of catalyst is up to four times without a major decline in its
catalytic activity.

18.4 Magnetic Nanoparticles


Nanomagnetic catalyst has increase attraction and identified as promising
for biodiesel application in recent years. This nanomagnetic catalyst is con-
sidered as promising due to its properties that can made it easily separated
from the reagents by an external magnetic field, which can lead to preven-
tion of catalyst loss and enhance its recovery rate during the separation
process.
As a matter of fact, magnetic materials have arisen as one of promising
alternatives to conventional heterogeneous supports over recent years. The
magnetic catalysts have the advantage over the non-magnetic catalysts in
the separation process. The excellent improvements had made on filtration
and purification process which magnetic separation had been introduced
to replace standard filtration and purification methods which it can avoid
the loss of catalysts and increase its reusability. The different types of fer-
rites such as Fe2O3 and Fe3O4 have been used as a magnetic catalyst or cat-
alyst support and between many types of different ferrites, and nanoscale
magnetite (Fe3O4) is a center of attention in this industry due to its stability,
large surface area, and super paramagnetic properties.
Liu et al. (2016) stated that the catalyst of MgFe2O4 coated by CaO using
a simple alkali precipitation method [34]. The MgFe2O4/CaO nanocatalyst
used for biodiesel synthesis by using soybean oil with 1.0 wt.% catalyst
dosage, 12:1 of methanol/oil ratio, 70°C reaction temperature, and 3 hours
of reaction time. The modify catalyst with magnetic properties shows a
better catalytic activity and better performance in resistance toward water
and acid in the reaction in comparison with single CaO catalyst. These
catalysts contain superparamagnetic behavior that could achieve a yield of
biodiesel 98.3% and reuse for five times with no remarks for any catalytic
activity decline. This shows that MgFe2O4 NP had excellent stability and
recyclability.
Another magnetic NP of CaO catalyst by chemical precipitation is CaO-
Fe3O4 [35]. This magnetic nanocatalyst contains CaO and Fe3O4 accom-
panied by CaFe2O4 for the transesterification of palm seed oil. About
69.7% yield of biodiesel was determined under the reaction conditions of
474 Biodiesel Technology and Applications

300-minute reaction time, 65°C temperature, 20 methanol/oil molar ratio,


and 10 wt.% of CaO-Fe3O4 catalyst loading.
Tang et al. (2012) reported that the catalyst of Ca/Al/Fe3O4 NP
obtained via chemical synthesis method provide high yield of conversion
of biodiesel which is 98.71%. The reaction was catalyzed by calcium alu-
minate and performed under reaction conditions as follows: 15:1 molar
ratio of methanol/oil, 6 wt.% catalyst dosage of rapeseed oil, 65°C reac-
tion temperature, the stirring rate of 270 rpm, and 3-hour reaction time
[21]. This catalyst can stand up to five cycles of catalysis due to improve-
ment toward the stability of catalyst by magnetic NP as supported by
the catalyst.
Moreover, a reactant can contact more effectively on the high surface
of supported magnetic NP and shows high dispersion of magnetic NP–
supported solid catalyst. Due to this, the outcome of the reaction shows
high catalytic activity and also the benefits of efficient separation and good
term in reusability.
The two-step process of esterification of oleic acid and transesterifica-
tion of glyceryl trioleate as model reactions for biodiesel production by
sulfamic acid and sulfonic acid functionalized silica-coated crystalline Fe/
Fe3O4 core/shell MNPs has been reported [36]. The magnetic nanocat-
alyst was prepared by thermal decomposition and the high reactivity of
95% conversion of biodiesel achieved throughout five cycles of reactions.
The MNPs are used as solid supports to produce nanocatalyst that pos-
sess magnetic properties for ease the separation process. Furthermore, due
to the high surface area–to-volume ratio of the MNPs, a large number of
active site can be hold on the catalyst surface. Moreover, the active catalyst
on the surface freely available to the reactant due to the nanocatalyst is
highly soluble in solvents. The mechanism of the related reaction is shown
in Figure 18.4.
Other than that, the cost of biodiesel production can be reduce by reus-
ability and efficient separation of the catalysts. From the nanocatalyst,
the acid-functionalized MNP shows high magnetic moments due to the
existence of the iron (0) core. This magnetic benefit can be seen during
the transesterification process. It occurs when the reaction is completed
and the stirring process has stopped, and the catalyst was gathered on the
magnetic stir and eases the separation between the product and catalyst.
After that, the used catalyst that had been collected from previous reaction
was used for new transesterification reaction. Both catalysts can be reused
for five times, but sulfamic acid-functionalized MNP shows exceptional
reusability and stability which results retaining over 95% yield of biodiesel
compared to sulfonic acid functionalized MNP.
NPs for the Enhanced Production of Biodiesel 475

SH
SO3H
HS
SO3H
HO3S O Si
O
O Si O O SH
Si Si
HO3S SO3H O O O O O Si O
O O O Si
Si O
Fe3O4 O Si O O Si Si O O Si
SO3H HS O O Si O O
O Si O O Si O O O Si O O
O Si O Si
HO3S Si O O Fe3O4 O O
O
Fe 1) H2O2 O O Si O Fe O Si O
Si
Si O O O O O
2) H2SO4 O O Si
Si Si O O O O Si
O O
O Si O O O Si O
Si SH
HO3S O
O O O Si Si Si O O Si O
SO3H
SiO2 Si
O Si O O OO O O O O
HS O Si O Si Si O Si
HO3S O O O
SO3H Si

SH
HO3S SO3H SO3H
MPTMS SH

O
O O O O
1) 1-octadecene O O Si
Si Si O
O
OHOH
oleylamine/HDA x HCI HOHO OH O Si O O O Si O
HO Fe O4 OH TEOS O
Si O
O O Si O
120ºC - 180 ºC 3
NH4OH O Fe O
3 4 O O
Fe(CO)5 HO Fe OH O Si O O Si O
Fe
2) controlled HO
OH O O O O
Si O
OH
oxidation HO
HO OH
Si
OO O
O
O Si
O
OH Si O O O
O O Si Si Si O
O OO O O

APTES
NH2
NHSO2H
HO2SHN NH2
NHSO2H
O Si
O
O Si O O NH2
APTES Si Si
Si O
HO2SHN NHSO2H O O O O OO O
O Si O O Si Si Si O O Si O
HO2SHN
Fe3O4 NHSO2H H2N O O Si O O
O Si
O Si O O Si O O O Si O Si O
O Si
Si O O O Fe3O4 O O O
CISO2H O
Fe O
Si
O Si O Fe O Si
O
Si
HO2SHN O O O O
O Si OO
Si Si O O Si
O O OO O Si
O Si
O
NH2
HO2SHN O Si O O
NHSO2H O O O Si Si Si O Si O
Si O O OO O O O O
SiO2 O Si Si Si
H2N O O Si O Si
O O O
HO2SHN NHSO2H Si
NH2
HO2SHN NHSO2H
NH2

Figure 18.4 Preparation of sulfonic acid functionalized MNPs and sulfamic acid
functionalized MNPs. Reproduced with permission [36]. Copyright 2015 American
Chemical Society.

18.5 How Nanoparticles Enhanced Biodiesel


Production?
NPs can enhance biodiesel production by acting as an enhancer to increase
biodiesel production during the transesterification process. For solid NPs,
there is much research regarding solid NPs as an enhancer in biodiesel
476 Biodiesel Technology and Applications

production. As stated by Badnore et al. [37], commercial calcium oxide has


a lower specific surface area which is 1.54 times lower than nanocrystalline
CaO which make its less favorable than nanocrystalline CaO to use as cat-
alyst. Due to its larger surface area, the nano-sized catalyst showed a higher
conversion rate that is 18.81% than commercial calcium oxide at the same
experiment condition. The high surface area of nanocrystalline CaO is due
to its small particle size and crystallite of the catalyst itself. Therefore, reac-
tion rate is increasing due to large surface area of nanocrystalline CaO.
Moreover, based on Banković–Ilić et al. [40], the research shows great
extensive information regarding the production of biodiesel by the appli-
cation of nano-sized CaO. Nano-sized CaO showed good reusability and
its ability to restrain to leaching makes it as one of good materials as cata-
lyst. The research also stated that the reusability of neat nano-CaO could
affect catalytic activity which it can be used up to three to eight times with-
out showing any decreasing effect. This is due to nano-CaO that consists
of a great number of crystallites with precise edges. When the catalyst was
reused, it can be seen that the aggregates of polycrystals that are less precise
edges was formed [41].
Other than that, Hebbar et al. [38] reported that by using CaO nano-
particle, a huge yield of biodiesel which is 96.2% is produced at moder-
ate reaction conditions. Hsiao et al. [39] also reported that when using
microwave irradiation for biodiesel production, nano-CaO was more effi-
cient than larger sized CaO. Based on the experiment conducted for the
research, high conversion of biodiesel that is 71.6% was obtained using
nano-CaO compared with large-sized CaO that only recorded a 2.3% con-
version rate in the same operating condition.
As for the nano-biocatalyst, the application of nanotechnology in the
biofuel industry and how its enhanced biodiesel production is enzyme
(biocatalysts) immobilization during lipase-catalyzed biodiesel and cellu-
losic ethanol production processes [42]. The advantages that nanostructure
offers for biodiesel production are the large surface area for high enzyme
loading, higher enzymatic stability, and also enzyme reusability, which will
lead to reducing the operational cost of biodiesel production plants [43].
Based on Ahmad and Sardar (2015), NPs can serve as good effective
support components for immobilization of enzyme, by cause of their
excellent properties for stabilizing the key factors that decide the efficiency
of biocatalysts, including specific surface area, mass transfer resistance,
and effective enzyme loading [44]. Other than that, diffusion is always a
problem when dealing with the macromolecular substrates, and NPs are
the ideal candidates for solving the problem [45]. Moreover, the enzyme
that binds with NPs shows the Brownian movement when immersed in
NPs for the Enhanced Production of Biodiesel 477

aqueous solutions, which providing better enzymatic performance than


the unbound enzyme [46].
Besides, for magnetic NPs, its additional hold advantage than other NPs
is that it had easy separation method by just using an external magnetic
field. Furthermore, the enzymes immobilization with the NPs can results in
reduce protein unfolding and performed enhancement in terms of stability
and catalytic performance [46]. Furthermore, Xie and Ma (2009) reported
that for enzyme immobilization, the use of magnetic NPs as supports is
highly recommended due to the following advantages: a higher specific
surface that lead to effective binding, lower mass transfer resistance and
less fouling, easy separation method by using the external magnetic field
thus lowering the operational cost, and also the application of a contin-
uous biocatalysis system [33]. As for the conclusion, NMs were success-
fully proven as one of the enhancers in biodiesel production by providing
high surface area for high catalyst activity and also high enzyme loading,
higher catalytic and enzymatic stability, and also in terms of reusability of
the catalyst.

18.6 Conclusion
The review of NP catalysts for biodiesel production was showed that the
usage of nanocatalysts contributed a better performance of the transes-
terification reaction. Overall, we can conclude that the application of NP
catalysts would be economical, recyclable, and functionable for high or low
FFA feedstock to produce high quality of biodiesel and contributed to the
environment clean.

References
1. Nouailhat A. An Introduction to Nanoscience and Nanotechnology. New York:
Wiley, 2010.
2. M. Aliofkharaei. Nanoparticles Technology. 2015.
3. Kesić Ž., Lukić I., Zdujić M., Mojović Lj., Skala D. Calcium oxide based cata-
lysts for biodiesel production: a review. Chem. Ind. Chem. Eng. Q 22:391–408,
2016.
4. Marinković D.M., Stanković M.V., Veličković A.V., Avramović J.M.,
Miladinović M.R., Stamenković O.S., Veljković V.B., Jovanović D.M. Calcium
oxide as a promising heterogeneous catalyst for biodiesel production: cur-
rent state and perspectives. Renew. Sustain. Energy. Rev. 56:1387–408, 2016.
478 Biodiesel Technology and Applications

5. Ivana B. Banković–Ilić, Marija R. Miladinović, Olivera S. Stamenković, Vlada


B. Veljković. Application of nano CaO–based catalysts in biodiesel synthesis.
Renewable and Sustainable Energy Reviews Volume 72, pp. 746–760, May
2017.
6. S.Siva, C.Marimuthu Production of Biodiesel by Transesterification of
Algae Oil with an assistance of Nano-CaO Catalyst derived from Egg Shell.
International Journal of ChemTech Research. Vol.7, No.4, pp 2112–2116.
7. Ming-Chien Hsiao, Chin-Chiuan Lin, Yung-Hung Chang. Microwave irra-
diation-assisted transesterification of soybean oil to biodiesel catalyzed by
nanopowder calcium oxide. Elsevier. Volume 90, Issue 5, Pages 1963–1967,
May 2011.
8. Reddy C.R.V., Oshel R., Verkade J.G., Room–temperature conversion of soybean
oil and poultry fat to biodiesel catalyzed by nanocrystalline calcium oxides.
Energy Fuel 20:1310–4, 2006.
9. Nebahat Degirmenbasi, Samet Coskun, Nezahat Boz, Dilhan M. Kalyon.
Biodiesel synthesis from canola oil via heterogeneous catalysis using func-
tionalized CaO nanoparticles. Elsevier Fuel 153, 620–627, 2015.
10. Libai Wen, Yun Wang, Donglian Lu, Shengyang Hu, Heyou Han. Preparation
of KF/CaO nanocatalyst and its application in biodiesel production from
Chinese tallow seed oil. Elsevier Fuel 89, 2267–2271, 2010.
11. Zhenzhong Wena, Xinhai Yu, Shan-Tung Tu, Jinyue Yan, Erik Dahlquist
Synthesis of biodiesel from vegetable oil with methanol catalyzed by Li-doped
magnesium oxide catalysts. Applied Energy 87, 743–748, 2010.
12. Nezahat Boz, Nebahat Degirmenbasi, Dilhan M. Kalyon. Conversion of bio-
mass to fuel: Transesterification of vegetable oil to biodiesel using KF loaded
nano-g-Al2O3 as catalyst. Applied Catalysis B: Environmental 89, 590–596,
2009.
13. Rajesh Madhuvilakku, Shakkthivel Piraman. Biodiesel synthesis by TiO2–
ZnO mixed oxide nanocatalyst catalysed palm oil transesterification process.
Bioresource Technology 150, 55–59, 2013.
14. Fatah H. Alhassan, Umer Rashid & Y. H. Taufiq-Yap. Ferric-Manganese
Doped Sulphated Zirconia Nanoparticles Catalyst for Single-Step Biodiesel
Production from Waste Cooking Oil: Characterization and Optimization.
International Journal of Green Energy.
15. Banerjee M., Dey B., Talukdar J., Kalita M.C., Production of biodiesel from
sunflower oil using highly catalytic bimetallic gold-silver core-shell nanopar-
ticle. Energy 69:695–9, 2014.
16. Chantamanee Poonjarernsilp, Noriaki Sano, Hajime Tamon. Simultaneous
esterification and transesterification for biodieselsynthesis by a catalyst con-
sisting of sulfonated single-walled carbonnanohorn dispersed with Fe/Fe2O3
nanoparticles. Applied Catalysis A: General 497, 145–152, 2015.
17. A.F. Rojas Gonzalez, E. Giron Gallego, H.G. Torres Castaneda, Ingenieria E
Investig. 29, 17–22, 2009.
18. S. Yan, S.O. Salley, K.Y. Simon Ng, Appl. Catal. A 353, 203–212, 2009.
NPs for the Enhanced Production of Biodiesel 479

19. B. Peng-Lim, S. Ganesan, G.P. Maniam, M. Khairuddean, Energy 46,


132–139, 2012.
20. K. Nuithitikul, Int. J. Chem. React. Eng., 9, 2011.
21. Shaokun Tang, Liping Wang, Yi Zhang, Shufen Li, Songjiang Tian, Boyang
Wang Study on preparation of Ca/Al/Fe3O4 magnetic composite solid catalyst
and its application in biodiesel transesterification. Fuel Processing Technology
95, 84–89, 2012.
22. H. Fukuda, S. Hama, S. Tamalampudi and H. Noda. Whole-cell biocatalysts
for biodiesel fuel production Volume 26, Issue 12, Pages 668–673, December
2008.
23. Ban, K. et al., Whole-cell biocatalyst for biodiesel fuel production utiliz-
ing Rhizopus oryzae cells immobilized within biomass support particles.
Biochem. Eng. J. 8, 39–43, 2001.
24. Fujita, Y. et al., Direct and efficient production of ethanol fromcellulosic
material with a yeast strain displaying cellulolytic enzymes. Appl. Environ.
Microbiol. 68, 5136–5141, 2002.
25. Narita, J. et al., Display of active enzymes on the cell surface of Escherichia
coli using PgsA anchor protein and their application to bioconversion. Appl.
Microbiol. Biotechnol. 70, 564–572, 2006.
26. Mailin Misson, Hu Zhang and Bo Jin. Nanobiocatalyst advancements and
bioprocessing applications. J. R. Soc. Interface 12: 20140891.
27. Plessis D.M., Botes M., Dicks L.M.T., Cloete T.E., Immobilization of com-
mercial hydrolytic enzymes on poly (acrylonitrile) nanofibers for anti-
biofilm activity. J. Chem. Technol. Biotechnol. 88, 585–593, 2012.
28. Wang L., Jiang R., Reversible His-tagged enzyme immobilization on func-
tionalized carbon nanotubes as nanoscale biocatalyst. Methods Mol. Biol.
743, 95–106, 2011.
29. Johnson P.A., Park H.J., Driscoll A.J., Enzyme nanoparticle fabrication: mag-
netic nanoparticle synthesis and enzyme immobilization. In Enzyme stabili-
zation and immobilization methods in molecular biology (ed. SD Minteer), pp.
183–191, 2011.
30. Tran D.T., Chen C.L., Chang J.S., Immobilization of Burkholderia sp. lipase
on a ferric silica nanocomposite for biodiesel production. J. Biotechnol. 158,
112–119, 2012.
31. Unisha Patel, Kishor Chauhan, and Shilpa Gupte Synthesis, characteri-
zation and application of lipase-conjugated citric acid-coated magnetic
nanoparticles for ester synthesis using waste frying oil. 3 Biotech., Apr; 8(4):
211, 2018.
32. Dang-Thuan Tran, Ching-Lung Chena, Jo-Shu Chang. Immobilization of
Burkholderia sp. lipase on a ferric silica nanocomposite for biodiesel pro-
duction. Journal of Biotechnology 158, 112–119, 2012.
33. Wenlei Xie and Ning Ma. Immobilized Lipase on Fe3O4 Nanoparticles as
Biocatalyst for Biodiesel Production. Energy & Fuels, 23, 1347–1353, 2009.
480 Biodiesel Technology and Applications

34. Yanlei Liu, Pingbo Zhang, Mingming Fan, Pingping Jiang. Biodiesel produc-
tion from soybean oil catalyzed by magnetic nanoparticle MgFe2O4@CaO.
Fuel 164, 314–321, 2016.
35. Mortadha A. Ali1, Imad A. Al-Hydary, Tahseen A. Al-Hattab. Nano-
Magnetic Catalyst CaO-Fe3O4 for Biodiesel Production from Date Palm Seed
Oil. Bulletin of Chemical Reaction Engineering & Catalysis, 12 (3), 460–468,
2017.
36. Hongwang Wang, Jose Covarrubias, Heidy Prock, Xiaorong Wu, Donghai
Wang, and Stefan H. Bossmann. Acid-functionalized Magnetic Nanoparticle
as Heterogeneous Catalysts for Biodiesel Synthesis. J. Phys. Chem. C, 119, 46,
26020–26028, 2015.
37. Badnore, A.U., Jadhav, N.L., Pinjari, D.V., Pandit, A.B.: Efficacy of newly
developed nano-crystalline calcium oxide for biodiesel production. Chem.
Eng. Process 133, 312–319, 2018.
38. Hebbar, H.R.H., Math, M.C., Yatish, K.V.: Optimization and kinetic study
of CaO nano-particles catalysed biodiesel production from Bombax coiba
oil. Energy 143, 25–34, 2018.
39. Hsiao, M.C., Lin, C.C., Chang, Y.H.: Microwave irradiation assisted trans-
esterification of soybean oil to biodiesel catalysed by nanopowder calcium
oxide. Fuel 90, 1963–1967, 2011.
40. Banković-Ilić, I.B., Miladinovic, M.R., Stamenkovic, O.S., Veljkovic, V.B.:
Application of nano CaO-based catalysts in biodiesel synthesis. Renew. Sust.
Energ. Rev. 72, 746–760, 2017.
41. Bharti, P., Singh, B., & Dey, R. K. Process optimization of biodiesel produc-
tion catalyzed by CaO nanocatalyst using response surface methodology.
Journal of Nanostructure in Chemistry, 9(4), 269–280, 2019.
42. Kim, K. H., Lee, O. K., & Lee, E. Y. Nano-immobilized biocatalysts for bio-
diesel production from renewable and sustainable resources. Catalysts, 8(2),
68, 2018.
43. Santos, C. R., Tonoli, C. C., Trindade, D. M., Betzel, C., Takata, H., Kuriki,
T., ... & Murakami, M. T. Structural basis for branching‐enzyme activity
of glycoside hydrolase family 57: Structure and stability studies of a novel
branching enzyme from the hyperthermophilic archaeon Thermococcus
Kodakaraensis KOD1. Proteins: Structure, Function, and Bioinformatics,
79(2), 547–557, 2011.
44. Ahmad, R., & Sardar, M. Enzyme immobilization: an overview on nanopar-
ticles as immobilization matrix. Biochemistry and Analytical Biochemistry,
4(2), 1, 2015.
45. Hwang E.T., Gu MB Enzyme stabilization by nano/microsized hybrid mate-
rials. Engineering in Life Sciences 13: 49–61(2013).
46. Gupta M.N., Kaloti M., Kapoor M., Solanki K. Nanomaterials as matrices for
enzyme immobilization. Artif. Cells Blood Substit. Immobil. Biotechnol. 39:
98–109, 2011.
Index

1,2-Propanediol, 137, 139, 141, Biodiesel, 3–8, 10, 11, 14, 86, 217, 342,
148–149, 159 362, 447–448, 451–452, 454–480
manufacturing, 137, 139–140
Acid catalyzed transesterification, 92 technology, 138
Acidic catalyst, 250–252 Biodiesel feedstock, 87, 242–248, 252
Acrylic resin, 22, 23, 32 Biodiesel production, 60–62
Activated carbon (AC), 111, 114–115 Biodiesel production from non-edible
Acyl acceptor, 4, 6, 9–15, 28–34 and waste lipid sources,
Acylglycerol, 19 algae, 406
Africa renewable energy initiative, 288 animal fat, 404
Alcohol, 2 azadirachta indica, neem oil, 398
Algae, 197, 203 calophyllum inophyllum, 397
Algal biomass, 450–451 castor oil, 402
Algal oil, 298, 299, 303 deacidification, 414
Alkali catalyzed transesterification, 94 jatropha, 394
Alkaline catalyst, 250 larva, 411
Alkaline earth metal oxides, 106–107 transesterification reaction, 392
Alkyl ester, 5, 7, 9, 167, 174, 180, 192 waste cooking oil, 405
Alternative fluid, 215 Bioethanol, 2, 3
Aluminum oxide (Alumina), 98, 111 Biofuel classification, 449
Amberlyst-15, 105–106 Biofuels, 2, 3, 38, 195, 196, 362
Anaerobic digestion, 195, 198, 201 Biogas, 3
Animal fat, 241, 242, 244 Biohydrogen, 3
Asthma, 2 Biomass support particles (BSPs), 23,
30
Bacterial lipases, 21 Biomechanical conversion, 195, 198,
Basicity, 468 201, 207
Batch reactor, 342–343 Biomethanol, 3
Benzene, 13, 27 Bioreactor, 10, 35, 37
Bimetallic, 468–469 Biorefinery design, 219
Biobutanol, 3 Blending, 173, 184, 189, 223
Biocatalysts, 8, 23, 30, 33, 38, 115, 232 Blending of plant-derived oils, 451
Biodegradability, 197 Box Behnken technique, 179

481
482 Index

Bubble column reactor (BCR), 342, Diacylglycerol (DAG), 7


347–349 Diglycerides, 5, 18, 19, 27
By-products, 169, 173 Dilution, 249
Dimethyl carbonate (DMC), 12, 14, 15,
Calcination, 142, 146, 149 24, 26
Carbon emissions, 169 Direct combustion, 195, 198, 201, 207
Carbon neutral biofuels, 289 Direct use of plant-derived, 451
Carboxylation, 141, 146
Catalysis, 7, 8, 17–19, 26, 27 EBD resins, 106
Catalyst, 137, 139–159, 453–455, 462–463 Edible oil, 243, 244, 250, 251, 253
heterogeneous, 364 Effect of initial water concentration,
homogeneous, 364 155
Catalyst selection, 139, 151 Encapsulation of enzyme, 116
Catalyst support, 111 Energy, 217
Catalytic, Energy generation, 168
dehydration of glycerol, 143 Environmental protection, 169
transformation, 137, 152, 159 Environment-friendly, 167
Catalytic methods, 89 Enzymatic, 471, 476–477
Catalytic transesterification, 174, 175, Enzymatic catalyst, 250, 252
190, 191 Enzymatic transesterification, 8, 14,
Cation exchange resins, 105–106 16, 17, 35, 38
Cavitation, 353 Enzyme immobilization, 115–116
Celite, 22 Enzymes, 146
Central composite design, 179, 190, 192 Epoxy acrylic, 22
Chromatography, 20 Esterification, 6, 7, 11, 19, 27, 365
Clays, 384 Ethanol, 10, 11, 14, 22, 24, 27, 30
Climate change, 138 Ethyl ester, 12, 17
Cloud point, 180
Comparative analysis, 21 Fatty acid profile, 245–247
Continuous stirred tank reactor Fatty acids, 6, 17–19
(CSTR), 342, 344–345 Fatty acids alkylesters (FAAEs), 3, 6, 7
Conversion of glycerol, 137, 139–141, Fatty acids methylesters (FAMEs), 3,
146–147, 151–152, 154–156, 11–14, 19, 25–27, 29, 31, 34, 465,
158–159 467, 469
Copenhagen accord, 288 Feed stock, 8, 9, 38
Corrosion, 457, 464 Feedstock sources, 215
Cost, 241, 243, 244, 251, 253, 256, 257 Fermentation, 195, 198, 201
Covalent bond, 472 First-generation biofuels, 288–289, 499
Covalently immobilization of enzymes, Fischer Tropsch, 196, 200, 202
116 Fixed bed reactor (FBR), 342,
Cracking, 174 346–347
Cross-linked enzyme aggregates Flash point, 170, 180, 183
(CLEA), 117–118 Fossil fuels, 2, 4, 38, 60
Crude oil refining industry, 147 Free fatty acids, 5
Index 483

Fuel security, 448 Immiscibility, 65–66


Fuel supply, 448–449 Impregnation, 467, 469
In situ method, 256
Gas chromatography, 183, Industrialization, 137–138
193 Influential parameters affecting
Gasification, 195, 198–200, 202 biodiesel production, 454
Genetic engineering/techniques Infrared spectroscopy, 182, 185, 193,
enhancing biodiesel production, 194
433 Intensification techniques,
GLC selectivity, 146 co-solvent method, 256
Global warming, 285–288, 312 microwave heating, 253, 254
Glyceric acid, 141–142, 159 supercritical alcohol, 253
Glycerol, 7–19, 33, 35, 137, 139–159 ultrasonic irradiation, 254
Gold, 140, 142–143, 151 International monetary fund, 287
Growth, 2, 5, 6, 20, 21
Jatropha, 203, 205
Helium, 137, 152 Jathropa oil, 22, 29–30
Heterogeneous, 465–466, 468, 473
Heterogeneous catalysts, 64–65, Kinetic, 468
252 Kyoto protocol, 288
Heterogeneous ultrasonic reaction,
72–78 Lactic acid, 148
Heterogenous catalyst, Leaching, 466
acid catalyst, 97 Lignocellulosic biomasses, 202, 289
base catalyst, 106 Lipases, 8, 9, 11, 13–15, 17–28, 31–35,
Heteropolyacids,101, 382 471–473, 476
High yield per hectre, 218 Lipozyme, 35, 37
High-performance liquid
chromatography, 184, 193 Magnetic, 466, 471–474, 477
Homogeneous catalyst, 63, 64, 250, Magnetic nanoparticles (MNPs), 111,
251 114, 116
acid catalyst, 89 Markets and economic considerations,
base catalyst, 93 455
Homogeneous ultrasonic reaction, Mechanism, 137, 157–159
69–72 Membrane reactor (MR), 256, 343,
Humanity, 237 350–352
Hybrid catalytic plasma reactor, 343, Membrane technology, 198, 202, 207
350 Mesoporous silica, 99–100, 112
Hydrogenolysis, 137, 139, 141, Metal oxides, 380, 465, 467
147–149, 151–154, 156–159 Methanol, 10–16, 22–24, 26
Hydrophobic, 469–470 Methanolysis, 11, 13, 15
Hydrothermal, 468, 473 Methodology involved in the
Hydrothermal liquefaction, 198, extraction of algae,
200 chemical solvent extraction, 439
484 Index

extraction by supercritical carbon Methyl acetate, 12, 15


dioxide, 439 Methyl ester, 9, 12, 13, 16, 17
extraction involving direct Microalgae, 5, 11, 12, 38
transesterification, 440 Microbial oil, 241, 242
extraction using biochemical Microchannel reactor (MCR), 342,
techniques, 439 350–351
extraction using transesterification Microemulsification, 173, 249, 292
techniques, 440 Microemulsion with excess methanol
Methods for biodiesel production, and ethanol, 451
269 Microtube microreactor (MM), 350,
dilution with hydrocarbons 352
blending, 269 Microwave, 33–35
micro-emulsion, 269–270 Microwave reactor, 255, 352
pyrolysis (thermal cracking), Millennium development goals,
270–271 288
transesterification (alcoholysis), 271 Mixed metal oxides, 110
conventional transesterification, Mutagenesis, 27
272–273
catalytic-based process, 273 Nafion NR50, 105
enzymatic-catalyzed Nano,
transesterification, 276 biocatalyst, 471
heterogeneous-catalyzed catalysis, 465–469, 472–474, 478,
transesterification, 480
275–276 materials, 465–466
homogeneous particles, 465–466, 472–473,
alkaline-catalyzed 475
transesterification, structure, 476
274–275 Nanotechnology in microalgae
homogeneous-acid- biodiesel, 434
catalyzed-esterification, Nature of feedstocks, 454
273–274 Non-edible oil, 241, 243, 244, 250, 251,
non-catalytic-based process, 253
276 Non-edible vegetable oils,
BIOX co-solvent process, jatropha, 296, 297, 301
277–278 karanja, 296, 297
supercritical alcohol process, linum usitassimum, 296
276–227 Non-toxic renewable fuels, 169
in-situ transesterification (reactive Novozyme, 11–15, 31, 32, 36
extraction), 271–272 Nuclear magnetic resonance
microwave/ultrasound-assisted spectroscopy, 184–185
transesterification, 278 Nucleophilic, 18, 19
variables affecting
transesterification reaction, Optimization, 137, 139, 151
278–281 Optimization techniques, 167
Index 485

Oscillatory flow reactor (OFR), 353 first-generation feedstock, 268


Oxalic acid, 141 second-generation feedstock,
268–269
Palladium, 140, 142–143, 151–152 third-generation feedstock, 269
Paramagnetic, 473 Serine, 18, 19
Parameters, 139–140, 153, 155 Silica gel, 11, 13
Paris agreement, 287, 288 Silver, 140
Penicillium, 28, 31 Sludge palm oil (SPO), 220
Pentapeptide, 18 SO3H as catalyst, 379
Petroleum, 137–138, 147, 149 Solvent, 149–151
Photobiological, 198, 201 Sonochemical reactor, 353
Physicochemical properties of Soybean oil, 22, 24, 27, 34
biodiesel from microalgae, Specific examples of biodiesel
431 production from microalgae,
Platinum, 140, 142, 145, 151–152 434
Polyurathane, 23, 27, 30 Spectrum, 148
Pour point, 180 Static mixers, 257
Production technologies, 248 Sulfated zirconia, 97
Psychrophilic Lipases, 26 Surface area, 466–467, 469, 472–474,
Pyrolysis, 224, 248, 368, 447, 452–453, 476–477
458, 462 Synthesis gas, 202
Pyrolysis of glycerol, 144
Tartronic acid, 141–142
Rancimat method, 185 Temperature, 137, 140, 144–148,
Reaction mechanism, 137, 157–159 150–151, 153–155, 157–159
Reaction parameters, 78–79 Thermal cracking, 292, 452
Reaction temperature, 454 Thermochemical conversion, 195, 198,
Reactive distillation, 257 199, 202, 207
Reactive distillation column (RDC), Third-generation biofuels, 288, 290,
343, 349 450, 460
Reactive distillation technology, 198, Transesterification, 6–17, 19, 21,
202, 207 25–28, 33, 38, 139, 145–146,
Reactor, 342, 369 195, 198, 199, 201, 202,
Renewable and sustainable energy, 249–251, 287, 292–300, 303,
138 365, 367, 447, 452–454, 458,
Renewable energy technologies, 138 462–463, 469–470, 472–475,
Reusability, 468, 473–474, 476–477 477
Triglyceride, 5, 6, 17–19, 31
Scientific, 138 Triolein, 22, 29, 32, 33
Second-generation biodiesel, 221 Tungstated zirconia, 99
Second-generation biofuels, 289, 290, Two step process, 251
449
Selection of feedstock for biodiesel, Ultrafiltration, 20, 21
267 Ultrasonicator, 255
486 Index

Ultrasonic, 33, 34 Waste biomass, 101


Ultrasonic energy, 66–68 Waste cooking oil, 295, 297, 298,
Ultrasonication, 33–35 301
Ultrasound, 35 Waste oil, 241, 242
Ultrasound technology, 198, 201, 202, 207 Water hyacinth, 290

Valorization, 137, 139–140 Zeolite, 378


Vegetable oil, 242, 248, 249 Zeolites, 102, 105
Also of Interest

Check out these other forthcoming and published titles


from Scrivener Publishing

Books by the same editor from Wiley-Scrivener


Applied Water Science Volume 1: Fundamentals and Applications, Edited by
Inamuddin, Mohd Imran Ahamed, Rajender Boddula and Tauseef Ahmad
Rangreez, ISBN 9781119724766. Edited by one of the most well-respected
and prolific engineers in the world and his team, this is the first volume in a
two-volume set that is the most thorough, up-to-date, and comprehensive
volume on applied water science available today. COMING IN SUMMER
2021

Applied Water Science Volume 2: Remediation Technologies, Edited by


Inamuddin, Mohd Imran Ahamed, Rajender Boddula and Tauseef Ahmad
Rangreez, ISBN 9781119724735. The second volume in a new two-volume
set on applied water science, this book provides understanding, occur-
rence, identification, toxic effects and control of water pollutants in aquatic
environment using green chemistry protocols. COMING IN SUMMER
2021

Potassium-Ion Batteries: Materials and Applications, Edited by Inamuddin,


Rajender Boddula, and Abdullah M. Asiri, ISBN 9781119661399. Edited
by one of the most well-respected and prolific engineers in the world and
his team, this is the most thorough, up-to-date, and comprehensive vol-
ume on potassium-ion batteries available today. NOW AVAILABLE!

Rechargeable Batteries: History, Progress, and Applications, edited by


Rajender Boddula, Inamuddin, Ramyakrishna Pothu, and Abdullah M.
Asiri, ISBN 9781119661191. Edited by one of the most well-respected and
prolific engineers in the world and his team, this is the most thorough,
up-to-date, and comprehensive volume on rechargeable batteries available
today. NOW AVAILABLE!
Other books on this subject from Wiley-Scrivener
Reverse Osmosis: Design, Applications, and Applications for Engineers 2nd
Edition, by Jane Kucera, ISBN 9781118639740. This is the most compre-
hensive and up-to-date coverage of the “green” process of reverse osmosis
in industrial applications, completely updated in this new edition to cover
all of the processes and equipment necessary to design, operate, and trou-
bleshoot reverse osmosis systems. NOW AVAILABLE!

Desalination: Water from Water 2nd Edition, by Jane Kucera, ISBN


9781119407744. This all-new revised edition of a modern classic is the
most comprehensive and up-to-date coverage of the “green” process of
desalination in industrial and municipal applications, covering all of the
processes and equipment necessary to design, operate, and troubleshoot
desalination systems. This is becoming increasingly more important for
not only our world’s industries, but our world’s populations, as pure water
becomes more and more scarce. NOW AVAILABLE!

Sustainable Water Purification, by Safiur Rahman and M. R. Islam, ISBN


9781119650997. This is the only book that takes a zero-waste approach
to propose 100% sustainable water purification techniques. NOW
AVAILABLE!

Electrochemical Water Processing, by Ralph Zito, ISBN 9781118098714.


Two of the most important issues facing society today and in the future
will be the global water supply and energy production. This book addresses
both of these important issues with the idea that non-usable water can
be purified through the use of electrical energy, instead of chemical or
mechanical methods. NOW AVAILABLE!

You might also like